quantum orders and symmetric spin liquids - massachusetts

36
Quantum Orders and Symmetric Spin Liquids Xiao-Gang Wen * Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139 (Dated: Aug., 2001) A concept – quantum order – is introduced to describe a new kind of orders that generally appear in quantum states at zero temperature. Quantum orders that characterize universality classes of quantum states (described by complex ground state wave-functions) is much richer then classical orders that characterize universality classes of finite temperature classical states (described by positive probability distribution functions). The Landau’s theory for orders and phase transitions does not apply to quantum orders since they cannot be described by broken symmetries and the associated order parameters. We introduced a mathematical object – projective symmetry group – to characterize quantum orders. With the help of quantum orders and projective symmetry groups, we construct hundreds of symmetric spin liquids, which have SU (2), U (1) or Z 2 gauge structures at low energies. We found that various spin liquids can be divided into four classes: (a) Rigid spin liquid – spinons (and all other excitations) are fully gaped and may have bosonic, fermionic, or fractional statistics. (b) Fermi spin liquid – spinons are gapless and are described by a Fermi liquid theory. (c) Algebraic spin liquid – spinons are gapless, but they are not described by free fermionic/bosonic quasiparticles. (d) Bose spin liquid – low lying gapless excitations are described by a free boson theory. The stability of those spin liquids are discussed in details. We find that stable 2D spin liquids exist in the first three classes (a–c). Those stable spin liquids occupy a finite region in phase space and represent quantum phases. Remarkably, some of the stable quantum phases support gapless excitations even without any spontaneous symmetry breaking. In particular, the gapless excitations in algebraic spin liquids interact down to zero energy and the interaction does not open any energy gap. We propose that it is the quantum orders (instead of symmetries) that protect the gapless excitations and make algebraic spin liquids and Fermi spin liquids stable. Since high Tc superconductors are likely to be described by a gapless spin liquid, the quantum orders and their projective symmetry group descriptions lay the foundation for spin liquid approach to high T c superconductors. PACS numbers: 73.43.Nq, 74.25.-q, 11.15.Ex Contents I. Introduction 2 A. Topological orders and quantum orders 2 B. Spin-liquid approach to high T c superconductors 3 C. Spin-charge separation in (doped) spin liquids 4 D. Organization 6 II. Projective construction of 2D spin liquids – a review of SU(2) slave-boson approach 6 III. Spin liquids from translationally invariant ansatz 9 IV. Quantum orders in symmetric spin liquids 12 A. Quantum orders and projective symmetry groups 12 B. Classification of symmetric Z 2 spin liquids 14 C. Classification of symmetric U (1) and SU (2) spin liquids 15 * URL: http://dao.mit.edu/~wen V. Continuous transitions and spinon spectra in symmetric spin liquids 17 A. Continuous phase transitions without symmetry breaking 17 B. Symmetric spin liquids around the U (1)-linear spin liquid U1Cn01n 17 C. Symmetric spin liquids around the SU (2)-gapless spin liquid SU2An0 20 D. Symmetric spin liquids around the SU (2)-linear spin liquid SU2Bn0 22 VI. Mean-field phase diagram of J 1 -J 2 model 23 VII. Physical measurements of quantum orders 24 VIII. Four classes of spin liquids and their stability 26 A. Rigid spin liquid 26 B. Bose spin liquid 26 C. Fermi spin liquid 27 D. Algebraic spin liquid 27 E. Quantum order and the stability of spin liquids 29 IX. Relation to previously constructed spin liquids 30

Upload: others

Post on 09-Feb-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Quantum Orders and Symmetric Spin Liquids - Massachusetts

Quantum Orders and Symmetric Spin Liquids

Xiao-Gang Wen∗Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139

(Dated: Aug., 2001)

A concept – quantum order – is introduced to describe a new kind of orders that generallyappear in quantum states at zero temperature. Quantum orders that characterize universalityclasses of quantum states (described by complex ground state wave-functions) is much richer thenclassical orders that characterize universality classes of finite temperature classical states (describedby positive probability distribution functions). The Landau’s theory for orders and phase transitionsdoes not apply to quantum orders since they cannot be described by broken symmetries and theassociated order parameters. We introduced a mathematical object – projective symmetry group –to characterize quantum orders. With the help of quantum orders and projective symmetry groups,we construct hundreds of symmetric spin liquids, which have SU(2), U(1) or Z2 gauge structuresat low energies. We found that various spin liquids can be divided into four classes: (a) Rigidspin liquid – spinons (and all other excitations) are fully gaped and may have bosonic, fermionic,or fractional statistics. (b) Fermi spin liquid – spinons are gapless and are described by a Fermiliquid theory. (c) Algebraic spin liquid – spinons are gapless, but they are not described by freefermionic/bosonic quasiparticles. (d) Bose spin liquid – low lying gapless excitations are describedby a free boson theory. The stability of those spin liquids are discussed in details. We find that stable2D spin liquids exist in the first three classes (a–c). Those stable spin liquids occupy a finite regionin phase space and represent quantum phases. Remarkably, some of the stable quantum phasessupport gapless excitations even without any spontaneous symmetry breaking. In particular, thegapless excitations in algebraic spin liquids interact down to zero energy and the interaction doesnot open any energy gap. We propose that it is the quantum orders (instead of symmetries) thatprotect the gapless excitations and make algebraic spin liquids and Fermi spin liquids stable. Sincehigh Tc superconductors are likely to be described by a gapless spin liquid, the quantum orders andtheir projective symmetry group descriptions lay the foundation for spin liquid approach to high Tc

superconductors.

PACS numbers: 73.43.Nq, 74.25.-q, 11.15.Ex

Contents

I. Introduction 2A. Topological orders and quantum orders 2B. Spin-liquid approach to high Tc

superconductors 3C. Spin-charge separation in (doped) spin

liquids 4D. Organization 6

II. Projective construction of 2D spin liquids– a review of SU(2) slave-boson approach 6

III. Spin liquids from translationally invariantansatz 9

IV. Quantum orders in symmetric spinliquids 12A. Quantum orders and projective symmetry

groups 12B. Classification of symmetric Z2 spin liquids 14C. Classification of symmetric U(1) and SU(2)

spin liquids 15

∗URL: http://dao.mit.edu/~wen

V. Continuous transitions and spinonspectra in symmetric spin liquids 17A. Continuous phase transitions without

symmetry breaking 17B. Symmetric spin liquids around the

U(1)-linear spin liquid U1Cn01n 17C. Symmetric spin liquids around the

SU(2)-gapless spin liquid SU2An0 20D. Symmetric spin liquids around the

SU(2)-linear spin liquid SU2Bn0 22

VI. Mean-field phase diagram of J1-J2 model 23

VII. Physical measurements of quantumorders 24

VIII. Four classes of spin liquids and theirstability 26A. Rigid spin liquid 26B. Bose spin liquid 26C. Fermi spin liquid 27D. Algebraic spin liquid 27E. Quantum order and the stability of spin

liquids 29

IX. Relation to previously constructed spinliquids 30

Page 2: Quantum Orders and Symmetric Spin Liquids - Massachusetts

2

X. Summary of the main results 31

A. General conditions on projectivesymmetry groups 33

References 35

I. INTRODUCTION

Due to its long length, we would like to first outlinethe structure of the paper so readers can choose to readthe parts of interests. The section X summarize the mainresults of the paper, which also serves as a guide of thewhole paper. The concept of quantum order is introducedin section I A. A concrete mathematical description ofquantum order is described in section IV A and sectionIV B. Readers who are interested in the background andmotivation of quantum orders may choose to read sec-tion I A. Readers who are familiar with the slave-bosonapproach and just want a quick introduction to quantumorders may choose to read sections IV A and IV B. Read-ers who are not familiar with the slave-boson approachmay find the review sections II and III useful. Read-ers who do not care about the slave-boson approach butare interested in application to high Tc superconductorsand experimental measurements of quantum orders maychoose to read sections I A, I B, VII and Fig. 1 - Fig. 15,to gain some intuitive picture of spinon dispersion andneutron scattering behavior of various spin liquids.

A. Topological orders and quantum orders

Matter can have many different states, such as gas,liquid, and solid. Understanding states of matter is thefirst step in understanding matter. Physicists find mattercan have much more different states than just gas, liquid,and solid. Even solids and liquids can appear in manydifferent forms and states. With so many different statesof matter, a general theory is needed to gain a deeperunderstanding of states of matter.

All the states of matter are distinguished by their in-ternal structures or orders. The key step in developingthe general theory for states of matter is the realizationthat all the orders are associated with symmetries (orrather, the breaking of symmetries). Based on the rela-tion between orders and symmetries, Landau developed ageneral theory of orders and the transitions between dif-ferent orders.[1, 2] Landau’s theory is so successful andone starts to have a feeling that we understand, at inprinciple, all kinds of orders that matter can have.

However, nature never stops to surprise us. In 1982,Tsui, Stormer, and Gossard[3] discovered a new kind ofstate – Fractional Quantum Hall (FQH) liquid.[4] Quan-tum Hall liquids have many amazing properties. A quan-tum Hall liquid is more “rigid” than a solid (a crystal),

in the sense that a quantum Hall liquid cannot be com-pressed. Thus a quantum Hall liquid has a fixed and well-defined density. When we measure the electron densityin terms of filling factor ν, we found that all discoveredquantum Hall states have such densities that the fillingfactors are exactly given by some rational numbers, suchas ν = 1, 1/3, 2/3, 2/5, .... Knowing that FQH liquidsexist only at certain magical filling factors, one cannothelp to guess that FQH liquids should have some inter-nal orders or “patterns”. Different magical filling fac-tors should be due to those different internal “patterns”.However, the hypothesis of internal “patterns” appearsto have one difficulty – FQH states are liquids, and howcan liquids have any internal “patterns”?

In 1989, it was realized that the internal orders inFQH liquids (as well as the internal orders in chiral spinliquids[5, 6]) are different from any other known ordersand cannot be observed and characterized in any con-ventional ways.[7, 8] What is really new (and strange)about the orders in chiral spin liquids and FQH liquidsis that they are not associated with any symmetries (orthe breaking of symmetries), and cannot be described byLandau’s theory using physical order parameters.[9] Thiskind of order is called topological order. Topological orderis a new concept and a whole new theory was developedto describe it.[9, 10]

Knowing FQH liquids contain a new kind of order –topological order, we would like to ask why FQH liquidsare so special. What is missed in Landau’s theory forstates of matter so that the theory fails to capture thetopological order in FQH liquids?

When we talk about orders in FQH liquids, we arereally talking about the internal structure of FQH liq-uids at zero temperature. In other words, we are talkingabout the internal structure of the quantum ground stateof FQH systems. So the topological order is a propertyof ground state wave-function. The Landau’s theory isdeveloped for system at finite temperatures where quan-tum effects can be ignored. Thus one should not be sur-prised that the Landau’s theory does not apply to statesat zero temperature where quantum effects are impor-tant. The very existence of topological orders suggeststhat finite-temperature orders and zero-temperature or-ders are different, and zero-temperature orders containricher structures. We see that what is missed by Landau’stheory is simply the quantum effect. Thus FQH liquidsare not that special. The Landau’s theory and symmetrycharacterization can fail for any quantum states at zerotemperature. As a consequence, new kind of orders withno broken symmetries and local order parameters (suchas topological orders) can exist for any quantum states atzero temperature. Because the orders in quantum statesat zero temperature and the orders in classical states atfinite temperatures are very different, here we would liketo introduce two concepts to stress their differences:[11](A) Quantum orders:[93] which describe the universal-ity classes of quantum ground states (ie the universalityclasses of complex ground state wave-functions with in-

Page 3: Quantum Orders and Symmetric Spin Liquids - Massachusetts

3

finity variables);(B)Classical orders: which describe the universalityclasses of classical statistical states (ie the universalityclasses of positive probability distribution functions withinfinity variables).From the above definition, it is clear that the quantumorders associated with complex functions are richer thanthe classical orders associated with positive functions.The Landau’s theory is a theory for classical orders,which suggests that classical orders may be characterizedby broken symmetries and local order parameters.[94]The existence of topological order indicates that quan-tum orders cannot be completely characterized by bro-ken symmetries and order parameters. Thus we need todevelop a new theory to describe quantum orders.

In a sense, the classical world described by positiveprobabilities is a world with only “black and white”. TheLandau’s theory and the symmetry principle for classicalorders are color blind which can only describe different“shades of grey” in the classical world. The quantumworld described by complex wave functions is a “colorful”world. We need to use new theories, such as the theory oftopological order and the theory developed in this paper,to describe the rich “color” of quantum world.

The quantum orders in FQH liquids have a specialproperty that all excitations above ground state have fi-nite energy gaps. This kind of quantum orders are calledtopological orders. In general, a topological order is de-fined as a quantum order where all the excitations aboveground state have finite energy gapes.

Topological orders and quantum orders are generalproperties of any states at zero temperature. Non trivialtopological orders not only appear in FQH liquids, theyalso appear in spin liquids at zero temperature. In fact,the concept of topological order was first introduced in astudy of spin liquids.[9] FQH liquid is not even the firstexperimentally observed state with non trivial topolog-ical orders. That honor goes to superconducting statediscovered in 1911.[12] In contrast to a common pointof view, a superconducting state cannot be characterizedby broken symmetries. It contains non trivial topologicalorders,[13] and is fundamentally different from a super-fluid state.

After a long introduction, now we can state the mainsubject of this paper. In this paper, we will study a newclass of quantum orders where the excitations above theground state are gapless. We believe that the gaplessquantum orders are important in understanding high Tc

superconductors. To connect to high Tc superconduc-tors, we will study quantum orders in quantum spin liq-uids on a 2D square lattice. We will concentrate on howto characterize and classify quantum spin liquids withdifferent quantum orders. We introduce projective sym-metry groups to help us to achieve this goal. The projec-tive symmetry group can be viewed as a generalizationof symmetry group that characterize different classicalorders.

B. Spin-liquid approach to high Tc superconductors

There are many different approaches to the high Tc

superconductors. Different people have different pointsof view on what are the key experimental facts for thehigh Tc superconductors. The different choice of the keyexperimental facts lead to many different approaches andtheories. The spin liquid approach is based on a point ofview that the high Tc superconductors are doped Mottinsulators.[14–16] (Here by Mott insulator we mean a in-sulator with an odd number of electron per unit cell.) Webelieve that the most important properties of the high Tc

superconductors is that the materials are insulators whenthe conduction band is half filled. The charge gap ob-tained by the optical conductance experiments is about2eV , which is much larger than the anti-ferromagnetic(AF) transition temperature TAF ∼ 250K, the super-conducting transition temperature Tc ∼ 100K, and thespin pseudo-gap scale ∆ ∼ 40meV.[17–19] The insulat-ing property is completely due to the strong correlationspresent in the high Tc materials. Thus the strong cor-relations are expect to play very important role in un-derstanding high Tc superconductors. Many importantproperties of high Tc superconductors can be directlylinked to the Mott insulator at half filling, such as (a)the low charge density[20] and superfluid density,[21] (b)Tc being proportional to doping Tc ∝ x,[22–24] (c) thepositive charge carried by the charge carrier,[20] etc .

In the spin liquid approach, the strategy is to try tounderstand the properties of the high Tc superconduc-tors from the low doping limit. We first study the spinliquid state at half filling and try to understand the par-ent Mott insulator. (In this paper, by spin liquid, wemean a spin state with translation and spin rotationsymmetry.) At half filling, the charge excitations canbe ignored due to the huge charge gap. Thus we canuse a pure spin model to describe the half filled system.After understand the spin liquid, we try to understandthe dynamics of a few doped holes in the spin liquidstates and to obtain the properties of the high Tc su-perconductors at low doping. One advantage of the spinliquid approach is that experiments (such as angle re-solved photo-emission,[17, 18, 25, 26] NMR,[27], neutronscattering,[28–30] etc ) suggest that underdoped cuper-ates have many striking and qualitatively new propertieswhich are very different from the well known Fermi liq-uids. It is thus easier to approve or disapprove a newtheory in the underdoped regime by studying those qual-itatively new properties.

Since the properties of the doped holes (such as theirstatistics, spin, effective mass, etc ) are completely de-termined by the spin correlation in the parent spin liq-uids, thus in the spin liquid approach, each possible spinliquid leads to a possible theory for high Tc supercon-ductors. Using the concept of quantum orders, we cansay that possible theories for high Tc superconductors inthe low doping limits are classified by possible quantumorders in spin liquids on 2D square lattice. Thus one

Page 4: Quantum Orders and Symmetric Spin Liquids - Massachusetts

4

way to study high Tc superconductors is to construct allthe possible spin liquids that have the same symmetriesas those observed in high Tc superconductors. Then an-alyze the physical properties of those spin liquids withdopings to see which one actually describes the high Tc

superconductor. Although we cannot say that we haveconstructed all the symmetric spin liquids, in this paperwe have found a way to construct a large class of symmet-ric spin liquids. (Here by symmetric spin liquids we meanspin liquids with all the lattice symmetries: translation,rotation, parity, and the time reversal symmetries.) Wealso find a way to characterize the quantum orders inthose spin liquids via projective symmetry groups. Thisgives us a global picture of possible high Tc theories.We would like to mention that a particular spin liquid– the staggered-flux/d-wave state[31, 32] – may be im-portant for high Tc superconductors. Such a state canexplain[33, 34] the highly unusual pseudo-gap metallicstate found in underdoped cuperates,[17, 18, 25, 26] aswell as the d-wave superconducting state[32].

The spin liquids constructed in this paper can be di-vided into four class: (a) Rigid spin liquid – spinons arefully gaped and may have bosonic, fermionic, or frac-tional statistics, (b) Fermi spin liquid – spinons are gap-less and are described by a Fermi liquid theory, (c) Alge-braic spin liquid – spinons are gapless, but they are notdescribed by free fermionic/bosonic quasiparticles. (d)Bose spin liquid – low lying gapless excitations are de-scribed by a free boson theory. We find some of the con-structed spin liquids are stable and represent stable quan-tum phases, while others are unstable at low energies dueto long range interactions caused by gauge fluctuations.The algebraic spin liquids and Fermi spin liquids are in-teresting since they can be stable despite their gaplessexcitations. Those gapless excitations are not protectedby symmetries. This is particularly striking for algebraicspin liquids since their gapless excitations interact downto zero energy and the states are still stable. We proposethat it is the quantum orders that protect the gaplessexcitations and ensure the stability of the algebraic spinliquids and Fermi spin liquids.

We would like to point out that both stable and unsta-ble spin liquids may be important for understanding highTc superconductors. Although at zero temperature highTc superconductors are always described stable quantumstates, some important states of high Tc superconductors,such as the pseudo-gap metallic state for underdopedsamples, are observed only at finite temperatures. Suchfinite temperature states may correspond to (doped) un-stable spin liquids, such as staggered flux state. Thuseven unstable spin liquids can be useful in understand-ing finite temperature metallic states.

There are many different approach to spin liquids.In addition to the slave-boson approach,[6, 15, 16,31–33, 35–40] spin liquids has been studied usingslave-fermion/σ-model approach,[41–46] quantum dimermodel,[47–51] and various numerical methods.[52–55] Inparticular, the numerical results and recent experimen-

tal results[56] strongly support the existence of quantumspin liquids in some frustrated systems. A 3D quantumorbital liquid was also proposed to exist in LaTiO3.[57]

However, I must point out that there is no generallyaccepted numerical results yet that prove the existenceof spin liquids with odd number of electron per unit cellfor spin-1/2 systems, despite intensive search in last tenyears. But it is my faith that spin liquids exist in spin-1/2 systems. For more general systems, spin liquids doexist. Read and Sachdev[43] found stable spin liquids ina Sp(N) model in large N limit. The spin-1/2 modelstudied in this paper can be easily generalized to SU(N)model with N/2 fermions per site.[31, 58] In the large Nlimit, one can easily construct various Hamiltonians[58,59] whose ground states realize various U(1) and Z2 spinliquids constructed in this paper. The quantum ordersin those large-N spin liquids can be described by themethods introduced in this paper.[58] Thus, despite theuncertainly about the existence of spin-1/2 spin liquids,the methods and the results presented in this paper arenot about (possibly) non-existing “ghost states”. Thosemethods and results apply, at least, to certain large-Nsystems. In short, non-trivial quantum orders exist intheory. We just need to find them in nature. (In fact, ourvacuum is likely to be a state with a non-trivial quantumorder, due to the fact that light exists.[58]) Knowing theexistence of spin liquids in large-N systems, it is not sucha big leap to go one step further to speculate spin liquidsexist for spin-1/2 systems.

C. Spin-charge separation in (doped) spin liquids

Spin-charge separation and the associated gauge the-ory in spin liquids (and in doped spin liquids) are very im-portant concepts in our attempt to understand the prop-erties of high Tc superconductors.[14–16, 39, 60] However,the exact meaning of spin-charge separation is differentfor different researchers. The term “spin-charge separa-tion” has at lease in two different interpretations. In thefirst interpretation, the term means that it is better to in-troduce separate spinons (a neutral spin-1/2 excitation)and holons (a spinless excitation with unit charge) to un-derstand the dynamical properties of high Tc supercon-ductors, instead of using the original electrons. However,there may be long range interaction (possibly, even con-fining interactions at long distance) between the spinonsand holons, and the spinons and holons may not be welldefined quasiparticles. We will call this interpretationpseudo spin-charge separation. The algebraic spin liquidshave the pseudo spin-charge separation. The essence ofthe pseudo spin-charge separation is not that spin andcharge separate. The pseudo spin-charge separation issimply another way to say that the gapless excitationscannot be described by free fermions or bosons. In thesecond interpretation, the term “spin-charge separation”means that there are only short ranged interactions be-tween the spinons and holons. The spinons and holons

Page 5: Quantum Orders and Symmetric Spin Liquids - Massachusetts

5

are well defined quasiparticles at least in the dilute limitor at low energies. We will call the second interpretationthe true spin-charge separation. The rigid spin liquidsand the Fermi spin liquids have the true spin-charge sep-aration.

Electron operator is not a good starting point to de-scribe states with pseudo spin-charge separation or truespin-charge separation. To study those states, we usu-ally rewrite the electron operator as a product of severalother operators. Those operators are called parton oper-ators. (The spinon operator and the holon operator areexamples of parton operators). We then construct mean-field state in the enlarged Hilbert space of partons. Thegauge structure can be determined as the most generaltransformations between the partons that leave the elec-tron operator unchanged.[61] After identifying the gaugestructure, we can project the mean-field state onto thephysical (ie the gauge invariant) Hilbert space and obtaina strongly correlated electron state. This procedure in itsgeneral form is called projective construction. It is a gen-eralization of the slave-boson approach.[15, 16, 33, 36–38, 40] The general projective construction and the re-lated gauge structure has been discussed in detail forquantum Hall states.[61] Now we see a third (but tech-nical) meaning of spin-charge separation: to construct astrongly correlated electron state, we need to use par-tons and projective construction. The resulting effectivetheory naturally contains a gauge structure.

Although, it is not clear which interpretation of spin-charge separation actually applies to high Tc supercon-ductors, the possibility of true spin-charge separation inan electron system is very interesting. The first con-crete example of true spin-charge separation in 2D isgiven by the chiral spin liquid state,[5, 6] where thegauge interaction between the spinons and holons be-comes short-ranged due to a Chern-Simons term. TheChern-Simons term breaks time reversal symmetry andgives the spinons and holons a fractional statistics. Laterin 1991, it was realized that there is another way tomake the gauge interaction short-ranged through theAnderson-Higgs mechanism.[38, 43] This led to a mean-field theory[38, 40] of the short-ranged Resonating Va-lence Bound (RVB) state[47, 48] conjectured earlier. Wewill call such a state Z2 spin liquid state, to stress theunconfined Z2 gauge field that appears in the low energyeffective theory of those spin liquids. (See remarks at theend of this section. We also note that the Z2 spin liquidsstudied in Ref. [43] all break the 90◦ rotation symmetryand are different from the short-ranged RVB state stud-ied Ref. [38, 40, 47, 48].) Since the Z2 gauge fluctuationsare weak and are not confining, the spinons and holonshave only short ranged interactions in the Z2 spin liquidstate. The Z2 spin liquid state also contains a Z2 vortex-like excitation.[38, 62] The spinons and holons can bebosons or fermions depending on if they are bound withthe Z2 vortex.

Recently, the true spin-charge separation, the Z2 gaugestructure and the Z2 vortex excitations were also pro-

posed in a study of quantum disordered superconduct-ing state in a continuum model[63] and in a Z2 slave-boson approach[64]. The resulting liquid state (whichwas named nodal liquid) has all the novel properties ofZ2 spin liquid state such as the Z2 gauge structure andthe Z2 vortex excitation (which was named vison). Fromthe point of view of universality class, the nodal liquid isone kind of Z2 spin liquids. However, the particular Z2

spin liquid studied in Ref. [38, 40] and the nodal liquidare two different Z2 spin liquids, despite they have thesame symmetry. The spinons in the first Z2 spin liquidhave a finite energy gap while the spinons in the nodalliquid are gapless and have a Dirac-like dispersion. In thispaper, we will use the projective construction to obtainmore general spin liquids. We find that one can constructhundreds of different Z2 spin liquids. Some Z2 spin liq-uids have finite energy gaps, while others are gapless.Among those gapless Z2 spin liquids, some have finiteFermi surfaces while others have only Fermi points. Thespinons near the Fermi points can have linear E(k) ∝ |k|or quadratic E(k) ∝ k2 dispersions. We find there aremore than one Z2 spin liquids whose spinons have a mass-less Dirac-like dispersion. Those Z2 spin liquids havethe same symmetry but different quantum orders. Theiransatz are give by Eq. (42), Eq. (39), Eq. (106), etc .

Both chiral spin liquid and Z2 spin liquid states areMott insulators with one electron per unit cell if notdoped. Their internal structures are characterized by anew kind of order – topological order, if they are gappedor if the gapless sector decouples. Topological order isnot related to any symmetries and has no (local) or-der parameters. Thus, the topological order is robustagainst all perturbations that can break any symmetries(including random perturbations that break translationsymmetry).[9, 10] (This point was also emphasized inRef. [65] recently.) Even though there are no order pa-rameters to characterize them, the topological orders canbe characterized by other measurable quantum numbers,such as ground state degeneracy in compact space as pro-posed in Ref. [9, 10]. Recently, Ref. [65] introduced avery clever experiment to test the ground state degen-eracy associated with the non-trivial topological orders.In addition to ground state degeneracy, there are otherpractical ways to detect topological orders. For example,the excitations on top of a topologically ordered statecan be defects of the under lying topological order, whichusually leads to unusual statistics for those excitations.Measuring the statistics of those excitations also allow usto measure topological orders.

The concept of topological order and quantum orderare very important in understanding quantum spin liq-uids (or any other strongly correlated quantum liquids).In this paper we are going to construct hundreds of dif-ferent spin liquids. Those spin liquids all have the samesymmetry. To understand those spin liquids, we need tofirst learn how to characterize those spin liquids. Thosestates break no symmetries and hence have no order pa-rameters. One would get into a wrong track if trying to

Page 6: Quantum Orders and Symmetric Spin Liquids - Massachusetts

6

find an order parameter to characterize the spin liquids.We need to use a completely new way, such as topologicalorders and quantum orders, to characterize those states.

In addition to the above Z2 spin liquids, in this paperwe will also study many other spin liquids with differentlow energy gauge structures, such as U(1) and SU(2)gauge structures. We will use the terms Z2 spin liq-uids, U(1) spin liquids, and SU(2) spin liquids to describethem. We would like to stress that Z2, U(1), and SU(2)here are gauge groups that appear in the low energy ef-fective theories of those spin liquids. They should not beconfused with the Z2, U(1), and SU(2) gauge group inslave-boson approach or other theories of the projectiveconstruction. The latter are high energy gauge groups.The high energy gauge groups have nothing to do withthe low energy gauge groups. A high energy Z2 gaugetheory (or a Z2 slave-boson approach) can have a lowenergy effective theory that contains SU(2), U(1) or Z2

gauge fluctuations. Even the t-J model, which has nogauge structure at lattice scale, can have a low energyeffective theory that contains SU(2), U(1) or Z2 gaugefluctuations. The spin liquids studied in this paper allcontain some kind of low energy gauge fluctuations. De-spite their different low energy gauge groups, all thosespin liquids can be constructed from any one of SU(2),U(1), or Z2 slave-boson approaches. After all, all thoseslave-boson approaches describe the same t-J model andare equivalent to each other. In short, the high energygauge group is related to the way in which we write downthe Hamiltonian, while the low energy gauge group is aproperty of ground state. Thus we should not regard Z2

spin liquids as the spin liquids constructed using Z2 slave-boson approach. A Z2 spin liquid can be constructedfrom the U(1) or SU(2) slave-boson approaches as well.A precise mathematical definition of the low energy gaugegroup will be given in section IV A.

D. Organization

In this paper we will use the method outlined inRef. [38, 40] to study gauge structures in various spinliquid states. In section II we review SU(2) mean-fieldtheory of spin liquids. In section III, we construct sim-ple symmetric spin liquids using translationally invari-ant ansatz. In section IV, projective symmetry group isintroduced to characterize quantum orders in spin liq-uids. In section V, we study the transition between dif-ferent symmetric spin liquids, using the results obtainedin Ref. [66], where we find a way to construct all thesymmetric spin liquids in the neighborhood of some wellknown spin liquids. We also study the spinon spectrumto gain some intuitive understanding on the dynamicalproperties of the spin liquids. Using the relation betweentwo-spinon spectrum and quantum order, we propose, insection VII, a practical way to use neutron scatteringto measure quantum orders. We study the stability ofFermi spin liquids and algebraic spin liquids in section

VIII. We find that both Fermi spin liquids and algebraicspin liquids can exist as zero temperature phases. This isparticularly striking for algebraic spin liquids since theirgapless excitations interacts even at lowest energies andthere are no free fermionic/bosonic quasiparticle excita-tions at low energies. We show how quantum order canprotect gapless excitations. Appendix A contains an al-gebraic description of projective symmetry groups, whichcan be used to classify projective symmetry groups.[66]Section X summarizes the main results of the paper.

II. PROJECTIVE CONSTRUCTION OF 2DSPIN LIQUIDS – A REVIEW OF SU(2)

SLAVE-BOSON APPROACH

In this section, we are going to use projective construc-tion to construct 2D spin liquids. We are going to reviewa particular projective construction, namely the SU(2)slave-boson approach.[15, 16, 33, 36–38, 40] The gaugestructure discovered by Baskaran and Anderson[16] inthe slave-boson approach plays a crucial role in our un-derstanding of strongly correlated spin liquids.

We will concentrate on the spin liquid states of a purespin-1/2 model on a 2D square lattice

Hspin =∑

<ij>

JijSi · Sj + ... (1)

where the summation is over different links (ie 〈ij〉 and〈ji〉 are regarded as the same) and ... represents possi-ble terms which contain three or more spin operators.Those terms are needed in order for many exotic spinliquid states introduced in this paper to become theground state. To obtain the mean-field ground state ofthe spin liquids, we introduce fermionic parton operatorsfiα, α = 1, 2 which carries spin 1/2 and no charge. Thespin operator Si is represented as

Si =12f†iασαβfiβ (2)

In terms of the fermion operators the Hamiltonian Eq. (1)can be rewritten as

H =∑

〈ij〉−1

2Jij

(f†iαfjαf†jβfiβ +

12f†iαfiαf†jβfjβ

)(3)

Here we have used σαβ · σα′β′ = 2δαβ′δα′β − δαβδα′β′ .We also added proper constant terms

∑i f

†iαfiα and∑

〈ij〉 f†iαfiαf†jβfjβ to get the above form. Notice that

the Hilbert space of Eq. (3) is generated by the partonoperators fα and is larger than that of Eq. (1). Theequivalence between Eq. (1) and Eq. (3) is valid only inthe subspace where there is exactly one fermion per site.Therefore to use Eq. (3) to describe the spin state weneed to impose the constraint[15, 16]

f†iαfiα = 1, fiαfiβεαβ = 0 (4)

Page 7: Quantum Orders and Symmetric Spin Liquids - Massachusetts

7

The second constraint is actually a consequence of thefirst one.

A mean-field ground state at “zeroth” order is obtainedby making the following approximations. First we replaceconstraint Eq. (4) by its ground-state average

〈f†iαfiα〉 = 1, 〈fiαfiβεαβ〉 = 0 (5)

Such a constraint can be enforced by including a sitedependent and time independent Lagrangian multiplier:al0(i)(f

†iαfiα − 1), l = 1, 2, 3, in the Hamiltonian. At the

zeroth order we ignore the fluctuations (ie the time de-pendence) of al

0. If we included the fluctuations of al0,

the constraint Eq. (5) would become the original con-straint Eq. (4).[15, 16, 36, 37] Second we replace theoperators f†iαfjβ and fiαfiβ by their ground-state ex-pectations value

ηijεαβ =− 2〈fiα fjβ〉, ηij =ηji

χijδαβ =2〈f†iαfjβ〉, χij =χ†ji (6)

again ignoring their fluctuations. In this way we obtainthe zeroth order mean-field Hamiltonian:

Hmean

=∑

〈ij〉−3

8Jij

[(χjif

†iαfjα + ηijf

†iαf†jβ εαβ + h.c)

−|χij |2 − |ηij |2]

(7)

+∑

i

[a30(f

†iαfiα − 1) + [(a1

0 + ia20)fiαfiβεαβ + h.c.]

]

χij and ηij in Eq. (7) must satisfy the self consistencycondition Eq. (6) and the site dependent fields al

0(i) arechosen such that Eq. (5) is satisfied by the mean-fieldground state. Such χij , ηij and al

0 give us a mean-fieldsolution. The fluctuations in χij , ηij and al

0(i) describethe collective excitations above the mean-field groundstate.

The Hamiltonian Eq. (7) and the constraints Eq. (4)have a local SU(2) symmetry.[36, 37] The local SU(2)symmetry becomes explicit if we introduce doublet

ψ =(

ψ1

ψ2

)=

(f↑f†↓

)(8)

and matrix

Uij =

(χ†ij ηij

η†ij −χij

)= U†

ji (9)

Using Eq. (8) and Eq. (9) we can rewrite Eq. (5) andEq. (7) as:

⟨ψ†iτ

lψi

⟩= 0 (10)

Hmean =∑

〈ij〉

38Jij

[12Tr(U†

ij Uij)− (ψ†i Uijψj + h.c.)]

+∑

i

al0ψ

†iτ

lψi (11)

where τ l, l = 1, 2, 3 are the Pauli matrices. From Eq. (11)we can see clearly that the Hamiltonian is invariant undera local SU(2) transformation W (i):

ψi → W (i) ψi

Uij → W (i) Uij W †(j) (12)

The SU(2) gauge structure is originated from Eq. (2).The SU(2) is the most general transformation betweenthe partons that leave the physical spin operator un-changed. Thus once we write down the parton expres-sion of the spin operator Eq. (2), the gauge structure ofthe theory is determined.[61] (The SU(2) gauge structurediscussed here is a high energy gauge structure.)

We note that both components of ψ carry spin-up.Thus the spin-rotation symmetry is not explicit in ourformalism and it is hard to tell if Eq. (11) describes aspin-rotation invariant state or not. In fact, for a generalUij satisfying Uij = U†

ji, Eq. (11) may not describe aspin-rotation invariant state. However, if Uij has a form

Uij = iρijWij ,

ρij = real number,Wij ∈ SU(2), (13)

then Eq. (11) will describe a spin-rotation invariant state.This is because the above Uij can be rewritten in a formEq. (9). In this case Eq. (11) can be rewritten as Eq. (7)where the spin-rotation invariance is explicit.

To obtain the mean-field theory, we have enlarged theHilbert space. Because of this, the mean-field theoryis not even qualitatively correct. Let |Ψ(Uij)

mean〉 be theground state of the Hamiltonian Eq. (11) with energyE(Uij , al

iτl). It is clear that the mean-field ground state

is not even a valid wave-function for the spin systemsince it may not have one fermion per site. Thus it isvery important to include fluctuations of al

0 to enforceone-fermion-per-site constraint. With this understand-ing, we may obtain a valid wave-function of the spin sys-tem Ψspin({αi}) by projecting the mean-field state to thesubspace of one-fermion-per-site:

Ψspin({αi}) = 〈0|∏

i

fiαi |Ψ(Uij)mean〉. (14)

Now the local SU(2) transformation Eq. (12) can have

a very physical meaning: |Ψ(Uij)mean〉 and |Ψ(W (i)UijW

†(j))mean 〉

give rise to the same spin wave-function after projection:

〈0|∏

i

fiαi |Ψ(Uij)mean〉 = 〈0|

i

fiαi |Ψ(W (i)UijW†(j))

mean 〉 (15)

Thus Uij and U ′ij = W (i)UijW †(j) are just two differ-

ent labels which label the same physical state. Within themean-field theory, a local SU(2) transformation changesa mean-field state |Ψ(Uij)

mean〉 to a different mean-field state

|Ψ(U ′ij)mean〉. If the two mean-field states always have the

Page 8: Quantum Orders and Symmetric Spin Liquids - Massachusetts

8

same physical properties, the system has a local SU(2)symmetry. However, after projection, the physical spinquantum state described by wave-function Ψspin({αi}) isinvariant under the local SU(2) transformation. A localSU(2) transformation just transforms one label, Uij , of aphysical spin state to another label, U ′

ij , which labels theexactly the same physical state. Thus after projection,local SU(2) transformations become gauge transforma-tions. The fact that Uij and U ′

ij label the same physicalspin state creates a interesting situation when we con-sider the fluctuations of Uij around a mean-field solution– some fluctuations of Uij do not change the physicalstate and are unphysical. Those fluctuations are calledthe pure gauge fluctuations.

The above discussion also indicates that in order forthe mean-field theory to make any sense, we must at leastinclude the SU(2) gauge (or other gauge) fluctuationsdescribed by al

0 and Wij in Eq. (13), so that the SU(2)gauge structure of the mean-field theory is revealed andthe physical spin state is obtained. We will include thegauge fluctuations to the zeroth-order mean-field theory.The new theory will be called the first order mean-fieldtheory. It is this first order mean-field theory that rep-resents a proper low energy effective theory of the spinliquid.

Here, we would like make a remark about “gauge sym-metry” and “gauge symmetry breaking”. We see thattwo ansatz Uij and U ′

ij = W (i)UijW †(j) have the samephysical properties. This property is usually called the“gauge symmetry”. However, from the above discussion,we see that the “gauge symmetry” is not a symmetry. Asymmetry is about two different states having the sameproperties. Uij and U ′

ij are just two labels that labelthe same state, and the same state always have the sameproperties. We do not usually call the same state hav-ing the same properties a symmetry. Because the samestate alway have the same properties, the “gauge symme-try” can never be broken. It is very misleading to call theAnderson-Higgs mechanism “gauge symmetry breaking”.With this understanding, we see that a superconductor isfundamentally different from a superfluid. A superfluidis characterized by U(1) symmetry breaking, while a su-perconductor has no symmetry breaking once we includethe dynamical electromagnetic gauge fluctuations. A su-perconductor is actually the first topologically orderedstate observed in experiments,[13] which has no symme-try breaking, no long range order, and no (local) orderparameter. However, when the speed of light c = ∞,a superconductor becomes similar to a superfluid and ischaracterized by U(1) symmetry breaking.

The relation between the mean-field state and thephysical spin wave function Eq. (14) allows us to con-struct transformation of the physical spin wave-functionfrom the mean-field ansatz. For example the mean-fieldstate |Ψ(U ′ij)

mean〉 with U ′ij = Ui−l,j−l produces a phys-

ical spin wave-function which is translated by a dis-tance l from the physical spin wave-function producedby |Ψ(Uij)

mean〉. The physical state is translationally sym-

metric if and only if the translated ansatz U ′ij and the

original ansatz Uij are gauge equivalent (it does not re-quire U ′

ij = Uij). We see that the gauge structure cancomplicates our analysis of symmetries, since the phys-ical spin wave-function Ψspin({αi}) may has more sym-metries than the mean-field state |Ψ(Uij)

mean〉 before projec-tion.

Let us discuss time reversal symmetry in more detail.A quantum system described by

i~∂tΨ(t) = HΨ(t) (16)

has a time reversal symmetry if Ψ(t) satisfying the equa-tion of motion implies that Ψ∗(−t) also satisfying theequation of motion. This requires that H = H∗. Wesee that, for time reversal symmetric system, if Ψ is aneigenstate, then Ψ∗ will be an eigenstate with the sameenergy.

For our system, the time reversal symmetry means thatif the mean-field wave function Ψ(Uij ,a

liτ

l)mean is a mean-field

ground state wave function for ansatz (Uij , aliτ

l), then(Ψ(Uij ,a

liτ

l)mean

)∗will be the mean-field ground state wave

function for ansatz (U∗ij , a

li(τ

l)∗). That is

(Ψ(Uij ,a

liτ

l)mean

)∗= Ψ

(U∗ij ,ali(τ

l)∗)mean (17)

For a system with time reversal symmetry, the mean-fieldenergy E(Uij , al

iτl) satisfies

E(Uij , aliτ

l) = E(U∗ij , a

li(τ

l)∗) (18)

Thus if an ansatz (Uij , aliτ

l) is a mean-field solution, then(U∗ij , a

li(τ

l)∗) is also a mean-field solution with the samemean-field energy.

From the above discussion, we see that under the timereversal transformation, the ansatz transforms as

Uij → U ′ij = (−iτ2)U∗

ij(iτ2) = −Uij ,

aliτ

l → a′li τl = (−iτ2)(al

iτl)∗(iτ2) = −al

iτl. (19)

Note here we have included an additional SU(2) gaugetransformation Wi = −iτ2. We also note that under thetime reversal transformation, the loop operator trans-forms as PC = eiθ+iθlτ l → (−iτ2)P ∗C(iτ2) = e−iθ+iθlτ l

.We see that the U(1) flux changes the sign while theSU(2) flux is not changed.

Before ending this review section, we would like topoint out that the mean-field ansatz of the spin liquidsUij can be divided into two classes: unfrustrated ansatzwhere Uij only link an even lattice site to an odd latticesite and frustrated ansatz where Uij are nonzero betweentwo even sites and/or two odd sites. An unfrustratedansatz has only pure SU(2) flux through each plaquette,while an frustrated ansatz has U(1) flux of multiple ofπ/2 through some plaquettes in addition to the SU(2)flux.

Page 9: Quantum Orders and Symmetric Spin Liquids - Massachusetts

9

III. SPIN LIQUIDS FROM TRANSLATIONALLYINVARIANT ANSATZ

In this section, we will study many simple examplesof spin liquids and their ansatz. Through those simpleexamples, we gain some understandings on what kind ofspin liquids are possible. Those understandings help usto develop the characterization and classification of spinliquids using projective symmetry group.

Using the above SU(2) projective construction, onecan construct many spin liquid states. To limit ourselves,we will concentrate on spin liquids with translation and90◦ rotation symmetries. Although a mean-field ansatzwith translation and rotation invariance always generatea spin liquid with translation and rotation symmetries,a mean-field ansatz without those invariances can alsogenerate a spin liquid with those symmetries.[95] Becauseof this, it is quite difficult to construct all the translationand rotation symmetric spin liquids. In this section wewill consider a simpler problem. We will limit ourselvesto spin liquids generated from translationally invariantansatz:

Ui+l,j+l = Uij , al0(i) = al

0 (20)

In this case, we only need to find the conditions underwhich the above ansatz can give rise to a rotationallysymmetric spin liquid. First let us introduce uij :

38JijUij = uij (21)

For translationally invariant ansatz, we can introduce ashort-hand notation:

uij = uµ−i+jτ

µ ≡ u−i+j (22)

where u1,2,3l are real, u0

l is imaginary, τ0 is the identitymatrix and τ1,2,3 are the Pauli matrices. The fermionspectrum is determined by Hamiltonian

H =−∑

〈ij〉

(ψ†iuj−iψj + h.c.

)+

i

ψ†ial0τ

lψi (23)

In k-space we have

H = −∑

k

ψ†k(uµ(k)− aµ

0 )τµψk (24)

where µ = 0, 1, 2, 3,

uµ(k) =∑

l

uµl e

il·k, (25)

a00 = 0, and N is the total number of site. The fermion

spectrum has two branches and is given by

E±(k) =u0(k)± E0(k)

E0(k) =√∑

l

(ul(k)− al0)2 (26)

The constraints can be obtained from ∂Eground

∂al0

= 0 andhave a form

N〈ψ†iτ lψi〉

=∑

k,E−(k)<0

ul(k)− al0

E0(k)−

k,E+(k)<0

ul(k)− al0

E0(k)= 0 (27)

which allow us to determine al0, l = 1, 2, 3. It is inter-

esting to see that if u0i = 0 and the ansatz is unfrus-

trated, then we can simply choose al0 = 0 to satisfy the

mean-field constraints (since uµ(k) = −uµ(k+(π, π)) forunfrustrated ansatz). Such ansatz always have time re-versal symmetry. This is because Uij and −Uij are gaugeequivalent for unfrustrated ansatz.

Now let us study some simple examples. First let usassume that only the nearest neighbor coupling ux anduy are non-zero. In order for the ansatz to describe arotationally symmetric state, the rotated ansatz must begauge equivalent to the original ansatz. One can easilycheck that the following ansatz has the rotation symme-try

al0 = 0

ux = χτ3 + ητ1

uy = χτ3 − ητ1 (28)

since the 90◦ rotation followed by a gauge transforma-tion Wi = iτ3 leave the ansatz unchanged. The aboveansatz also has the time reversal symmetry, since timereversal transformation uij → −uij followed by a gaugetransformation Wi = iτ2 leave the ansatz unchanged.

To understand the gauge fluctuations around the abovemean-field state, we note that the mean-field ansatzmay generate non-trivial SU(2) flux through plaquettes.Those flux may break SU(2) gauge structure down toU(1) or Z2 gauge structures as discussed in Ref. [38, 40].In particular, the dynamics of the gauge fluctuations inthe break down from SU(2) to Z2 has been discussed indetail in Ref. [40]. According to Ref. [38, 40], the SU(2)flux plays a role of Higgs fields. A non-trivial SU(2) fluxcorrespond to a condensation of Higgs fields which canbreak the gauge structure and give SU(2) and/or U(1)gauge boson a mass. Thus to understand the dynamicsof the gauge fluctuations, we need to find the SU(2) flux.

The SU(2) flux is defined for loops with a base point.The loop starts and ends at the base point. For example,we can consider the following two loops C1,2 with thesame base point i: C1 = i → i+x → i+x+y → i+y → iand C2 is the 90◦ rotation of C1: C2 = i → i + y →i− x+ y → i− x → i. The SU(2) flux for the two loopsis defined as

PC1 ≡ ui,i+yui+y,i+x+yui+x+y,i+xui+x,i = u†yu†xuyux

PC2 ≡ ui,i−xui−x,i−x+yui−x+y,i+yui+y,i = uxu†yu†xuy(29)

As discussed in Ref. [38, 40], if the SU(2) flux PC for allloops are trivial: PC ∝ τ0, then the SU(2) gauge struc-ture is unbroken. This is the case when χ = η or when

Page 10: Quantum Orders and Symmetric Spin Liquids - Massachusetts

10

η = 0 in the above ansatz Eq. (28). The spinon in thespin liquid described by η = 0 has a large Fermi surface.We will call this state SU(2)-gapless state (This statewas called uniform RVB state in literature). The statewith χ = η has gapless spinons only at isolated k points.We will call such a state SU(2)-linear state to stress thelinear dispersion E ∝ |k| near the Fermi points. (Such astate was called the π-flux state in literature). The lowenergy effective theory for the SU(2)-linear state is de-scribed by massless Dirac fermions (the spinons) coupledto a SU(2) gauge field.

After proper gauge transformations, the SU(2)-gaplessansatz can be rewritten as

ux = iχ

uy = iχ (30)

and the SU(2)-linear ansatz as

ui,i+x = iχ

ui,i+y = i(−)ixχ (31)

In these form, the SU(2) gauge structure is explicit sinceuij ∝ iτ0. Here we would also like to mention that underthe projective-symmetry-group classification, the SU(2)-gapless ansatz Eq. (30) is labeled by SU2An0 and theSU(2)-linear ansatz Eq. (31) by SU2Bn0 (see Eq. (100)).

When χ 6= η 6= 0, The flux PC is non trivial. How-ever, PC commute with PC′ as long as the two loopsC and C ′ have the same base point. In this case theSU(2) gauge structure is broken down to a U(1) gaugestructure.[38, 40] The gapless spinon still only appearat isolated k points. We will call such a state U(1)-linear state. (This state was called staggered flux stateand/or d-wave pairing state in literature.) After a propergauge transformation, the U(1)-linear state can also bedescribed by the ansatz

ui,i+x = iχ− (−)iητ3

ui,i+y = iχ + (−)iητ3 (32)

where the U(1) gauge structure is explicit. Under theprojective-symmetry-group classification, such a state islabeled by U1Cn01n (see section IVC). The low energyeffective theory is described by massless Dirac fermions(the spinons) coupled to a U(1) gauge field.

The above results are all known before. In the follow-ing we are going to study a new class of translation androtation symmetric ansatz, which has a form

al0 =0

ux =iητ0 − χ(τ3 − τ1)

uy =iητ0 − χ(τ3 + τ1) (33)

with χ and η non-zero. The above ansatz describes theSU(2)-gapless spin liquid if χ = 0, and the SU(2)-linearspin liquid if η = 0.

After a 90◦ rotation R90, the above ansatz becomes

ux = −iητ0 − χ(τ3 + τ1)uy = iητ0 − χ(τ3 − τ1) (34)

The rotated ansatz is gauge equivalent to the origi-nal ansatz under the gauge transformation GR90(i) =(−)ix(1 − iτ2)/

√2. After a parity x → −x transforma-

tion Px, Eq. (33) becomes

ux = −iητ0 − χ(τ3 − τ1)uy = iητ0 − χ(τ3 + τ1) (35)

which is gauge equivalent to the original ansatz under thegauge transformation GPx(i) = (−)ixi(τ3 + τ1)/

√2. Un-

der time reversal transformation T , Eq. (33) is changedto

ux = −iητ0 + χ(τ3 − τ1)uy = −iητ0 + χ(τ3 + τ1) (36)

which is again gauge equivalent to the original ansatz un-der the gauge transformation GT (i) = (−)i. (In fact anyansatz which only has links between two non-overlappingsublattices (ie the unfrustrated ansatz) is time reversalsymmetric if al

0 = 0 .) To summarize the ansatz Eq. (33)is invariant under the rotation R90, parity Px, and timereversal transformation T , followed by the following gaugetransformations

GR90(i) =(−)ix(1− iτ2)/√

2

GPx(i) =(−)ixi(τ3 + τ1)/√

2

GT (i) =(−)i (37)

Thus the ansatz Eq. (33) describes a spin liquid whichtranslation, rotation, parity and time reversal symme-tries.

Using the time reversal symmetry we can show thatthe vanishing al

0 in our ansatz Eq. (33) indeed satisfy theconstraint Eq. (27). This is because al

0 → −al0 under the

time reversal transformation. Thus ∂Emean

∂al0

= 0 when

al0 = 0 for any time reversal symmetric ansatz, including

the ansatz Eq. (33).The spinon spectrum is given by (see Fig. 5a)

E± = 2η(sin(kx)+sin(ky))±2|χ|√

2 cos2(kx) + 2 cos2(ky)(38)

The spinons have two Fermi points and two small Fermipockets (for small η). The SU(2) flux is non-trivial. Fur-ther more PC1 and PC2 do not commute. Thus the SU(2)gauge structure is broken down to a Z2 gauge structureby the SU(2) flux PC1 and PC2 .[38, 40] We will call thespin liquid described by Eq. (33) Z2-gapless spin liquid.The low energy effective theory is described by mass-less Dirac fermions and fermions with small Fermi sur-faces, coupled to a Z2 gauge field. Since the Z2 gaugeinteraction is irrelevant at low energies, the spinons are

Page 11: Quantum Orders and Symmetric Spin Liquids - Massachusetts

11

free fermions at low energies and we have a true spin-charge separation in the Z2-gapless spin liquid. The Z2-gapless spin liquid is one of the Z2 spin liquids classifiedin Ref. [66]. Its projective symmetry group is labeledby Z2Aτ13

− τ13+ τ3τ0

− or equivalently by Z2Ax2(12)n (seesection IV B and Eq. (85)).

Now let us include longer links. First we still limitourselves to unfrustrated ansatz. An interesting ansatzis given by

al0 = 0

ux = χτ3 + ητ1

uy = χτ3 − ητ1

u2x+y = λτ2

u−x+2y = −λτ2

u2x−y = λτ2

ux+2y = −λτ2 (39)

By definition, the ansatz is invariant under translationand parity x → −x. After a 90◦ rotation, the ansatz ischanged to

ux = −χτ3 − ητ1

uy = −χτ3 + ητ1

u2x+y = −λτ2

u−x+2y = +λτ2

u2x−y = −λτ2

ux+2y = +λτ2 (40)

which is gauge equivalent to Eq. (39) under the gaugetransformation GR90(i) = iτ3. Thus the ansatz describea spin liquid with translation, rotation, parity and thetime reversal symmetries. The spinon spectrum is givenby (see Fig. 1c)

E± = ±√

ε1(k)2 + ε2(k)2 + ε3(k)2

ε1 = −2χ(cos(kx) + cos(ky))ε2 = −2η(cos(kx)− cos(ky))ε3 = −2λ[cos(2kx + ky) + cos(2kx − ky)

− cos(kx − 2ky)− cos(kx + 2ky)] (41)

Thus the spinons are gapless only at four k points(±π/2,±π/2). We also find that PC3 and PC4 do notcommute, where the loops C3 = i → i + x → i + 2x →i + 2x + y → i and C4 = i → i + y → i + 2y →i + 2y− x → i. Thus the SU(2) flux PC3 and PC4 breakthe SU(2) gauge structure down to a Z2 gauge struc-ture. The spin liquid described by Eq. (39) will be calledthe Z2-linear spin liquid. The low energy effective theoryis described by massless Dirac fermions coupled to a Z2

gauge field. Again the Z2 coupling is irrelevant and thespinons are free fermions at low energies. We have a truespin-charge separation. According to the classificationscheme summarized in section IVB, the above Z2-linearspin liquid is labeled by Z2A003n.

Next let us discuss frustrated ansatz. A simple Z2

spin liquid can be obtained from the following frustratedansatz

a30 6=0, a1,2

0 = 0

ux =χτ3 + ητ1

uy =χτ3 − ητ1

ux+y =γτ3

u−x+y =γτ3 (42)

The ansatz has translation, rotation, parity, and the timereversal symmetries. When a3

0 6= 0, χ 6= ±η and χη 6= 0,al0τ

l does not commute with the loop operators. Thus theansatz breaks the SU(2) gauge structure to a Z2 gaugestructure. The spinon spectrum is given by (see Fig. 1a)

E± =±√

ε2(k) + ∆2(k)

ε(k) =2χ(cos(kx) + cos(ky)) + a30

2γ(cos(kx + ky) + cos(kx − ky))

∆(k) =2η(cos(kx)− cos(ky)) + a30 (43)

which is gapless only at four k points with a linear dis-persion. Thus the spin liquid described by Eq. (42) is aZ2-linear spin liquid, which has a true spin-charge separa-tion. The Z2-linear spin liquid is described by the projec-tive symmetry group Z2A0032 or equivalently Z2A0013.(see section IV B.) From the above two examples of Z2-linear spin liquids, we find that it is possible to obtaintrue spin-charge separation with massless Dirac points(or nodes) within a pure spin model without the chargefluctuations. We also find that there are more than oneway to do it.

A well known frustrated ansatz is the ansatz for thechiral spin liquid[6]

ux = −χτ3 − χτ1

uy = −χτ3 + χτ1

ux+y = ητ2

u−x+y = −ητ2

al0 = 0 (44)

The chiral spin liquid breaks the time reversal and paritysymmetries. The SU(2) gauge structure is unbroken.[38]The low energy effective theory is an SU(2) Chern-Simons theory (of level 1). The spinons are gaped andhave a semionic statistics.[5, 6] The third interesting frus-trated ansatz is given in Ref. [38, 40]

ux =uy = −χτ3

ux+y =ητ1 + λτ2

u−x+y =ητ1 − λτ2

a2,30 =0, a1

0 6= 0 (45)

This ansatz has translation, rotation, parity and the timereversal symmetries. The spinons are fully gaped and

Page 12: Quantum Orders and Symmetric Spin Liquids - Massachusetts

12

the SU(2) gauge structure is broken down to Z2 gaugestructure. We may call such a state Z2-gapped spin liquid(it was called sRVB state in Ref. [38, 40]). It is describedby the projective symmetry group Z2Axx0z. Both thechiral spin liquid and the Z2-gapped spin liquid have truespin-charge separation.

IV. QUANTUM ORDERS IN SYMMETRICSPIN LIQUIDS

A. Quantum orders and projective symmetrygroups

We have seen that there can be many different spinliquids with the same symmetries. The stability analysisin section VIII shows that many of those spin liquids oc-cupy a finite region in phase space and represent stablequantum phases. So here we are facing a similar situa-tion as in quantum Hall effect: there are many distinctquantum phases not separated by symmetries and orderparameters. The quantum Hall liquids have finite energygaps and are rigid states. The concept of topologicalorder was introduced to describe the internal order ofthose rigid states. Here we can also use the topologicalorder to describe the internal orders of rigid spin liquids.However, we also have many other stable quantum spinliquids that have gapless excitations.

To describe internal orders in gapless quantum spinliquids (as well as gapped spin liquids), we have intro-duced a new concept – quantum order – that describesthe internal orders in any quantum phases. The key pointin introducing quantum orders is that quantum phases,in general, cannot be completely characterized by bro-ken symmetries and local order parameters. This pointis illustrated by quantum Hall states and by the stablespin liquids constructed in this paper. However, to makethe concept of quantum order useful, we need to findconcrete mathematical characterizations the quantum or-ders. Since quantum orders are not described by symme-tries and order parameters, we need to find a completelynew way to characterize them. Here we would like topropose to use Projective Symmetry Group to character-ize quantum (or topological) orders in quantum spin liq-uids. The projective symmetry group is motivated fromthe following observation. Although ansatz for differ-ent symmetric spin liquids all have the same symmetry,the ansatz are invariant under symmetry transformationsfollowed by different gauge transformations. We can usethose different gauge transformations to distinguish dif-ferent spin liquids with the same symmetry. In the fol-lowing, we will introduce projective symmetry group ina general and formal setting.

We know that to find quantum numbers that charac-terize a phase is to find the universal properties of thephase. For classical systems, we know that symmetryis a universal property of a phase and we can use sym-metry to characterize different classical phases. To find

universal properties of quantum phases we need to finduniversal properties of many-body wave functions. Thisis too hard. Here we want to simplify the problem by lim-iting ourselves to a subclass of many-body wave functionswhich can be described by ansatz (uij , al

0τl) via Eq. (14).

Instead of looking for the universal properties of many-body wave functions, we try to find the universal prop-erties of ansatz (uij , al

0τl). Certainly, one may object

that the universal properties of the ansatz (or the sub-class of wave functions) may not be the universal prop-erties of spin quantum phase. This is indeed the case forsome ansatz. However, if the mean-field state describedby ansatz (uij , al

0τl) is stable against fluctuations (ie the

fluctuations around the mean-field state do not cause anyinfrared divergence), then the mean-field state faithfullydescribes a spin quantum state and the universal proper-ties of the ansatz will be the universal properties of thecorrespond spin quantum phase. This completes the linkbetween the properties of ansatz and properties of phys-ical spin liquids. Motivated by the Landau’s theory forclassical orders, here we whould like to propose that theinvariance group (or the “symmetry” group) of an ansatzis a universal property of the ansatz. Such a group willbe called the projective symmetry group (PSG). We willshow that PSG can be used to characterize quantum or-ders in quantum spin liquids.

Let us give a detailed definition of PSG. A PSG is aproperty of an ansatz. It is formed by all the transfor-mations that keep the ansatz unchanged. Each trans-formation (or each element in the PSG) can be writtenas a combination of a symmetry transformation U (suchas translation) and a gauge transformation GU . The in-variance of the ansatz under its PSG can be expressedas

GUU(uij) =uij

U(uij) ≡uU(i),U(j)

GU (uij) ≡GU (i)uijG†U (j)

GU (i) ∈SU(2) (46)

for each GUU ∈ PSG.Every PSG contains a special subgroup, which will be

called invariant gauge group (IGG). IGG (denoted by G)for an ansatz is formed by all the gauge transformationsthat leave the ansatz unchanged:

G = {Wi|WiuijW†j = uij ,Wi ∈ SU(2)} (47)

If we want to relate IGG to a symmetry transformation,then the associated transformation is simply an identitytransformation.

If IGG is non-trivial, then for a fixed symmetry trans-formation U , there are can be many gauge transforma-tions GU that leave the ansatz unchanged. If GUU is inthe PSG of uij , GGUU will also be in the PSG iff G ∈ G.Thus for each symmetry transformation U , the differentchoices of GU have a one to one correspondence with theelements in IGG. From the above definition, we see that

Page 13: Quantum Orders and Symmetric Spin Liquids - Massachusetts

13

the PSG, the IGG, and the symmetry group (SG) of anansatz are related:

SG = PSG/IGG (48)

This relation tells us that a PSG is a projective repre-sentation or an extension of the symmetry group.[96] (Inappendix A we will introduce a closely related but differ-ent definition of PSG. To distinguish the two definitions,we will call the PSG defined above invariant PSG andthe PSG defined in appendix A algebraic PSG.)

Certainly the PSG’s for two gauge equivalent ansatzuij and W (i)uijW †(j) are related. From WGUU(uij) =W (uij), where W (uij) ≡ W (i)uijW †(j), we findWGUUW−1W (uij) = WGUW−1

U UW (uij) = W (uij),where WU ≡ UWU−1 is given by WU (i) = W (U(i)).Thus if GUU is in the PSG of ansatz uij , then(WGUWU )U is in the PSG of gauge transformed ansatzW (i)uijW †(j). We see that the gauge transformationGU associated with the symmetry transformation U ischanged in the following way

GU (i) → W (i)GU (i)W †(U(i)) (49)

after a gauge transformation W (i).Since PSG is a property of an ansatz, we can group

all the ansatz sharing the same PSG together to form aclass. We claim that such a class is formed by one orseveral universality classes that correspond to quantumphases. (A more detailed discussion of this importantpoint is given in section VIII E.) It is in this sense we saythat quantum orders are characterized by PSG’s.

We know that a classical order can be described byits symmetry properties. Mathematically, we say that aclassical order is characterized by its symmetry group.Using projective symmetry group to describe a quantumorder, conceptually, is similar to using symmetry groupto describe a classical order. The symmetry descriptionof a classical order is very useful since it allows us toobtain many universal properties, such as the number ofNambu-Goldstone modes, without knowing the details ofthe system. Similarly, knowing the PSG of a quantumorder also allows us to obtain low energy properties ofa quantum system without knowing its details. As anexample, we will discuss a particular kind of the low en-ergy fluctuations – the gauge fluctuations – in a quantumstate. We will show that the low energy gauge fluctua-tions can be determined completely from the PSG. In factthe gauge group of the low energy gauge fluctuations isnothing but the IGG of the ansatz.

To see this, let us assume that, as an example, anIGG G contains a U(1) subgroup which is formed by thefollowing constant gauge transformations

{Wi = eiθτ3 |θ ∈ [0, 2π)} ⊂ G (50)

Now we consider the following type of fluctuationsaround the mean-field solution uij : uij = uije

ia3ijτ

3.

Since uij is invariant under the constant gauge trans-formation eiθτ3

, a spatial dependent gauge transforma-tion eiθiτ

3will transform the fluctuation a3

ij to a3ij =

a3ij+θi−θj . This means that a3

ij and a3ij label the same

physical state and a3ij correspond to gauge fluctuations.

The energy of the fluctuations has a gauge invarianceE({a3

ij}) = E({a3ij}). We see that the mass term of the

gauge field, (a3ij)

2, is not allowed and the U(1) gaugefluctuations described by a3

ij will appear at low energies.If the U(1) subgroup of G is formed by spatial depen-

dent gauge transformations

{Wi = eiθni·τ |θ ∈ [0, 2π), |ni| = 1} ⊂ G, (51)

we can always use a SU(2) gauge transformation to ro-tate ni to the z direction on every site and reduce theproblem to the one discussed above. Thus, regardlessif the gauge transformations in IGG have spatial depen-dence or not, the gauge group for low energy gauge fluc-tuations is always given by G.

We would like to remark that some times low energygauge fluctuations not only appear near k = 0, butalso appear near some other k points. In this case, wewill have several low energy gauge fields, one for each kpoints. Examples of this phenomenon are given by someansatz of SU(2) slave-boson theory discussed in sectionVI, which have an SU(2) × SU(2) gauge structures atlow energies. We see that the low energy gauge struc-ture SU(2)×SU(2) can even be larger than the high en-ergy gauge structure SU(2). Even for this complicatedcase where low energy gauge fluctuations appear arounddifferent k points, IGG still correctly describes the lowenergy gauge structure of the corresponding ansatz. IfIGG contains gauge transformations that are indepen-dent of spatial coordinates, then such transformationscorrespond to the gauge group for gapless gauge fluctua-tions near k = 0. If IGG contains gauge transformationsthat depend on spatial coordinates, then those transfor-mations correspond to the gauge group for gapless gaugefluctuations near non-zero k. Thus IGG gives us a unifiedtreatment of all low energy gauge fluctuations, regardlesstheir momenta.

In this paper, we have used the terms Z2 spin liquids,U(1) spin liquids, SU(2) spin liquids, and SU(2)×SU(2)spin liquids in many places. Now we can have a pre-cise definition of those low energy Z2, U(1), SU(2), andSU(2) × SU(2) gauge groups. Those low energy gaugegroups are nothing but the IGG of the correspondingansatz. They have nothing to do with the high energygauge groups that appear in the SU(2), U(1), or Z2 slave-boson approaches. We also used the terms Z2 gaugestructure, U(1) gauge structure, and SU(2) gauge struc-ture of a mean-field state. Their precise mathematicalmeaning is again the IGG of the corresponding ansatz.When we say a U(1) gauge structure is broken down toa Z2 gauge structure, we mean that an ansatz is changedin such a way that its IGG is changed from U(1) to Z2

group.

Page 14: Quantum Orders and Symmetric Spin Liquids - Massachusetts

14

B. Classification of symmetric Z2 spin liquids

As an application of PSG characterization of quantumorders in spin liquids, we would like to classify the PSG’sassociated with translation transformations assuming theIGG G = Z2. Such a classification leads to a classificationof translation symmetric Z2 spin liquids.

When G = Z2, it contains two elements – gauge trans-formations G1 and G2:

G ={G1, G2}G1(i) =τ0, G2(i) = −τ0. (52)

Let us assume that a Z2 spin liquid has a translation sym-metry. The PSG associated with the translation groupis generated by four elements ±GxTx, ±GyTx where

Tx(uij) = ui−x,j−x, Ty(uij) = ui−y,j−y. (53)

Due to the translation symmetry of the ansatz, we canchoose a gauge in which all the loop operators of theansatz are translation invariant. That is PC1 = PC2 ifthe two loops C1 and C2 are related by a translation.We will call such a gauge uniform gauge.

Under transformation GxTx, a loop operator PC

based at i transforms as PC → Gx(i′)PTxCG†x(i′) =Gx(i′)PCG†x(i′) where i′ = Txi is the base point of thetranslated loop Tx(C). We see that translation invarianceof PC in the uniform gauge requires

Gx(i) = ±τ0, Gy(i) = ±τ0. (54)

since different loop operators based at the same basepoint do not commute for Z2 spin liquids. We note thatthe gauge transformations of form W (i) = ±τ0 do notchange the translation invariant property of the loop op-erators. Thus we can use such gauge transformationsto further simplify Gx,y through Eq. (49). First we canchoose a gauge to make

Gy(i) = τ0. (55)

We note that a gauge transformation satisfying W (i) =W (ix) does not change the condition Gy(i) = τ0. Wecan use such kind of gauge transformations to make

Gx(ix, iy = 0) = τ0. (56)

Since the translations in x- and y-direction commute,Gx,y must satisfy (for any ansatz, Z2 or not Z2)

GxTxGyTy(GxTx)−1(GyTy)−1 =

GxTxGyTyT−1x G−1

x T−1y G−1

y ∈ G. (57)

That means

Gx(i)Gy(i− x)G−1x (i− y)Gy(i)−1 ∈ G (58)

For Z2 spin liquids, Eq. (58) reduces to

Gx(i)G−1x (i− y) = +τ0 (59)

or

Gx(i)G−1x (i− y) = −τ0 (60)

When combined with Eq. (55) and Eq. (56), we find thatthere are only two gauge inequivalent extensions of thetranslation group when IGG is G = Z2. The two PSG’sare given by

Gx(i) =τ0, Gy(i) =τ0 (61)

and

Gx(i) =(−)iyτ0, Gy(i) =τ0 (62)

Thus, under PSG classification, there are only two typesof Z2 spin liquids if they have only the translation sym-metry and no other symmetries. The ansatz that satisfyEq. (61) have a form

ui,i+m =um (63)

and the ones that satisfy Eq. (62) have a form

ui,i+m =(−)myixum (64)

Through the above example, we see that PSG is a verypowerful tool. It can lead to a complete classification of(mean-field) spin liquids with prescribed symmetries andlow energy gauge structures.

In the above, we have studied Z2 spin liquids whichhave only the translation symmetry and no other sym-metries. We find there are only two types of such spinliquids. However, if spin liquids have more symmetries,then they can have much more types. In Ref. [66], wegive a classification of symmetric Z2 spin liquids us-ing PSG. Here we use the term symmetric spin liquidto refer to a spin liquid with the translation symmetryTx,y, the time reversal symmetry T : uij → −uij , andthe three parity symmetries Px: (ix, iy) → (−ix, iy),Py: (ix, iy) → (ix,−iy), and Pxy: (ix, iy) → (iy, ix).The three parity symmetries also imply the 90◦ rota-tion symmetry. The classification is obtained by notic-ing that the gauge transformations Gx,y, GPx,Py,Pxy

and GT must satisfy certain algebraic relations (see ap-pendix A). Solving those algebraic relations and fac-toring out gauge equivalent solutions,[66], we find thatthere are 272 different extensions of the symmetry group{Tx,y, Px,y,xy, T} if IGG G = Z2. Those PSG’s are gen-erated by (GxTx, GyTy, GT T,GPxPx, GPyPy, GPxyPxy).The PSG’s can be divided into two classes. The firstclass is given by

Gx(i) =τ0, Gy(i) =τ0

GPx(i) =ηixxpxηiy

xpygPx GPy (i) =ηixxpyηiy

xpxgPy

GPxy (i) =gPxy GT (i) =ηitgT (65)

and the second class by

Gx(i) =(−)iyτ0, Gy(i) =τ0

GPx(i) =ηixxpxηiy

xpygPx GPy (i) =ηixxpyηiy

xpxgPy

GPxy (i) =(−)ixiygPxy GT (i) =ηitgT (66)

Page 15: Quantum Orders and Symmetric Spin Liquids - Massachusetts

15

Here the three η’s can independently take two values±1. g’s have 17 different choices which are given by (seeRef. [66])

gPxy =τ0 gPx =τ0 gPy =τ0 gT =τ0; (67)

gPxy =τ0 gPx=iτ3 gPy

=iτ3 gT =τ0; (68)

gPxy =iτ3 gPx =τ0 gPy =τ0 gT =τ0; (69)

gPxy =iτ3 gPx =iτ3 gPy =iτ3 gT =τ0; (70)

gPxy =iτ3 gPx=iτ1 gPy

=iτ1 gT =τ0; (71)

gPxy =τ0 gPx =τ0 gPy =τ0 gT =iτ3; (72)

gPxy =τ0 gPx=iτ3 gPy

=iτ3 gT =iτ3; (73)

gPxy =τ0 gPx=iτ1 gPy

=iτ1 gT =iτ3; (74)

gPxy =iτ3 gPx =τ0 gPy =τ0 gT =iτ3; (75)

gPxy =iτ3 gPx =iτ3 gPy =iτ3 gT =iτ3; (76)

gPxy =iτ3 gPx =iτ1 gPy =iτ1 gT =iτ3; (77)

gPxy =iτ1 gPx =τ0 gPy =τ0 gT =iτ3; (78)

gPxy =iτ1 gPx =iτ3 gPy =iτ3 gT =iτ3; (79)

gPxy =iτ1 gPx =iτ1 gPy =iτ1 gT =iτ3; (80)

gPxy =iτ1 gPx =iτ2 gPy =iτ2 gT =iτ3; (81)

gPxy =iτ12 gPx =iτ1 gPy =iτ2 gT =iτ0; (82)

gPxy =iτ12 gPx =iτ1 gPy =iτ2 gT =iτ3; (83)

where

τab =τa + τ b

√2

, τab =τa − τ b

√2

. (84)

Thus there are 2× 17× 23 = 272 different PSG’s. Theycan potentially lead to 272 different types of symmetricZ2 spin liquids on 2D square lattice.

To label the 272 PSG’s, we propose the followingscheme:

Z2A(gpx)ηxpx(gpy)ηxpygpxy(gt)ηt , (85)

Z2B(gpx)ηxpx(gpy)ηxpygpxy(gt)ηt . (86)

The label Z2A... correspond to the case Eq. (65), andthe label Z2B... correspond to the case Eq. (66). A typi-cal label will looks like Z2Aτ1

+τ2−τ12τ3

−. We will also usean abbreviated notation. An abbreviated notation is ob-tained by replacing (τ0, τ1, τ2, τ3) or (τ0

+, τ1+, τ2

+, τ3+) by

(0, 1, 2, 3) and (τ0−, τ1

−, τ2−, τ3

−) by (n, x, y, z). For exam-ple, Z2Aτ1

+τ0−τ12τ3

− can be abbreviated as Z2A1n(12)z.Those 272 different Z2 PSG’s, strictly speaking, are the

so called algebraic PSG’s. The algebraic PSG’s are de-fined as extensions of the symmetry group. They can be

calculated through the algebraic relations listed in ap-pendix A. The algebraic PSG’s are different from theinvariant PSG’s which are defined as a collection of alltransformations that leave an ansatz uij invariant. Al-though an invariant PSG must be an algebraic PSG, analgebraic PSG may not be an invariant PSG. This is be-cause certain algebraic PSG’s have the following proper-ties: any ansatz uij that is invariant under an algebraicPSG may actually be invariant under a larger PSG. Inthis case the original algebraic PSG cannot be an invari-ant PSG of the ansatz. The invariant PSG of the ansatzis really given by the larger PSG. If we limit ourselvesto the spin liquids constructed through the ansatz uij ,then we should drop the algebraic PSG’s are not invari-ant PSG’s. This is because those algebraic PSG’s do notcharacterize mean-field spin liquids.

We find that among the 272 algebraic Z2 PSG’s, atleast 76 of them are not invariant PSG’s. Thus the 272 al-gebraic Z2 PSG’s can at most lead to 196 possible Z2 spinliquids. Since some of the mean-field spin liquid statesmay not survive the quantum fluctuations, the number ofphysical Z2 spin liquids is even smaller. However, for thephysical spin liquids that can be obtained through themean-field states, the PSG’s do offer a characterizationof the quantum orders in those spin liquids.

C. Classification of symmetric U(1) and SU(2) spinliquids

In addition to the Z2 symmetric spin liquids studiedabove, there can be symmetric spin liquids whose lowenergy gauge structure is U(1) or SU(2). Such U(1) andSU(2) symmetric spin liquids (at mean-field level) areclassified by U(1) and SU(2) symmetric PSG’s. The U(1)and SU(2) symmetric PSG’s are calculated in Ref. [66].In the following we just summarize the results.

We find that the PSG’s that characterize mean-fieldsymmetric U(1) spin liquids can be divided into fourtypes: U1A, U1B, U1C and U1m

n . There are 24 typeU1A PSG’s:

Gx =g3(θx), Gy = g3(θy),

GPx =ηiyypxg3(θpx), GPy = ηix

ypxg3(θpy)

GPxy =g3(θpxy), g3(θpxy)iτ1

GT =ηitg3(θt)|ηt=−1, ηitg3(θt)iτ1 (87)

and

Gx =g3(θx), Gy = g3(θy),

GPx =ηixxpxg3(θpx)iτ1, GPy = ηiy

xpxg3(θpy)iτ1

GPxy =g3(θpxy), g3(θpxy)iτ1

GT =ηitg3(θt)|ηt=−1, ηitg3(θt)iτ1 (88)

where

ga(θ) ≡ eiθτa

. (89)

Page 16: Quantum Orders and Symmetric Spin Liquids - Massachusetts

16

We will use U1Aaηxpxbηypxcdηt to label the 24 PSG’s.a, b, c, d are associated with GPx

, GPy, GPxy

, GT re-spectively. They are equal to τ1 if the corresponding Gcontains a τ1 and equal to τ0 otherwise. A typical nota-tion looks like U1Aτ1

−τ1τ0τ1− which can be abbreviated

as U1Ax10x.There are also 24 type U1B PSG’s:

Gx =(−)iyg3(θx), Gy = g3(θy),

GPx =ηiyypxg3(θpx), GPy = ηix

ypxg3(θpy)

(−)ixiyGPxy=g3(θpxy), g3(θpxy)iτ1

GT =ηitg3(θt)|ηt=−1, ηitg3(θt)iτ1 (90)

and

Gx =(−)iyg3(θx), Gy = g3(θy),

GPx=ηix

xpxg3(θpx)iτ1, GPy= ηiy

xpxg3(θpy)iτ1

(−)ixiyGPxy =g3(θpxy), g3(θpxy)iτ1

GT =ηitg3(θt)|ηt=−1, ηitg3(θt)iτ1 (91)

We will use U1Baηxpxbηypxcdηt to label the 24 PSG’s.The 60 type U1C PSG’s are given by

Gx =g3(θx)iτ1, Gy = g3(θy)iτ1,

GPx =ηixxpxηiy

ypxg3(θpx), GPy = ηixypxηiy

xpxg3(θpy)

GPxy =ηixpxyg3(ηipxy

π

4+ θpxy),

GT =ηitg3(θt)|ηt=−1, ηixpxyg3(θt)iτ1 (92)

Gx =g3(θx)iτ1, Gy = g3(θy)iτ1,

GPx =ηixxpxg3(θpx)iτ1, GPy = ηiy

xpxηipxyg3(θpy)iτ1

GPxy =ηixpxyg3(ηipxy

π

4+ θpxy),

GT =ηitg3(θt)|ηt=−1, ηixpxyηitg3(θt)iτ1 (93)

Gx =g3(θx)iτ1, Gy = g3(θy)iτ1,

GPx =ηixxpxηiy

ypxg3(θpx), GPy = ηixypxηiy

xpxg3(θpy)

GPxy =g3(θpxy)iτ1

GT =ηitg3(θt)|ηt=−1 (94)

Gx =g3(θx)iτ1, Gy = g3(θy)iτ1,

GPx =ηixxpxηiy

ypxg3(θpx), GPy = ηixypxηiy

xpxg3(θpy)

GPxy =g3(ηipxy

π

4+ θpxy)iτ1

GT =ηixpxyηitg3(θt)iτ1 (95)

Gx =g3(θx)iτ1, Gy = g3(θy)iτ1,

GPx=ηix

xpxg3(θpx)iτ1, GPy= ηiy

xpxηipxyg3(θpy)iτ1

GPxy=g3(ηipxy

π

4+ θpxy)iτ1

GT =ηitg3(θt)|ηt=−1, ηitηixpxyg3(θt)iτ1 (96)

which will be labeled by U1Caηxpxbηypxcηpxydηt .The type U1m

n PSG’s have not been classified. How-ever, we do know that for each rational number m/n ∈(0, 1), there exist at least one mean-field symmetric spinliquid, which is described by the ansatz

ui,i+x = χτ3, ui,i+y = χg3(mπ

nix)τ3 (97)

It has πm/n flux per plaquette. Thus there are infinitemany type U1m

n spin liquids.We would like to point out that the above 108

U1[A,B,C] PSG’s are algebraic PSG’s. They are only asubset of all possible algebraic U(1) PSG’s. However,they do contain all the invariant U(1) PSG’s of typeU1A, U1B and U1C. We find 46 of the 108 PSG’s arealso invariant PSG’s. Thus there are 46 different mean-field U(1) spin liquids of type U1A, U1B and U1C. Theiransatz and labels are given in Ref. [66].

To classify symmetric SU(2) spin liquids, we find 8different SU(2) PSG’s which are given by

Gx(i) =gx, Gy(i) = gy

GPx(i) =ηixxpxηiy

xpygPx , GPy (i) = ηixxpyηiy

xpxgPy

GPxy (i) =gPxy , GT (i) = (−)igT (98)

and

Gx(i) =(−)iygx, Gy(i) = gy

GPx(i) =ηixxpxηiy

xpygPx , GPy (i) = ηixxpyηiy

xpxgPy

GPxy (i) =(−)ixiygPxy , GT (i) = (−)igT (99)

where g’s are in SU(2). We would like to use the followingtwo notations

SU2Aτ0ηxpx

τ0ηxpy

SU2Bτ0ηxpx

τ0ηxpy

(100)

to denote the above 8 PSG’s. SU2Aτ0ηxpx

τ0ηxpy

is forEq. (98) and SU2Bτ0

ηxpxτ0ηxpy

for Eq. (99). We find only4 of the 8 SU(2) PSG’s, SU2A[n0, 0n] and SU2B[n0, 0n],leads to SU(2) symmetric spin liquids. The SU2An0state is the uniform RVB state and the SU2Bn0 stateis the π-flux state. The other two SU(2) spin liquids aregiven by SU2A0n:

ui,i+2x+y = + iχτ0

ui,i−2x+y =− iχτ0

ui,i+x+2y = + iχτ0

ui,i−x+2y = + iχτ0 (101)

Page 17: Quantum Orders and Symmetric Spin Liquids - Massachusetts

17

and SU2B0n:

ui,i+2x+y = + i(−)ixχτ0

ui,i−2x+y =− i(−)ixχτ0

ui,i+x+2y = + iχτ0

ui,i−x+2y = + iχτ0 (102)

The above results give us a classification of symmet-ric U(1) and SU(2) spin liquids at mean-field level. Ifa mean-field state is stable against fluctuations, it willcorrespond to a physical U(1) or SU(2) symmetric spinliquids. In this way the U(1) and the SU(2) PSG’s alsoprovide an description of some physical spin liquids.

V. CONTINUOUS TRANSITIONS ANDSPINON SPECTRA IN SYMMETRIC SPIN

LIQUIDS

A. Continuous phase transitions without symmetrybreaking

After classifying mean-field symmetric spin liquids, wewould like to know how those symmetric spin liquidsare related to each other. In particular, we would liketo know which spin liquids can change into each otherthrough a continuous phase transition. This problem isstudied in detail in Ref. [66], where the symmetric spinliquids in the neighborhood of some important symmet-ric spin liquids were obtained. After lengthy calculations,we found all the mean-field symmetric spin liquids aroundthe Z2-linear state Z2A001n in Eq. (39), the U(1)-linearstate U1Cn01n in Eq. (32), the SU(2)-gapless stateSU2An0 in Eq. (30), and the SU(2)-linear state SU2Bn0in Eq. (31). We find that, at the mean-field level, theU(1)-linear spin liquid U1Cn01n can continuously changeinto 8 different Z2 spin liquids, the SU(2)-gapless spinliquid SU2An0 can continuously change into 12 U(1) spinliquids and 52 Z2 spin liquids, and the SU(2)-linear spinliquid SU2Bn0 can continuously change into 12 U(1) spinliquids and 58 Z2 spin liquids.

We would like to stress that the above results on thecontinuous transitions are valid only at mean-field level.Some of the mean-field results survive the quantum fluc-tuations while others do not. One need to do a case bycase study to see which mean-field results can be validbeyond the mean-field theory. In Ref. [40], a mean-fieldtransition between a SU(2) × SU(2)-linear spin liquidand a Z2-gapped spin liquid was studied. In particularthe effects of quantum fluctuations were discussed.

We would also like to point out that all the abovespin liquids have the same symmetry. Thus the contin-uous transitions between them, if exist, represent a newclass of continuous transitions which do not change anysymmetries.[67]

B. Symmetric spin liquids around the U(1)-linearspin liquid U1Cn01n

The SU(2)-linear state SU2Bn0 (the π-flux state), theU(1)-linear state U1Cn01n (the staggered-flux/d-wavestate), and the SU(2)-gapless state SU2An0 (the uniformRVB state), are closely related to high Tc superconduc-tors. They reproduce the observed electron spectra func-tion for undoped, underdoped, and overdoped samplesrespectively. However, theoretically, those spin liquidsare unstable at low energies due to the U(1) or SU(2)gauge fluctuations. Those states may change into morestable spin liquids in their neighborhood. In the next afew subsections, we are going to study those more sta-ble spin liquids. Since there are still many different spinliquids involved, we will only present some simplified re-sults by limiting the length of non-zero links. Those spinliquids with short links should be more stable for simplespin Hamiltonians. The length of a link between i and jis defined as |ix − jx|+ |iy − jy|. By studying the spinondispersion in those mean-field states, we can understandsome basic physical properties of those spin liquids, suchas their stability against the gauge fluctuations and thequalitative behaviors of spin correlations which can bemeasured by neutron scattering. Those results allow usto identify them, if those spin liquids exist in certain sam-ples or appear in numerical calculations. We would liketo point out that we will only study symmetric spin liq-uids here. The above three unstable spin liquids may alsochange into some other states that break certain symme-tries. Such symmetry breaking transitions actually havebeen observed in high Tc superconductors (such as thetransitions to antiferromagnetic state, d-wave supercon-ducting state, and stripe state).

First, let us consider the spin liquids around theU(1)-linear state U1Cn01n. In the neighborhood of theU1Cn01n ansatz Eq. (32), there are 8 different spin liq-uids that break the U(1) gauge structure down to a Z2

gauge structure. Those 8 spin liquids are labeled by dif-ferent PSG’s despite they all have the same symmetry.In the following, we will study those 8 Z2 spin liquids inmore detail. In particular, we would like to find out thespinon spectra in them.

The first one is labeled by Z2A0013 and takes the fol-lowing form

ui,i+x =χτ1 − ητ2

ui,i+y =χτ1 + ητ2

ui,i+x+y = + γ1τ1

ui,i−x+y = + γ1τ1

ui,i+2x =γ2τ1 + λ2τ

2

ui,i+2y =γ2τ1 − λ2τ

2

a10 6=0, a2,3

0 = 0 (103)

It has the same quantum order as that in the ansatzEq. (42). The label Z2A0013 tells us the PSG that char-

Page 18: Quantum Orders and Symmetric Spin Liquids - Massachusetts

18

acterizes the spin liquid. The second ansatz is labeled byZ2Azz13:

ui,i+x =χτ1 − ητ2

ui,i+y =χτ1 + ητ2

ui,i+x+y =− γ1τ1

ui,i−x+y = + γ1τ1

ui,i+2x =ui,i+2y = 0

a1,2,30 = 0 (104)

The third one is labeled by Z2A001n (or equivalentlyZ2A003n):

al0 =0

ui,i+x =χτ1 + ητ2

ui,i+y =χτ1 − ητ2

ui,i+2x+y =λτ3

ui,i−x+2y =− λτ3

ui,i+2x−y =λτ3

ui,i+x+2y =− λτ3 (105)

Such a spin liquid has the same quantum order asEq. (39). The fourth one is labeled by Z2Azz1n:

al0 =0

ui,i+x =χτ1 + ητ2

ui,i+y =χτ1 − ητ2

ui,i+2x+y =χ1τ1 + η1τ

2 + λτ3

ui,i−x+2y =χ1τ1 − η1τ

2 + λτ3

ui,i+2x−y =χ1τ1 + η1τ

2 − λτ3

ui,i+x+2y =χ1τ1 − η1τ

2 − λτ3 (106)

The above four ansatz have translation invariance. Thenext four Z2 ansatz do not have translation invariance.(But they still describe translation symmetric spin liquidsafter the projection.) Those Z2 spin liquids areZ2B0013:

ui,i+x =χτ1 − ητ2

ui,i+y =(−)ix(χτ1 + ητ2)

ui,i+2x =− γ2τ1 + λ2τ

2

ui,i+2y =− γ2τ1 − λ2τ

2

a10 6=0, a2,3

0 = 0, (107)

Z2Bzz13:

ui,i+x =χτ1 − ητ2

ui,i+y =(−)ix(χτ1 + ητ2)

ui,i+2x+2y =− γ1τ1

ui,i−2x+2y =γ1τ1

a1,2,30 = 0, (108)

(a) (b)

(c) (d)

FIG. 1: Contour plot of the spinon dispersion E+(k) as afunction of (kx/2π, ky/2π) for the Z2-linear spin liquids. (a)is for the Z2A0013 state in Eq. (103), (b) for the Z2Azz13state in Eq. (104), (c) for the Z2A001n state in Eq. (105), (d)for the Z2Azz1n state in Eq. (106).

Z2B001n:

u ˆi,i+x =χτ1 + ητ2

u ˆi,i+y =(−)ix(χτ1 − ητ2)

u ˆi,i+2x+y =(−)ixλτ3

u ˆi,i−x+2y =− λτ3

u ˆi,i+2x−y =(−)ixλτ3

u ˆi,i+x+2y =− λτ3

al0 =0, (109)

and Z2Bzz1n:

ux =χτ1 + ητ2

uy =(−)ix(χτ1 − ητ2)

u ˆ2x+y =(−)ix(χ1τ1 + η1τ

2 + λτ3)

u ˆ−x+2y =χ1τ1 − η1τ

2 + λτ3

u ˆ2x−y =(−)ix(χ1τ1 + η1τ

2 − λτ3)

u ˆx+2y =χ1τ1 − η1τ

2 − λτ3

al0 =0. (110)

The spinons are gapless at four isolated points witha linear dispersion for the first four Z2 spin liquidsEq. (103), Eq. (104), Eq. (105), and Eq. (106). (SeeFig. 1) Therefore the four ansatz describe symmetricZ2-linear spin liquids. The single spinon dispersion forthe second Z2 spin liquid Z2Azz13 is quite interesting. It

Page 19: Quantum Orders and Symmetric Spin Liquids - Massachusetts

19

(a) (b)

(c) (d)

FIG. 2: Contour plot of the spinon dispersionmin(E1(k), E2(k)) as a function of (kx/2π, ky/2π) forthe Z2-linear states. (a) is for the Z2B0013 state inEq. (107), (b) for the Z2Bzz13 state in Eq. (108), (c) for theZ2B001n state in Eq. (109), (d) for the Z2Bzz1n state inEq. (110).

has the 90◦ rotation symmetry around k = (0, π) and theparity symmetry about k = (0, 0). One very importantthing to notice is that the spinon dispersions for the fourZ2-linear spin liquids, Eq. (103), Eq. (104), Eq. (105),and Eq. (106) have some qualitative differences betweenthem. Those differences can be used to physically mea-sure quantum orders (see section VII).

Next let us consider the ansatz Z2B0013 in Eq. (107).The spinon spectrum for ansatz Eq. (107) is determinedby

H =− 2χ cos(kx)Γ0 − 2η cos(kx)Γ2

− 2χ cos(ky)Γ1 + 2η cos(ky)Γ3 + λΓ4 (111)

where kx ∈ (0, π), ky ∈ (−π, π) and

Γ0 =τ1 ⊗ τ3, Γ1 =τ1 ⊗ τ1,

Γ2 =τ2 ⊗ τ3, Γ3 =τ2 ⊗ τ1,

Γ4 =τ1 ⊗ τ0. (112)

assuming γ1,2 = λ2 = 0. The four bands of spinondispersion have a form ±E1(k), ±E2(k). We find thespinon spectrum vanishes at 8 isolated points near k =(π/2,±π/2). (See Fig. 2a.) Thus the state Z2B0013 is aZ2-linear spin liquid.

Knowing the translation symmetry of the above Z2-linear spin liquid, it seems strange to find that the spinonspectrum is defined only on half of the lattice Brillouinzone. However, this is not inconsistent with translationsymmetry since the single spinon excitation is not phys-ical. Only two-spinon excitations correspond to physical

excitations and their spectrum should be defined on thefull Brillouin zone. Now the problem is that how to ob-tain two-spinon spectrum defined on the full Brillouinzone from the single-spinon spectrum defined on half ofthe Brillouin zone. Let |k, 1〉 and |k, 2〉 be the two eigen-states of single spinon with positive energies E1(k) andE2(k) (here kx ∈ (−π/2, π/2) and ky ∈ (−π, π)). Thetranslation by x (followed by a gauge transformation)change |k, 1〉 and |k, 2〉 to the other two eigenstates withthe same energies:

|k, 1〉 →|k + πy, 1〉|k, 2〉 →|k + πy, 2〉 (113)

Now we see that momentum and the energy of two-spinonstates |k1, α1〉|k2, α2〉 ± |k1 + πy, α1〉|k2 + πy, α2〉 aregiven by

E2spinon =Eα1(k1) + Eα2(k2)k =k1 + k2, k1 + k2 + πx (114)

Eq. (114) allows us to construct two-spinon spectrumfrom single-spinon spectrum.

Now let us consider the ansatz Z2Bzz13 in Eq. (108).The spinon spectrum for ansatz Eq. (108) is determinedby

H =− 2χ cos(kx)Γ0 − 2η cos(kx)Γ2

− 2χ cos(ky)Γ1 + 2η cos(ky)Γ3 (115)− 2γ1 cos(2kx + 2ky)Γ4 + 2γ1 cos(2kx − 2ky)Γ4

where kx ∈ (0, π), ky ∈ (−π, π) and

Γ0 =τ1 ⊗ τ3, Γ1 =τ1 ⊗ τ1,

Γ2 =τ2 ⊗ τ3, Γ3 =τ2 ⊗ τ1,

Γ4 =τ1 ⊗ τ0. (116)

We find the spinon spectrum to vanish at 2 isolated pointsk = (π/2,±π/2). (See Fig. 2b.) The state Z2Bzz13 is aZ2-linear spin liquid.

The spinon spectrum for the ansatz Z2B001n inEq. (109) is determined by

H =− 2χ cos(kx)Γ0 − 2η cos(kx)Γ2

− 2χ cos(ky)Γ1 + 2η cos(ky)Γ3

+ 2λ(cos(kx + 2ky) + cos(−kx + 2ky))Γ4

− 2λ(cos(2kx + ky) + cos(2kx − ky))Γ5 (117)

where kx ∈ (0, π), ky ∈ (−π, π) and

Γ0 =τ1 ⊗ τ3, Γ1 =τ1 ⊗ τ1,

Γ2 =τ2 ⊗ τ3, Γ3 =τ2 ⊗ τ1,

Γ4 =τ3 ⊗ τ3, Γ5 =τ3 ⊗ τ1. (118)

The spinon spectrum vanishes at 2 isolated points k =(π/2,±π/2). (See Fig. 2c.) The state Z2B001n is also aZ2-linear spin liquid.

Page 20: Quantum Orders and Symmetric Spin Liquids - Massachusetts

20

The spinon spectrum for the ansatz Z2Bzz1n inEq. (110) can be obtained from

H =− 2χ cos(kx)Γ0 − 2η cos(kx)Γ2

− 2χ cos(ky)Γ1 + 2η cos(ky)Γ3

+ 2λ(cos(kx + 2ky)− cos(−kx + 2ky))Γ4

− 2λ(cos(2kx + ky)− cos(2kx − ky))Γ5 (119)

where kx ∈ (0, π), ky ∈ (−π, π) and

Γ0 =τ1 ⊗ τ3, Γ1 =τ1 ⊗ τ1,

Γ2 =τ2 ⊗ τ3, Γ3 =τ2 ⊗ τ1,

Γ4 =τ3 ⊗ τ3, Γ5 =τ3 ⊗ τ1. (120)

We have also assumed that χ1 = η1 = 0. The spinonspectrum vanishes at 2 isolated points k = (π/2,±π/2).(See Fig. 2d.) The state Z2Bzz1n is again a Z2-linearspin liquid.

C. Symmetric spin liquids around theSU(2)-gapless spin liquid SU2An0

There are many types of symmetric ansatz in theneighborhood of the SU(2)-gapless state Eq. (30). Letus first consider the 12 classes of symmetric U(1) spinliquids around the SU(2)-gapless state. Here we justpresent the simple cases where uij are non-zero only forlinks with length ≤ 2. Among the 12 classes of symmet-ric ansatz, We find that 5 classes actually give us theSU(2)-gapless spin liquid when the link length is limitedto ≤ 2. The other 7 symmetric U(1) spin liquids aregiven bellow.

The U1Cn01n state:

ui,i+x =χτ1 − ητ2 ui,i+y =χτ1 + ητ2

a1,2,30 =0

Gx =Gy = τ0, GPx =GPy = τ0,

GPxy =iτ1, GT =(−)iτ0 (121)

In the above, we have also listed the gauge transforma-tions Gx,y, GPx,Py,Pxy and GT associated translation,parity and time reversal transformations. Those gaugetransformations define the PSG that characterizes theU(1) spin liquid. In section IVC, we have introduced anotation U1Cn01n to label the above PSG. We will usethe same notation to label the ansatz.

U1Cn00x state:

ui,i+x =χτ1 ui,i+y =χτ1

ui,i+x+y =η1τ3 ui,i−x+y =η1τ

3

ui,i+2x =η2τ3 ui,i+2y =η2τ

3

a30 =η3, a1,2

0 =0

Gx =Gy = τ0, GPx =GPy = τ0,

GPxy =τ0, GT =iτ2 (122)

U1Cn01x state:

ui,i+x =χτ1 ui,i+y =χτ1

ui,i+2x =− η2τ3 ui,i+2y =η2τ

3

a1,2,30 =0

Gx =Gy = τ0, GPx=GPy

= τ0,

GPxy=iτ1, GT =iτ2 (123)

U1Cx10x state:

ui,i+x =χτ1 ui,i+y =χτ1

ui,i+x+y =− ητ3 ui,i−x+y =ητ3

a1,2,30 =0

Gx =Gy = τ0, GPx =GPy = iτ1,

GPxy=τ0, GT =iτ2 (124)

U1A0001 state:

ui,i+x =iχτ0 ui,i+y =iχτ0

ui,i+x+y =− η1τ3 ui,i−x+y =η1τ

3

ui,i+2x =η2τ3 ui,i+2y =η2τ

3

a1,2,30 =0

Gx =Gy = τ0, (−)ixGPx =(−)ixGPy = τ0,

GPxy =τ0, GT =i(−)iτ1 (125)

U1A0011 state:

ui,i+x =iχτ0 ui,i+y =iχτ0

ui,i+2x =− η2τ3 ui,i+2y =η2τ

3

a1,2,30 =0

Gx =Gy = τ0, (−)ixGPx =(−)ixGPy = τ0,

GPxy =iτ1, GT =i(−)iτ1 (126)

U1Ax10x state:

ui,i+x =iχτ0 ui,i+y =iχτ0

ui,i+x+y =ητ3 ui,i−x+y =ητ3

a1,2,30 =0

Gx =Gy = τ0, (−)ixGPx =(−)ixGPy = iτ1,

GPxy =τ0, GT =i(−)iτ1 (127)

In addition to the labels, we also explicitly list thegauge transformations Gx,y, GPx,Py,Pxy and GT for eachansatz. Note that when we define the labels of U(1) PSGfrom the gauge transformations Gx,y, GPx,Py,Pxy and GT ,we have chosen a particular gauge called the canonicalgauge. In the canonical gauge, the IGG is generated byconstant gauge transformation eiθτ3

. Some of the aboveansatz are given in the canonical gauge while other arenot. For the latter ansatz, the listed gauge transforma-tions Gx,y, GPx,Py,Pxy and GT are different from those inthe canonical gauge.

Page 21: Quantum Orders and Symmetric Spin Liquids - Massachusetts

21

(a) (b)

FIG. 3: Contour plot of the spinon dispersion E+(k) as a func-tion of (kx/2π, ky/2π) for (a) the U(1)-linear state U1Cn00xin Eq. (122), and (b) the U(1)-quadratic state U1Cx10x inEq. (124). In the U(1)-quadratic state, the spinon energyvanishes as ∆k2 near two points k = (π, 0), (0, π).

(a) (b)

(c) (d)

FIG. 4: Contour plot of the spinon dispersion E+(k) as afunction of (kx/2π, ky/2π) for the U(1)-gapless states. (a) isfor the U1Cn01x state Eq. (123), (b) for the U1A0001 stateEq. (125), (c) for the U1A0011 state Eq. (126), and (d) forthe U1Ax10x state Eq. (127).

Eq. (121) is the U1Cn01n U(1)-linear state (the stag-gered flux state) studied before. After examining thespinon dispersion, we find that the U1Cn00x state inEq. (122) can be a U(1)-linear or a U(1)-gapped statedepending on the value of a3

0. If it is a U(1)-linear state,it will have 8 isolated Fermi points (see Fig. 3a). TheU1Cn01x state in Eq. (123) is a U(1)-gapless state (seeFig. 4a). The U1Cx10x state in Eq. (124) has twoFermi points at k1 = (π, 0) and k2 = (0, π). (see Fig.3b). However, the spinon energy has a quadratic formE(k) ∝ (k − k1,2)2 near k1 and k2. Thus we call theU1Cx10x spin liquid Eq. (124) a U(1)-quadratic state.The U1A0001 state in Eq. (125), the U1A0011 state inEq. (126), and the U1Ax10x state in Eq. (127) are U(1)-

(a) (b)

FIG. 5: Contour plot of the spinon dispersion E+(k) as afunction of (kx/2π, ky/2π) for the Z2 spin liquids. (a) is forZ2-gapless state Z2Ax2(12)n in Eq. (33), and (b) is for Z2-quadratic state Z2Bx2(12)n in Eq. (128). Despite the lack ofrotation and parity symmetries in the single spinon dispersionin (a), the two-spinon spectrum does have those symmetries.

gapless states (see Fig. 4). Again the spinon dispersionsfor the U(1) spin liquids have some qualitative differencesbetween each other, which can be used to detect differentquantum orders in those U(1) spin liquids.

We next consider the 52 classes of symmetric Z2 spinliquids around the SU(2)-gapless state. Here we justpresent the simplest case where uij are non-zero only forlinks with length ≤ 1. We find that 48 out of 52 classesof ansatz reduce to U(1) or SU(2) spin liquids when thelink’s length is ≤ 1. In the following we discuss the 4remaining Z2 ansatz.

The first one is Z2 spin liquid Z2Ax2(12)n describedby Eq. (33). The second one is Z2 spin liquid Z2A0013described by Eq. (103) or Eq. (42). The third one isZ2 spin liquid Z2By1(12)n (note Z2By1(12)n is gaugeequivalent to Z2Bx2(12)n ):

ui,i+x =iχτ0 + η1τ1

ui,i+y =(−)ix(iχτ0 + η1τ2)

a1,2,30 =0 (128)

The fourth one is Z2 spin liquid Z2B0013, which is de-scribed by Eq. (107).

The ansatz Z2Bx2(12)n in Eq. (128) is a new Z2 spinliquid. The spinon spectrum for ansatz Eq. (128) is de-termined by

H =− 2χ sin(kx)Γ0 + 2η cos(kx)Γ2

− 2χ sin(ky)Γ1 + 2η cos(ky)Γ3 (129)

where kx ∈ (−π/2, π/2), ky ∈ (−π, π) and

Γ0 =τ0 ⊗ τ3, Γ2 =τ1 ⊗ τ3,

Γ1 =τ0 ⊗ τ1, Γ3 =τ2 ⊗ τ1. (130)

The spinon spectrum can be calculated exactly andits four branches take a form ±E1(k) and ±E2(k).The spinon energy vanishes at two isolated points k =(0, 0), (0, π). Near k = 0 the low energy spectrum is given

Page 22: Quantum Orders and Symmetric Spin Liquids - Massachusetts

22

by (see Fig. 5b)

E = ±η−1√

(χ2 + η2)2(k2x − k2

y)2 + 4χ4k2xk2

y (131)

It is interesting to see that the energy does not vanishlinearly as k → 0, instead it vanishes like k2.

We find that the loop operators for the following loopsi → i + x → i + x + y → i + y → i and i → i + y →i− x + y → i− x → i do not commute as long as bothχ and η are non-zero. Thus the spin liquid describedby Eq. (128) indeed has a Z2 gauge structure. We willcall such a state Z2-quadratic spin liquid to stress thequadratic E ∝ k2 dispersion. Such a state cannot beconstructed from translation invariant ansatz. The two-spinon spectrum is still related to the one-spinon spec-trum through Eq. (114).

D. Symmetric spin liquids around the SU(2)-linearspin liquid SU2Bn0

Last, we consider symmetric states in the neighbor-hood of the SU(2)-linear state Eq. (31). We would liketo use the following result proved in Ref. [66]. Given aPSG generated by Gx,y,T and GPx,Py,Pxy , the followinggenerators

Gx(i) =(−)iyGx(i), Gy(i) =Gy(i),

GPx(i) =GPx(i), GPy (i) =GPy (i),

GPxy (i) =(−)ixiyGPxy (i), GT (i) =GT (i). (132)

generate a new PSG. The new PSG has the same IGGand is an extension of the same symmetry group asthe original PSG. The PSG’s for states in the neighbor-hood of the SU(2)-linear state can be obtained throughabove mapping from the PSG’s of symmetric spin liquidsaround the SU(2)-gapless spin liquid.

Here we will only consider the 12 classes of symmet-ric U(1) spin liquids around the SU(2)-linear state. Wewill just present the simple cases where uij are non-zeroonly for links with length ≤ 2. We find that 7 of 12classes of ansatz reduce SU(2)-gapless spin liquids whenthe link length is ≤ 2. Thus we only obtain the following5 symmetric U(1) spin liquids.

U1Cn01n ansatz:

ui,i+x =χτ1 − ητ2 ui,i+y =χτ1 + ητ2

a1,2,30 =0

Gx =Gy = τ0, GPx =GPy = τ0,

GPxy =iτ1, GT =(−)iτ0 (133)

which has the same quantum order as in the U(1)-linearstate Eq. (32).

(a) (b)

FIG. 6: Contour plot of the spinon dispersion E+(k) as a func-tion of (kx/2π, ky/2π) for (a) the U(1)-linear state U1Cn0x1in Eq. (134) and (b) the U(1)-linear state Eq. (142).

U1Cn0x1 ansatz:

ui,i+x =χτ2 ui,i+y =χτ1

ui,i+2x =− ητ3 ui,i+2y =ητ3

a1,2,30 =0

Gx =Gy = τ0, GPx=GPy

= τ0,

GPxy =iτ12, GT =(−)iy iτ1 (134)

U1Cn0n1 ansatz:

ui,i+x =χτ2, ui,i+y = χτ1

ui,i+2x =ητ3, ui,i+2y = ητ3

a30 =η1, a1,2

0 = 0

Gx =Gy = τ0, GT = (−)iy iτ1

GPx =GPy = τ0,

GPxy =(−)ixiyg3(((−)ix − (−)iy )π/4). (135)

U1B0001 ansatz:

ui,i+x =iχτ0 ui,i+y =i(−)ixχτ0

ui,i+2x =ητ3 ui,i+2y =ητ3

a30 =η1, a1,2

0 =0

(−)iyGx =Gy = τ0, (−)ixGPx =(−)ixGPy = τ0,

GPxy =(−)ixiyτ0, GT =i(−)iτ1 (136)

U1B0011 ansatz:

ui,i+x =iχτ0 ui,i+y =i(−)ixχτ0

ui,i+2x =− ητ3 ui,i+2y =ητ3

a1,2,30 =0

(−)iyGx =Gy = τ0, (−)ixGPx =(−)ixGPy = τ0,

GPxy =i(−)ixiyτ1, GT =i(−)iτ1 (137)

Now let us discuss spinon dispersions in the above U(1)spin liquids. The spinon in the U1Cn0x1 state Eq. (134)has 4 linear nodes at (±π/2,±π/2). Thus U1Cn0x1 stateis a U(1)-linear spin liquid. The U1Cn0n1 state Eq. (135)

Page 23: Quantum Orders and Symmetric Spin Liquids - Massachusetts

23

(a) (b)

FIG. 7: Contour plot of the spinon dispersionmin(E1(k), E2(k)) as a function of (kx/2π, ky/2π) forthe U(1) spin liquid states. (a) is for the U(1)-gapless stateU1B0001 in Eq. (136) and (b) is for the U(1)-linear stateU1B0011 in Eq. (137).

has fully gapped spinons and is a U(1)-gapped spin liq-uid.

The four spinon bands in the U1B0001 state Eq. (136)are given by (see Fig. 7a)

±2χ√

sin2(kx) + sin2(ky)± (2η cos(2kx) + 2η cos(2ky) + η1)(138)

We find that the U1B0001 state is a U(1)-gapless spinliquid. The four spinon bands in the U1B0011 stateEq. (137) are given by (see Fig. 7b)

±2χ√

sin2(kx) + sin2(ky)± 2η(cos(2kx)− cos(2ky))(139)

Hence, the U1B0011 state is a U(1)-linear spin liquid.To summarize we list all the spin liquids discussed so

far in the following table:Z2-gapped Z2Axx0z

Z2-linear Z2A0013, Z2Azz13, Z2A001n

Z2Azz1n, Z2B0013, Z2Bzz13Z2B001n, Z2Bzz1n

Z2-quadratic Z2Bx2(12)nZ2-gapless Z2Ax2(12)n

U(1)-gapped U1Cn00x

U(1)-linear U1B0011, U1Cn00x, U1Cn01n

U1Cn0x1U(1)-quadratic U1Cx10x

U(1)-gapless U1A0001, U1A0011, U1Ax10x

U1B0001, U1Cn01x

SU(2)-linear SU2Bn0SU(2)-gapless SU2An0

VI. MEAN-FIELD PHASE DIAGRAM OF J1-J2

MODEL

To see which of the Z2, U(1), and SU(2) spin liquidsdiscussed in the last section have low ground energies

A

ID

B

C

GE

H

J 2

F

−0.5

−0.45

−0.4

−0.35

−0.3

−0.25

−0.2

−0.15

−0.1

−0.05

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−0.5

−0.45

−0.35

−0.3

−0.25

−0.2

−0.15

−0.1

−0.05

0 0.2 0.4 0.6 0.8 1

FIG. 8: The mean-field energies for various phases in a J1-J2 spin system. (A) the π-flux state (the SU(2)-linear stateSU2Bn0). (B) the SU(2)× SU(2)-gapless state in Eq. (140).(C) the SU(2) × SU(2)-linear state in Eq. (141). (D) thechiral spin state (an SU(2)-gapped state). (E) the U(1)-linearstate Eq. (142) which breaks 90◦ rotation symmetry. (F) theU(1)-gapped state U1Cn00x in Eq. (122). (G) the Z2-linearstate Z2Azz13 in Eq. (104). (H) the Z2-linear state Z2A0013in Eq. (103). (I) the uniform RVB state (the SU(2)-gaplessstate SU2An0).

and may appear in real high Tc superconductors, we cal-culate the mean-field energy of a large class of transla-tion invariant ansatz. In Fig. 8, we present the result-ing mean-field phase diagram for a J1-J2 spin system.Here J1 is the nearest-neighbor spin coupling and J2 isthe next-nearest-neighbor spin coupling. We have fixedJ1 + J2 = 1. The y-axis is the mean-field energy persite (multiplied by a factor 8/3). The phase (A) is theπ-flux state (the SU2Bn0 SU(2)-linear state) Eq. (31).The phase (B) is a state with two independent uniformRVB states on the diagonal links. It has SU(2)× SU(2)gauge fluctuations at low energies and will be called anSU(2)× SU(2)-gapless state. Its ansatz is given by

ui,i+x+y =χτ3

ui,i+x−y =χτ3

al0 =0 (140)

The phase (C) is a state with two independent π-fluxstates on the diagonal links. It has SU(2)×SU(2) gaugefluctuations at low energies and will be called an SU(2)×SU(2)-linear state. Its ansatz is given by

ui,i+x+y =χ(τ3 + τ1)

ui,i+x−y =χ(τ3 − τ1)

al0 =0 (141)

The phase (D) is the chiral spin state Eq. (44). The

Page 24: Quantum Orders and Symmetric Spin Liquids - Massachusetts

24

(a) (b)

FIG. 9: Contour plot of the dispersion for spin-1 excitation,E2s(k), as a function of (kx/2π, ky/2π) for (a) the SU(2)-linear spin liquid SU2Bn0 in Eq. (31) (the π-flux phase) and(b) the U(1)-quadratic spin liquid U1Cx10x in Eq. (124).

phase (E) is described by an ansatz

ui,i+x+y =χ1τ1 + χ2τ

2

ui,i+x−y =χ1τ1 − χ2τ

2

ui,i+y =ητ3

al0 =0 (142)

which break the 90◦ rotation symmetry and is a U(1)-linear state (see Fig. 6b). The phase (F) is describedby the U1Cn00x ansatz in Eq. (122). The U1Cn00xstate can be a U(1)-linear or a U(1)-gapped state. Thestate for phase (F) turns out to be a U(1)-gapped state.The phase (G) is described by the Z2Azz13 ansatz inEq. (104) which is a Z2-linear state. The phase (H) isdescribed by the Z2A0013 ansatz in Eq. (103) and is alsoa Z2-linear state. The phase (I) is the uniform RVB state(the SU(2)-gapless state SU2An0 Eq. (30)).

From Fig. 8, we see continuous phase transitions (atmean-field level) between the following pairs of phases:(A,D), (A,G), (B,G), (C,E), and (B,H). The three contin-uous transitions (B,G), (B,H) and (A,G) do not changeany symmetries. We also note that the SU(2) gaugestructure in the phase (A) breaks down to Z2 in thecontinuous transition from the phase (A) to the phase(G). The SU(2) × SU(2) gauge structure in the phase(B) breaks down to Z2 in the two transitions (B,G) and(B,H).

VII. PHYSICAL MEASUREMENTS OFQUANTUM ORDERS

After characterizing the quantum orders using PSGmathematically, we would like to ask how to measurequantum orders in experiments. The quantum ordersin gapped states are related to the topological orders.The measurement of topological orders are discussed inRef. [9, 10, 65]. The quantum order in a state with gap-less excitations can be measured, in general, by the dy-namical properties of gapless excitation. However, notall dynamical properties are universal. Thus we need to

(a) (b)

(c) (d)

FIG. 10: Contour plot of E2s(k) as a function of(kx/2π, ky/2π) for the Z2-linear spin liquids. (a) is for theZ2A0013 state in Eq. (103), (b) for the Z2Azz13 state inEq. (104), (c) for the Z2A001n state in Eq. (105), (d) for theZ2Azz1n state in Eq. (106).

identify the universal properties of gapless excitations,before using them to characterize and measure quantumorders. The PSG characterization of quantum orders al-lows us obtain those universal properties. We simplyneed to identify the common properties of gapless exci-tations that are shared by all the ansatz with the samePSG.

To demonstrate the above idea, we would like to studythe spectrum of two-spinon excitations. We note thatspinons can only be created in pairs. Thus the one-spinon spectrum is not physical. We also note that thetwo-spinon spectrum include spin-1 excitations which canbe measured in experiments. At a given momentum,the two-spinon spectrum is distributed in one or severalranges of energy. Let E2s(k) be the lower edge of thetwo-spinon spectrum at momentum k. In the mean-fieldtheory, the two-spinon spectrum can be constructed fromthe one-spinon dispersion

E2-spinon(k) = E1-spinon(q) + E1-spinon(k − q) (143)

In Fig. 9 – 15 we present mean-field E2s for some simplespin liquids. If the mean-field state is stable against thegauge fluctuations, we expect the mean-field E2s shouldqualitatively agrees with the real E2s.

Among our examples, there are eight Z2-linear spinliquids (see Fig. 10 and Fig. 11). We see that some ofthose eight different Z2-linear spin liquids (or eight dif-ferent quantum orders) have different number of gaplesspoints. The gapless points of some spin liquids are pinnedat position k = (π, π) and/or k = (π, 0), (0, π). By mea-suring the low energy spin excitations (say using neutron

Page 25: Quantum Orders and Symmetric Spin Liquids - Massachusetts

25

(a) (b)

(c) (d)

FIG. 11: Contour plot of E2s(k) as a function of(kx/2π, ky/2π) for the Z2-linear spin liquids. (a) is for theZ2B0013 state in Eq. (107), (b) for the Z2Bzz13 state inEq. (108), (c) for the Z2B001n state in Eq. (109), (d) forthe Z2Bzz1n state in Eq. (110).

(a) (b)

FIG. 12: Contour plot of E2s(k) as a function of(kx/2π, ky/2π) for (a) the Z2-gapless state Z2Ax2(12)n inEq. (33), and (b) the Z2-quadratic state Z2Bx2(12)n inEq. (128).

scattering), we can distinguish those Z2 spin liquids. Wenote that all the two-spinon spectra have rotation andparity symmetries around k = 0. This is expected. Sincethe two-spinon spectra are physical, they should have allthe symmetries the spin liquids have.

We also have four U(1)-linear spin liquids. Some ofthem can be distinguished by their different numbers ofgapless points. It is interesting to note that all the U(1)spin liquids discussed here have a gapless point in thetwo-spinon spectrum pinned at position k = (π, π). TheU(1)-linear spin liquids are also different from the Z2-linear spin liquids in that the spin-spin correlations havedifferent decay exponents once the U(1) gauge fluctua-tions are included. We also see that E2s has a quadraticform E2s ∝ k2 for the U(1)-quadratic spin liquid. E2s

(a) (b)

FIG. 13: Contour plot of E2s(k) as a function of(kx/2π, ky/2π) for two U(1)-linear spin liquids. (a) is for theU1Cn01n state Eq. (32) (the staggered flux phase), and (b)for the U1Cn00x state Eq. (122) in the gapless phase.

(a) (b)

FIG. 14: Contour plot of the two-spinon dispersion E2s(k)as a function of (kx/2π, ky/2π) for (a) the U(1)-linear spinliquid state U1Cn0x1 in Eq. (134) and (b) the U(1)-linearspin liquid Eq. (142).

vanishes in two finite regions in k-space for the Z2-gaplessspin liquids.

Neutron scattering experiments probe the two-spinonsector. Thus low energy neutron scattering allows us tomeasure quantum orders in high Tc superconductors.

Let us discuss the U(1) linear state U1Cn01n (thestaggered-flux state) in more detail. The U1Cn01nstate is proposed to describe the pseudo-gap metallicstate in underdoped high Tc superconductors.[33, 34] TheU1Cn01n state naturally explains the spin pseudo-gap

(a) (b)

FIG. 15: Contour plot of the two-spinon dispersion E2s(k) asa function of (kx/2π, ky/2π) for the U(1) spin liquid states.(a) is for the U(1)-gapless state U1B0001 in Eq. (136) and (b)is for the U(1)-linear state U1B0011 in Eq. (137).

Page 26: Quantum Orders and Symmetric Spin Liquids - Massachusetts

26

in the underdoped metallic state. As an algebraic spinliquid, the U1Cn01n state also explain the Luttinger-likeelectron spectral function[34] and the enhancement of the(π, π) spin fluctuations[68] in the pseudo-gap state. FromFig. 13a, we see that gapless points of the spin-1 exci-tations in the U1Cn01n state are always at k = (π, π),(0, 0), (π, 0) and (0, π). The equal energy contour for theedge of the spin-1 continuum has a shape of two over-lapped ellipses at all the four k points. Also the energycontours are not perpendicular to the zone boundary. Allthose are the universal properties of the U1Cn01n state.Measuring those properties in neutron scattering experi-ments will allow us to determine if the pseudo-gap metal-lic state is described by the U1Cn01n (the staggered-flux)state or not.

We have seen that at low energies, the U1Cn01nstate is unstable due to the instanton effect. Thus theU1Cn01n state has to change into some other states, suchas the 8 Z2 spin liquids discussed in section V or someother states not discussed in this paper. From Fig. 10a,we see that the transition from the U1Cn01n state to theZ2-linear state Z2A0013 can be detected by neutron scat-tering if one observe the splitting of the node at (π, π)into four nodes at (π ± δ, π ± δ) and the splitting of thenodes at (π, 0) and (0, π) into two nodes at (π ± δ, 0)and (0, π ± δ). From Fig. 10b, we see that, for thetransition from the U1Cn01n state to the Z2-linear stateZ2Azz13, the node at (π, π) still splits into four nodesat (π ± δ, π ± δ). However, the nodes at (π, 0) and (0, π)split differently into two nodes at (π,±δ) and (±δ, π). Wecan also study the transition from the U1Cn01n state toother 6 Z2 spin liquids. We find the spectrum of spin-1 excitations all change in certain characteristic ways.Thus by measuring the spin-1 excitation spectrum andits evolution, we not only can detect a quantum transi-tion that do not change any symmetries, we can also tellwhich transition is happening.

The neutron scattering on high Tc superconductor in-deed showed a splitting of the scattering peak at (π, π)into four peaks at (π ± δ, π), (π, π ± δ) [30, 69–75] orinto two peaks at (π, π) → (π + δ, π − δ), (π − δ, π + δ)[28, 76] as we lower the energy. This is consistent withour belief that the U1Cn01n state is unstable at low en-ergies. However, it is still unclear if we can identify theposition of the neutron scattering peak as the position ofthe node in the spin-1 spectrum. If we do identify thescattering peak as the node, then non of the 8 Z2 spinliquids in the neighborhood of the U1Cn01n state canexplain the splitting pattern (π ± δ, π), (π, π ± δ). Thiswill imply that the U1Cn01n state change into anotherstate not studied in this paper. This example illustratesthat detailed neutron scattering experiments are power-ful tools in detecting quantum orders and studying newtransitions between quantum orders that may not changeany symmetries.

VIII. FOUR CLASSES OF SPIN LIQUIDS ANDTHEIR STABILITY

We have concentrated on the mean-field states ofspin liquids and presented many examples of mean-fieldansatz for symmetric spin liquids. In order for thosemean-field states to represent real physical spin liquids,we need to include the gauge fluctuations. We alsoneed to show that the inclusion of the gauge fluctuationsdoes not destabilize the mean-field states at low ener-gies. This requires that (a) the gauge interaction is nottoo strong and (b) the gauge interaction is not a rele-vant perturbation. (The gauge interaction, however, canbe a marginal perturbation.) The requirement (a) canbe satisfied through large N limit and/or adjustment ofshort-range spin couplings in the spin Hamiltonian, ifnecessary. Here we will mainly consider the requirement(b). We find that, at least in certain large N limits,many (but not all) mean-field states do correspond toreal quantum spin liquids which are stable at low ener-gies. In this case, the characterization of the mean-fieldstates by PSG’s correspond to the characterization of realquantum spin liquids.

All spin liquids (with odd number of electron per unitcell) studied so far can be divided into four classes. Inthe following we will study each classes in turn.

A. Rigid spin liquid

In rigid spin liquids, by definition, the spinons and allother excitations are fully gaped. The gapped gauge fieldonly induces short range interaction between spinons dueto Chern-Simons terms or Anderson-Higgs mechanism.By definition, the rigid spin liquids are locally stable andself consistent. The rigid spin liquids are characterizedby topological orders and they have the true spin-chargeseparation. The low energy effective theories for rigidspin liquids are topological field theories. The Z2-gappedspin liquid and chiral spin liquid are examples of rigidspin liquids.

B. Bose spin liquid

The U(1)-gapped spin liquid discussed in the last sec-tion is not a rigid spin liquid. It is a Bose spin liquid.Although the spinon excitations are gapped, the U(1)gauge fluctuations are gapless in the U(1)-gapped spinliquid. The dynamics of the gapless U(1) gauge fluctua-tions are described by low energy effective theory

L =12g

(fµν)2 (144)

where fµν is the field strength of the U(1) gauge field.However, in 1+2 dimension and after including the in-stanton effect, the U(1) gauge fluctuations will gain an

Page 27: Quantum Orders and Symmetric Spin Liquids - Massachusetts

27

energy gap.[77] The properties of the resulting quantumstate remain to be an open problem.

C. Fermi spin liquid

The Fermi spin liquids have gapless excitations that aredescribed by spin 1/2 fermions. Those gapless excitationshave only short range interactions between them. TheZ2-linear, Z2-quadratic and the Z2-gapless spin liquiddiscussed above are examples of the Fermi spin liquids.

The spinons have a massless Dirac dispersion in Z2-linear spin liquids. Thus Z2-linear spin liquids are lo-cally stable since short range interactions between mass-less Dirac fermions are irrelevant at 1+2 dimensions. Wewould like to point out that the massless Dirac dispersionof the Z2-linear spin liquids are protected by the PSG (orthe quantum order). That is any perturbations around,for example, the Z2-linear ansatz Z2A003z in Eq. (39)cannot destroy the massless Dirac dispersion as long asthe PSG are not changed by the perturbations. To un-derstand this result, we start with the most general formof symmetric perturbations (see Ref. [66])

ui,i+m =ulmτ l|l=1,2,3

u1,2Pxy(m) =− u1,2

m

u3Pxy(m) =u3

m

u1,2,3Px(m) =u1,2,3

m

u1,2,3Py(m) =u1,2,3

m

um =0, for m = even (145)

around the Z2-linear ansatz Eq. (39). We find thatsuch perturbations vanish in the momentum space atk = (±π/2,±π/2). The translation, parity, and thetime reversal symmetries do not allow any mass terms orchemical potential terms. Thus the Z2-linear spin liquidis a phase that occupy a finite region in the phase space(at T = 0). One does not need any fine tuning of couplingconstants and uij to get massless Dirac spectrum.

Now let us consider the stability of the Z2-quadraticspin liquid Z2Bx2(12)n in Eq. (128). The spinons have agapless quadratic dispersion in the Z2-quadratic spin liq-uid. The gapless quadratic dispersion of the Z2-quadraticspin liquid is also protected by the symmetries. The mostgeneral form of symmetric perturbations around the Z2-quadratic ansatz Eq. (128) is given by (see Ref. [66])

ui,i+m =(−)myix(u0mτ0 + u1

mτ1 + u2mτ2)

u0,1,2m =0, for m = even (146)

In the momentum space, the most general symmetric Z2-quadratic ansatz give rise to the following Hamiltonian

(after considering the 90◦ rotation symmetry)

H =− 2∑

χmn[sin(nkx −mky)Γ0 + sin(mkx + nky)Γ1]

+ 2∑

ηmn[cos(nkx −mky)Γ2 + cos(mkx + nky)Γ3]

+ 2∑

λmn[cos(nkx −mky)Γ4 + cos(mkx + nky)Γ5](147)

where

Γ0 =τ0 ⊗ τ3, Γ2 =τ1 ⊗ τ3,

Γ1 =τ0 ⊗ τ1, Γ3 =τ2 ⊗ τ1,

Γ4 =− τ2 ⊗ τ3, Γ5 =τ1 ⊗ τ1. (148)

and the summation is over m =even, n =odd. We findthat the spinon dispersion still vanish at k = (0, 0), (0, π)and the energy still satisfy E ∝ k2. The translation, par-ity, and the time reversal symmetric perturbations do notchange the qualitative behavior of the low energy spinondispersion. Thus, at mean-field level, the Z2-quadraticspin liquid is a phase that occupy a finite region in thephase space (at T = 0). One does not need any fine tun-ing of coupling constants to get gapless quadratic disper-sion of the spinons. However, unlike the Z2-linear spinliquid, the short range four-fermion interactions betweenthe gapless spinons in the Z2-quadratic state are marginalat 1+2 dimensions. Further studies are needed to under-stand the dynamical stability of the Z2-quadratic spinliquid beyond the mean-field level.

The Z2-gapless spin liquid is as stable as Fermi liquid in1+2 dimensions. Again we expect Z2-gapless spin liquidto be a phase that occupy a finite region in the phasespace, at least at mean-field level.

D. Algebraic spin liquid

U(1)-linear spin liquids are examples of algebraic spinliquids. Their low lying excitations are described bymassless Dirac fermions coupled to U(1) gauge field.Although the massless Dirac fermions are protected byquantum orders, the gauge couplings remain large at lowenergies. Thus the low lying excitations in the U(1)-linear spin liquids are not described by free fermions.This makes the discussion on the stability of those statesmuch more difficult.

Here we would like to concentrate on the U(1)-linearspin liquid U1Cn01n in Eq. (32). The spinons have amassless Dirac dispersion in the U(1)-linear spin liquid.First we would like to know if the massless Dirac dis-persion is generic property of the U(1)-linear spin liquid,ie if the massless Dirac dispersion is a property shared byall the spin liquids that have the same quantum order asthat in Eq. (32). The most general perturbations aroundthe U(1)-linear ansatz Eq. (32) are given by

δui,i+m = δu0mτ0 + (−)iδu3

mτ3 (149)

Page 28: Quantum Orders and Symmetric Spin Liquids - Massachusetts

28

ui,i+m =u0mτ0 + (−)iu3

mτ3

u0,3m =0, for m = even

u0Pxy(m) =u0

m

u3Pxy(m) =− u3

m

u0,3Px(m) =(−)mxu0,3

m

u0,3Py(m) =(−)myu0,3

m (150)

if the perturbations respect translation, parity, and thetime reversal symmetries, and if the perturbation do notbreak the U(1) gauge structure. Since δu3

m = δu0m =

0 for m = even, their contributions in the momentumspace vanish at k = (0, 0) and k = (0, π). The spinonenergy also vanish at those points for the ansatz Eq. (32).Thus the massless Dirac dispersion is protected by thesymmetries and the U(1) gauge structure in the U(1)-linear spin liquid Eq. (32). In other words, the masslessDirac dispersion is protected by the quantum order inthe U(1)-linear spin liquid.

Next we consider if the symmetries and the U(1) gaugestructure in the U(1)-linear spin liquid can be brokenspontaneously due to interactions/fluctuations at low en-ergy. The low energy effective theory is described by La-grangian (in imaginary time)

L =∑a,µ

ψ†aγ0[vµ,aγµ(∂µ + iaµ)]ψa (151)

where µ = 0, 1, 2, a = 1, 2, γµ are 4 × 4 γ-matrices,v0

a = 1, and (v1,a, v2,a) are velocities for aθ fermion inx and y directions. We make a large N generalizationof the above effective theory and allow a = 1, 2, .., N .Our first concern is about whether the self energy fromthe gauge interaction can generate any mass/chemical-potential term, due to infrared divergence. It turns outthat, in the 1/N expansion, the gauge fluctuations repre-sent an exact marginal perturbation that does not gener-ate any mass/chemical=potential term.[78] Instead thegauge interaction changes the quantum fixed point de-scribed by free massless Dirac fermions to a new quantumfixed point which has no free fermionic excitations at lowenergies.[34, 78] The new quantum fixed point has gaplessexcitations and correlation functions all have algebraicdecay. Such a quantum fixed point was called algebraicspin liquid.[34] Actually, it is easy to understand why thegauge fluctuations represent an exact marginal perturba-tion. This is because the conserved current that couple tothe gauge potential cannot have any anomalous dimen-sions. Thus if the gauge interaction is marginal at firstorder, then it is marginal at all orders. Gauge interactionas an exact marginal perturbation is also supported bythe following results. The gauge invariant Green’s func-tion of ψ is found to be gapless after coupling to gaugefield, to all orders in 1/N expansion.[78] Recently it wasargued that the U(1) gauge interaction do not generateany mass perturbatively even when N is as small as 2.[79]

Now let us discuss other possible instabilities. First wewould like consider a possible instability that change the

U(1)-linear state to the Z2-linear state. To study such aninstability we add a charge-2 Higgs field to our effectivetheory

L =ψ†aγ0[γµ(∂µ + iaµ)]ψa (152)

+ |(∂0 − 2ia0)φ|2 + v2|(∂i − 2iai)φ|2 + V (φ)

where V (φ) has its minimum at φ = 0 and we have as-sumed v1,a = v2,a = 1 for simplicity. (Note that φ corre-sponds to λ in Eq. (39). It is a non-zero λ that break theU(1) gauge structure down to the Z2 gauge structure.) Ifafter integrating out ψ and aµ, the resulting effective po-tential Veff (|φ|) has its minimum at a non zero φ, thenthe U(1)-linear state has a instability towards the Z2-linear state.

To calculate Veff (|φ|), we first integrate out ψ and get

L =12aµπµνaν (153)

+ |(∂0 − 2ia0)φ|2 + v2|(∂i − 2iai)φ|2 + V (φ)

where

πµν =N

8(p2)−1/2(p2δµν − pµpν) (154)

Now the effective potential Veff (|φ|) can be obtained byintegrating out aµ (in the a0 = 0 gauge) and the phase θof the φ field, φ = ρeiθ:[97]

Veff (φ)− V (φ) (155)

=∫ ∞

0

π

∫d2k

(2π)212Im

(ln[−K⊥(iω)] + ln[−K||(iω)]

)

=∫ ∞

0

π

∫d2k

(2π)2Im ln(−N

8(k2 − ω2 − 0+)1/2 − 4|φ|2)

where

K⊥ =N

8(ω2 + k2)1/2 + 4|φ|2 (156)

K|| =N

8(ω2 + k2)−1/2ω2 + 4|φ|2 ω2

ω2 + v2k2

=ω2

ω2 + v2k2

(N

8(ω2 + k2)1/2 + 4|φ|2

)

We find that Veff = V −C1|φ|6 ln |φ| where C1 is a con-stant. Now it is clear that the gapless gauge fluctuationscannot shift the minimum of V from φ = 0 and the U(1)-linear state is stable against spontaneously changing intothe Z2-linear state.

So far we only considered the effects of perturbativefluctuations. The non-perturbative instanton effects canalso cause instability of the algebraic spin liquid. Theinstanton effects have been discussed in Ref. [60] for thecase v1

a = v2a. It was found that the instanton effects

represent a relevant perturbation which can destabilizethe algebraic spin liquid when N < 24. In the following,

Page 29: Quantum Orders and Symmetric Spin Liquids - Massachusetts

29

we will generalize the analysis of Ref. [60] to v1a 6= v2

a

case. First we rewrite

S =∫

d3k

(2π)312aµ(−k)πµνaν(k)

=∫

d3k

(2π)312fµ(−k)Kµνfν(k) (157)

where

fµ = εµνλ∂νaλ (158)

When πµν = k2δµν − kµkν , we find Kµν = δµν . Whenπµν = (k2δµν − kµkν)/

√k2, we may assume Kµν =

δµν/√

k2. When v1,a 6= v2,a we have

(Kµν) =∑

a

18(ω2 + v2

1,ak21 + v2

2,ak22)−1/2×

v1,av2,a 0 00 v2,a/v1,a 00 0 v1,a/v2,a

(159)

The instanton field fµ minimize the action Eq. (157) andsatisfies

Kµνfν = c(k)kµ (160)

where c(k) is chosen such that kµfµ = 2iπ We find that

f0 =8cω∑

a(ω2 + v21,ak2

1 + v22,ak2

2)−1/2v1,av2,a

f1 =8ck1∑

a(ω2 + v21,ak2

1 + v22,ak2

2)−1/2v2,a/v1,a

f2 =8ck2∑

a(ω2 + v21,ak2

1 + v22,ak2

2)−1/2v1,a/v2,a(161)

and

c =2iπ

(8ω2

∑a(ω2 + v2

1,ak21 + v2

2,ak22)−1/2v1,av2,a

+8k2

1∑a(ω2 + v2

1,ak21 + v2

2,ak22)−1/2v2,a/v1,a

+8k2

2∑a(ω2 + v2

1,ak21 + v2

2,ak22)−1/2v1,a/v2,a

)−1

(162)

Using the above solution, we can calculate the action fora single instanton, which has a form

Sinst =N

2α(v2/v1) ln(L) (163)

where L is the size of the system and we have assumedthat N/2 fermions have velocity (vx, vy) = (v1, v2) andthe other N/2 fermions have velocity (vx, vy) = (v2, v1).We find α(1) = 1/4 + O(1/N) and α(0.003) = 3. +

O(1/N). When N2 α(v2/v1) > 3, the instanton effect is

irrelevant. We see that even for the case N = 2, the in-stanton effect can be irrelevant for small enough v2/v1.Therefore, the algebraic spin liquid exists and can be sta-ble.

It has been proposed that the pseudo-gap metallicstate in underdoped high Tc superconductors is describedby the (doped) staggered flux state (the U(1)-linearstate U1Cn01n) which contains a long range U(1) gaugeinteraction.[33, 34] From the above result, we see that,for realistic v2/v1 ∼ 0.1 in high Tc superconductors, theU1Cn01n spin liquid is unstable at low energies. How-ever, this does not mean that we cannot not use the al-gebraic spin liquid U1Cn01n to describe the pseudo-gapmetallic state. It simply means that, at low temper-atures, the algebraic spin liquid will change into otherstable quantum states, such as superconducting state orantiferromagnetic state[80] as observed in experiments.

The unstable algebraic spin liquid can be viewed asan unstable quantum fixed point. Thus the algebraic-spin-liquid approach to the pseudo-gap metallic statein underdoped samples looks similar to the quantum-critical-point approach[81, 82]. However, there is an im-portant distinction between the two approaches. Thequantum-critical-point approach assumes a nearby con-tinuous phase transition that changes symmetries andstrong fluctuations of local order parameters that causethe criticality. The algebraic-spin-liquid approach doesnot require any nearby symmetry breaking state andthere is no local order parameter to fluctuate.

E. Quantum order and the stability of spin liquids

After introducing quantum orders and PSG, we canhave a deeper discussion on the stability of mean-fieldstates. The existence of the algebraic spin liquid is avery striking phenomenon, since gapless excitations in-teract down to zero energy and cannot be described byfree fermions or free bosons. According to a conventionalwisdom, if bosons/fermions interact at low energies, theinteraction will open an energy gap for those low lying ex-citations. This implies that a system can either has freebosonic/fermionic excitations at low energies or has nolow energy excitations at all. According to the discussionin section VIII, such a conventional wisdom is incorrect.But it nevertheless rises an important question: whatprotects gapless excitations (in particular when they in-teract at all energy scales). There should be a “reason”or “principle” for the existence of the gapless excitations.Here we would like to propose that it is the quantum or-der that protects the gapless excitations. We would liketo stress that gapless excitations in the Fermi spin liq-uids and in the algebraic spin liquids exist even withoutany spontaneous symmetry breaking and they are notprotected by symmetries. The existence of gapless ex-citations without symmetry breaking is a truly remark-able feature of quantum ordered states. In addition to

Page 30: Quantum Orders and Symmetric Spin Liquids - Massachusetts

30

the gapless Nambu-Goldstone modes from spontaneouscontinuous symmetry breaking, quantum orders offer an-other origin for gapless excitations.

We have seen from several examples discussed in sec-tion V that the quantum order (or the PSG) not onlyprotect the zero energy gap, it also protects certain qual-itative properties of the low energy excitations. Thoseproperties include the linear, quadratic, or gapless disper-sions, the k locations where the 2-spinon energy E2s(k)vanishes, etc .

Since quantum order is a generic property for anyquantum state at zero temperature, we expect thatthe existence of interacting gapless excitations is also ageneric property of quantum state. We see that algebraicstate is a norm. It is the Fermi liquid state that is special.

In the following, we would like to argue that the PSGcan be a stable (or universal) property of a quantumstate. It is robust against perturbative fluctuations.Thus, the PSG, as a universal property, can be used tocharacterize a quantum phase. From the examples dis-cussed in sections VIII C and VIII D, we see that PSGprotects gapless excitations. Thus, the stability of PSGalso imply the stability of gapless excitations.

We know that a mean-field spin liquid state is char-acterized by Uij = 〈ψiψ†j〉. If we include perturbativefluctuations around the mean-field state, we expect Uijto receive perturbative corrections δUij . Here we wouldlike to argue that the perturbative fluctuations can onlychange Uij in such a way that Uij and Uij + δUij havethe same PSG.

First we would like to note the following well knowfacts: the perturbative fluctuations cannot change thesymmetries and the gauge structures. For example, ifUij and the Hamiltonian have a symmetry, then δUijgenerated by perturbative fluctuations will have the samesymmetry. Similarly, the perturbative fluctuations can-not generate δUij that, for example, break a U(1) gaugestructure down to a Z2 gauge structure.

Since both the gauge structure (described by the IGG)and the symmetry are part of the PSG, it is reasonableto generalize the above observation by saying that notonly the IGG and the symmetry in the PSG cannot bechanged, the whole PSG cannot be changed by the per-turbative fluctuations. In fact, the mean-field Hamilto-nian and the mean-field ground state are invariant underthe transformations in the PSG. Thus in a perturbativecalculation around a mean-field state, the transforma-tions in the PSG behave just like symmetry transforma-tions. Therefore, the perturbative fluctuations can onlygenerate δUij that are invariant under the transforma-tions in the PSG.

Since the perturbative fluctuations (by definition) donot change the phase, Uij and Uij + δUij describe thesame phase. In other words, we can group Uij into classes(which are called universality classes) such that Uij ineach class are connected by the the perturbative fluctu-ations and describe the same phase. We see that if theabove argument is true then the universality classes are

classified by the PSG’s (or quantum orders).We would like to point out that we have assumed the

perturbative fluctuations to have no infrared divergencein the above discussion. The infrared divergence impliesthe perturbative fluctuations to be relevant perturba-tions, which cause phase transitions.

IX. RELATION TO PREVIOUSLYCONSTRUCTED SPIN LIQUIDS

Since the discovery of high Tc superconductor in 1987,many spin liquids were constructed. After classifying andconstructing a large class of spin liquids, we would liketo understand the relation between the previously con-structed spin liquids and the spin liquids constructed inthis paper.

Anderson, Baskaran, and Zou[14–16] first used theslave-boson approach to construct uniform RVB state.The uniform RVB state is a symmetric spin liquid whichhas all the symmetries of the lattice. It is a SU(2)-gaplessstate characterized by the PSG SU2An0. Later two morespin liquids were constructed using the same U(1) slave-boson approach. One is the π-flux phase and the otheris the staggered-flux/d-wave state.[31, 32, 83] The π-fluxphase is a SU(2)-linear symmetric spin liquid character-ized by PSG SU2Bn0. The staggered-flux/d-wave stateis a U(1)-linear symmetric spin liquid characterized byPSG U1Cn01n. The U1Cn01n state was found to be themean-field ground for underdoped samples. Upon dopingthe U1Cn01n state becomes a metal with a pseudo-gapat high temperatures and a d-wave superconductor at lowtemperatures.

It is amazing to see that the slave-boson approach,which is regarded as a very unreliable approach, pre-dicted the d-wave superconducting state 5 years before itsexperimental confirmation.[21, 84–86] Maybe predictingthe d-wave superconductor is not a big deal. After all,the d-wave superconductor is a commonly known stateand the paramagnon approach[87, 88] predicted d-wavesuperconductor before the slave-boson approach. How-ever, what is really a big deal is that the slave bosonapproach also predicted the pseudo-gap metal which is acompletely new state of matter. It is very rare in con-densed matter physics to predict a new state of matterbefore experiments.

The above U(1) and SU(2) spin liquids are likely tobe unstable at low energies and may not appear as theground states of spin systems. The first known stablespin liquid is the chiral spin liquid.[5, 6]. It has a truespin-charge separation. The spinons and holons carryfractional statistics. Such a state breaks the time rever-sal and parity symmetries and is a SU(2)-gapped state.The SU(2) gauge fluctuations in the chiral spin state doesnot cause any instability since the gauge fluctuations aresuppressed and become massive due to the Chern-Simonsterm. Due to the broken time reversal and parity sym-metries, the chiral spin state does not fit within our clas-

Page 31: Quantum Orders and Symmetric Spin Liquids - Massachusetts

31

sification scheme.Spin liquids can also be constructed using the slave-

fermion/σ-model approach.[41, 42] Some gapped spin liq-uids were constructed using this approach.[43, 44] Thosestates turn out to be Z2 spin liquids. But they are notsymmetric spin liquids since the 90◦ rotation symmetryis broken. Thus they do not fit within our classifica-tion scheme. Later, a Z2-gapped symmetric spin liquidwas constructed using the SU(2) slave-boson approach(or the SU(2) projective construction).[38] The PSG forsuch a state is Z2Axx0z. Recently, another Z2 state wasconstructed using slave-boson approach.[63, 64] It is aZ2-linear symmetric spin liquid. Its PSG is given byZ2A0013. New Z2 spin liquids were also obtained re-cently using the slave-fermion/σ-model approach.[46] Itappears that most of those states break certain symme-tries and are not symmetric spin liquids. We would liketo mention that Z2 spin liquids have a nice property thatthey are stable at low energies and can appear as theground states of spin systems.

Many spin liquids were also obtained in quantum dimermodel,[47–51] and in various numerical approach.[52–55]It is hard to compare those states with the spin liquidsconstructed here. This is because either the spectrum ofspin-1 excitations was not calculated or the model has avery different symmetry than the model discussed here.We need to generalized our classification to models withdifferent symmetries so that we can have a direct com-parison with those interesting results and with the non-symmetric spin liquids obtained in the slave-fermion/σ-model approach. In quantum dimer model and in nu-merical approach, we usually know the explicit form ofground state wave function. However, at moment, we donot know how to obtain PSG from ground wave function.Thus, knowing the explicit ground state wave functiondoes not help us to obtain PSG. We see that it is im-portant to understand the relation between the groundstate wave function and PSG so that we can understandquantum order in the states obtained in numerical calcu-lations.

X. SUMMARY OF THE MAIN RESULTS

In the following we will list the main results obtainedin this paper. The summary also serves as a guide of thewhole paper.

(1) A concept of quantum order is introduced. Thequantum order describes the orders in zero-temperaturequantum states. The opposite of quantum order – classi-cal order describes the orders in finite-temperature clas-sical states. Mathematically, the quantum order charac-terizes universality classes of complex ground state wave-functions. It is richer then the classical order that char-acterizes the universality classes of positive distributionfunctions. Quantum orders cannot be completely de-scribed by symmetries and order parameters. Landau’stheory of orders and phase transitions does not apply to

quantum orders. (See section IA)(2) Projective symmetry group is introduced to de-

scribe different quantum orders. It is argued that PSGis a universal property of a quantum phase. PSG ex-tends the symmetry group description of classical ordersand can distinguishes different quantum orders with thesame symmetries. (See section IVA and VIII E)

(3) As an application of the PSG description of quan-tum phases, we propose the following principle that gov-ern the continuous phase transition between quantumphases. Let PSG1 and PSG2 be the PSG’s of the twoquantum phases on the two sides of a transition, andPSGcr be the PSG that describes the quantum criti-cal state. Then PSG1 ⊆ PSGcr and PSG2 ⊆ PSGcr.We note that the two quantum phases may have thesame symmetry and continuous quantum phase transi-tions are possible between quantum phases with samesymmetry.[67] The continuous transitions between dif-ferent mean-field symmetric spin liquids are discussedin section V and in Ref. [66] which demonstrate theabove principle. However, for continuous transitions be-tween mean-field states, we have an additional conditionPSG1 = PSGcr or PSG2 = PSGcr.

(4) With the help of PSG, we find that, within theSU(2) mean-field slave-boson approach, there are 4 sym-metric SU(2) spin liquids and infinite many symmetricU(1) spin liquids. There are at least 103 and at most196 symmetric Z2 spin liquids. Those symmetric spinliquids have translation, rotation, parity and the time re-versal symmetries. Although the classifications are donefor the mean-field states, they apply to real physical spinliquids if the corresponding mean-field states turn out tobe stable against fluctuations. (See section IV)

(5) The stability of mean-field spin liquid states are dis-cussed in detail. We find many gapless mean-field spinliquids to be stable against quantum fluctuations. Theycan be stable even in the presence of long range gauge in-teractions. In that case the mean-field spin liquid statesbecome algebraic spin liquids where the gapless excita-tions interact down to zero energy. (See section VIII)

(6) The existence of algebraic spin liquids is a strikingphenomenon since there is no spontaneous broken sym-metry to protect the gapless excitations. There should bea “principle” that prevents the interacting gapless exci-tations from opening an energy gap and makes the alge-braic spin liquids stable. We propose that quantum orderis such a principle. To support our idea, we showed thatjust like the symmetry group of a classical state deter-mines the gapless Nambu-Goldstone modes, the PSG ofa quantum state determines the structure of gapless exci-tations. The gauge group of the low energy gauge fluctu-ations is given by the IGG, a subgroup of the PSG. ThePSG also protects massless Dirac fermions from gaininga mass due to radiative corrections. We see that the sta-bilities of algebraic spin liquids and Fermi spin liquidsare protected by their PSG’s. The existence of gaplessexcitations (the gauge bosons and gapless fermions) with-out symmetry breaking is a truly remarkable feature of

Page 32: Quantum Orders and Symmetric Spin Liquids - Massachusetts

32

quantum ordered states. The gapless gauge and fermionexcitations are originated from the quantum orders, justlike the phonons are originated from translation symme-try breaking. (See sections VIII C, VIII D, VIII E anddiscussions below Eq. (49))

(7) Many Z2 spin liquids are constructed. Their lowenergy excitations are described by free fermions. SomeZ2 spin liquids have gapless excitations and others havefinite energy gap. For those gapless Z2 spin liquids somehave Fermi surface while others have only Fermi points.The spinon dispersion near the Fermi points can be lin-ear E ∝ |k| (which gives us Z2-linear spin liquids) orquadratic E ∝ k2 (which gives us Z2-quadratic spin liq-uids). In particular, we find there can be many Z2-linearspin liquids with different quantum orders. All those dif-ferent Z2-linear spin liquids have nodal spinon excita-tions. (See section III and V)

(8) Many U(1) spin liquids are constructed. Some U(1)spin liquids have gapless excitations near isolated Fermipoint with a linear dispersion. Those U(1) linear statescan be stable against quantum fluctuations. Due to longrange U(1) gauge fluctuations, the gapless excitations in-teract at low energies. The U(1)-linear spin liquids canbe concrete realizations of algebraic spin liquids.[34, 68](See section III and V)

(9) Spin liquids with the same symmetry and differ-ent quantum orders can have continuous phase transi-tions between them. Those phase transitions are verysimilar to the continuous topological phase transitionsbetween quantum Hall states.[67, 89–91] We find that,at mean-field level, the U1Cn01n spin liquid in Eq. (32)(the staggered flux phase) can continuously change into8 different symmetric Z2 spin liquids. The SU2An0 spinliquid in Eq. (30) (the uniform RVB state) can contin-uously change into 12 symmetric U(1) spin liquids and52 symmetric Z2 spin liquids. The SU2Bn0 spin liquidin Eq. (31) (the π-flux phase) can continuously changeinto 12 symmetric U(1) spin liquids and 58 symmetricZ2 spin liquids. (See Ref. [66])

(10) We show that spectrum of spin-1 excitations(ie the two-spinon spectrum), which can be probed inneutron scattering experiments, can be used to measurequantum orders. The gapless points of the spin-1 ex-citations in the U1Cn01n (the staggered-flux) state arealways at k = (π, π), (0, 0), (π, 0) and (0, π). In thepseudo-gap metallic phase of underdoped high Tc super-conductors, the observed splitting of the neutron scat-tering peak (π, π) → (π ± δ, π), (π, π ± δ) [30, 69–75] or(π, π) → (π + δ, π − δ), (π − δ, π + δ) [28, 76] at low en-ergies indicates a transition of the U1Cn01n state intoa state with a different quantum order, if we can indeedidentify the scattering peak as the gapless node. Non ofthe 8 symmetric Z2 spin liquids in the neighborhood ofthe the U1Cn01n state can explain the splitting pattern.Thus we might need to construct a new low energy stateto explain the splitting. This illustrates that detailedneutron scattering experiments are powerful tools in de-tecting quantum orders and studying transitions between

quantum orders. (See section VII.)(11) The mean-field phase diagram Fig. 8 for a J1-

J2 spin system is calculated. (Only translation sym-metric states are considered.) We find four mean-fieldground states as we change J2/J1: the π-flux state (theSU2An0 state), the chiral spin state (an SU(2)-gappedstate), the U(1)-linear state Eq. (142) which breaks 90◦rotation symmetry, and the SU(2) × SU(2)-linear stateEq. (141). We also find several locally stable mean-fieldstates: the U(1)-gapped state U1Cn00x in Eq. (122) andtwo Z2-linear states Z2Azz13 in Eq. (104) and Z2A0013in Eq. (103). Those spin liquids have a better chanceto appear in underdoped high Tc superconductors. TheZ2A0013 Z2-linear state has a spinon dispersion very sim-ilar to electron dispersion observed in underdoped sam-ples. The spinon dispersion in the Z2Azz13 Z2-linearstate may also be consistent with electron dispersion inunderdoped samples. We note that the two-spinon spec-trum for the two Z2-linear states have some qualitativedifferences (see Fig. 10a and Fig. 10b and note the posi-tions of the nodes). Thus we can use neutron scatteringto distinguish the two states. (See section VI.)

Next we list some remarks/comments that may clar-ify certain confusing points and help to avoid possiblemisunderstanding.

(A) Gauge structure is simply a redundant labelingof quantum states. The “gauge symmetries” (referringdifferent labels of same physical state give rise to thesame result) are not symmetries and can never be broken.(See the discussion below Eq. (15))

(B) The gauge structures referred in this paper (suchas in Z2, U(1), or SU(2) spin liquids) are “low energy”gauge structures. They are different from the “highenergy” gauge structure that appear in Z2, U(1), andSU(2) slave-boson approaches. The “low energy” gaugestructures are properties of the quantum orders in theground state of a spin system. The “high energy” gaugestructure is a particular way of writing down the Hamil-tonian of spin systems. The two kinds of gauge structureshave nothing to do with each other. (See discussions atthe end of section I C and at the end of section IV A)

(C) There are (at least) two different interpretations ofspin-charge separation. The first interpretation (pseudospin-charge separation) simply means that the low energyexcitations cannot be described by electron-like quasipar-ticles. The second interpretation (true spin-charge sepa-ration) means the existence of free spin-1/2 neutral quasi-particles and spin-0 charged quasiparticles. In this paperboth interpretations are used. The algebraic spin liquidshave a pseudo spin-charge separation. The Z2 and chi-ral spin liquids have a true spin-charge separation. (Seesection I C)

(D) Although in this paper we stress that quantumorders can be characterized by the PSG’s, we need topoint out that the PSG’s do not completely characterizequantum orders. Two different quantum orders may becharacterized by the same PSG. As an example, we haveseen that the ansatz Eq. (122) can be a U(1)-linear state

Page 33: Quantum Orders and Symmetric Spin Liquids - Massachusetts

33

or a U(1)-gapped state depending on the values of pa-rameters in the ansatz. Both states are described by thesame PSG U1Cn00x. Thus the PSG can not distinguishthe different quantum orders carried by the U(1)-linearstate and the U(1)-gapped state.

(E) The unstable spin liquids can be important inunderstanding the finite temperature states in high Tc

superconductors. The pseudo-gap metallic state in un-derdoped samples is likely to be described by the un-stable U1Cn01n algebraic spin liquid (the staggeredflux state) which contains a long range U(1) gaugeinteraction.[33, 34] (See discussions at the end of sectionVIII.)

(F) Although we have been concentrated on the char-acterization of stable quantum states, quantum order andthe PSG characterization can also be used to describethe internal order of quantum critical states. Here wedefine “quantum critical states” as states that appearat the continuous phase transition points between twostates with different symmetries or between two stateswith different quantum orders (but the same symme-try). We would like to point out that “quantum criti-cal states” thus defined are more general than “quantumcritical points”. “Quantum critical points”, by defini-tion, are the continuous phase transition points betweentwo states with different symmetries. The distinctionis important. “Quantum critical points” are associatedwith broken symmetries and order parameters. Thus thelow energy excitations at “quantum critical points” comefrom the strong fluctuations of order parameters. While“quantum critical states” may not be related to brokensymmetries and order parameters. In that case it is im-possible to relate the gapless fluctuations in a “quantumcritical state” to fluctuations of an order parameter. Theunstable spin liquids mentioned in (E) can be more gen-eral quantum critical states. Since some finite tempera-ture phases in high Tc superconductors may be describedby quantum critical states or stable algebraic spin liquids,their characterization through quantum order and PSG’sis useful for describing those finite temperature phases.

(G) In this paper, we only studied quantum ordersand topological orders at zero temperature. However, wewould like point out that topological orders and quan-tum orders may also apply to finite temperature systems.Quantum effect can be important even at finite tempera-tures. In Ref. [13], a dimension index (DI) is introducedto characterize the robustness of the ground state degen-eracy of a topologically ordered state. We find that ifDI≤ 1 topological orders cannot exist at finite tempera-ture. However, if DI> 1, topological order can exist atfinite temperatures and one expect a finite-temperaturephase transition without any change of symmetry. Topo-logical orders in FQH states have DI=1, and they cannotexist at finite temperatures. The topological order in 3Dsuperconductors has DI=2. Such a topological order canexist at finite temperatures, and we have a continuousfinite-temperature superconductor-metal transition thatdo not change any symmetry.

Although we mainly discussed quantum orders in 2Dspin systems, the concept of quantum order is not limitedto 2D spin systems. The concept applies to any quantumsystems in any dimensions. Actually, a superconductor isthe simplest example of a state with non trivial quantumorder if the dynamical electromagnetic fluctuations areincluded. A superconductor breaks no symmetries andcannot be characterized by order parameters. An s-waveand a d-wave superconductors, having the same symme-try, are distinguished only by their different quantumorders. The gapless excitations in a d-wave supercon-ductor are not produced by broken symmetries, but byquantum orders. We see that a superconductor has manyproperties characteristic of quantum ordered states, andit is a quantum ordered state. The quantum orders inthe superconducting states can also be characterized us-ing PSG’s. The IGG G = Z2 if the superconductingstate is caused by electron-pair condensation, and theIGG G = Z4 if the superconducting state is caused byfour-electron-cluster condensation. The different quan-tum orders in an s-wave and a d-wave superconductorscan be distinguished by their different PSG’s. The ansatzof the s-wave superconductor is invariant under the 90◦rotation, while the ansatz of the d-wave superconduc-tor is invariant under the 90◦ rotation followed by gaugetransformations ci → ±eiπ/2ci.

It would be interesting to study quantum orders in3D systems. In particular, it is interesting to find outthe quantum order that describes the physical vacuumthat we all live in. The existence of light – a masslessexcitation – without any sign of spontaneous symmetrybreaking suggests that our vacuum contains a non-trivialquantum order that protect the massless photons. Thusquantum order provides an origin of light.[58]

I would like to thank P. A. Lee, J. Moore, E. Frad-kin, W. Rantner, T. Senthil for many helpful discussions.This work is supported by NSF Grant No. DMR–97–14198 and by NSF-MRSEC Grant No. DMR–98–08941.

APPENDIX A: GENERAL CONDITIONS ONPROJECTIVE SYMMETRY GROUPS

The transformations in a symmetry group satisfy vari-ous algebraic relations so that they form a group. Thosealgebraic relations leads to conditions on the elementsof the PSG. Solving those conditions for a given sym-metry group and a given IGG allows us to find pos-sible extensions of the symmetry group, or in anotherword, to find possible PSG’s associated with the sym-metry group. In section IV A, we have seen that therelation TxTyT−1

x T−1y = 1 between translations in x- and

y-directions leads to condition

GxTxGyTy(GxTx)−1(GyTy)−1 =

GxTxGyTyT−1x G−1

x T−1y G−1

y ∈ G; (A1)

Page 34: Quantum Orders and Symmetric Spin Liquids - Massachusetts

34

or

Gx(i)Gy(i− x)G−1x (i− y)Gy(i)−1 ∈ G (A2)

on elements GxTx and GyTy of the PSG. Here G is theIGG. This condition allows us to determine that thereare only two different extensions (given by Eq. (61) andEq. (62)) for the translation group generated by Tx andTy, if G = Z2.

However, a bigger symmetry groups can have manymore extensions. In the following we are going to con-sider PSG’s for the symmetry group generated by twotranslations Tx,y, three parity transformations Px,y,xy,and the time reversal transformation T . Since transla-tions and the time reversal transformation commute wehave,

(GxTx)−1(GT T )−1GxTxGT T ∈ G(GyTy)−1(GT T )−1GyTyGT T ∈ G (A3)

which reduces to the following two conditions on Gx,y(i)and GT (i)

G−1x (i)G−1

T (i)Gx(i)GT (i− x) ∈ GG−1

y (i)G−1T (i)Gy(i)GT (i− y) ∈ G (A4)

Since T−1P−1x TPx = 1, T−1P−1

y TPy = 1, andT−1P−1

xy TPxy = 1, one can also show that

G−1T (Px(i))G−1

Px(i)GT (i)GPx(i) ∈G

G−1T (Py(i))G−1

Py(i)GT (i)GPy (i) ∈G

G−1T (Pxy(i))G−1

Pxy(i)GT (i)GPxy (i) ∈G (A5)

From the relation between the translations and the par-ity transformations, TxP−1

x TxPx = T−1y P−1

x TyPx =TyP−1

y TyPy = T−1x P−1

y TxPy = T−1y P−1

xy TxPxy =T−1

x P−1xy TyPxy = 1, we find that

(GxTx)(GPxPx)−1GxTxGPxPx ∈G(GyTy)−1(GPxPx)−1GyTyGPxPx ∈G (A6)

(GyTy)(GPyPy)−1GyTyGPyPy ∈G(GxTx)−1(GPyPy)−1GxTxGPyPy ∈G (A7)

(GyTy)−1(GPxyPxy)−1GxTxGPxyPxy ∈G(GxTx)−1(GPxyPxy)−1GyTyGPxyPxy ∈G (A8)

or

Gx(Px(i))G−1Px

(i + x)Gx(i + x)GPx(i) ∈GG−1

y (Px(i))G−1Px

(i)Gy(i)GPx(i− y) ∈G (A9)

Gy(Py(i))G−1Py

(i + y)Gy(i + y)GPy (i) ∈GG−1

x (Py(i))G−1Py

(i)Gx(i)GPy (i− x) ∈G (A10)

G−1y (Pxy(i))G−1

Pxy(i)Gx(i)GPxy

(i− x) ∈GG−1

x (Pxy(i))G−1Pxy

(i)Gy(i)GPxy(i− y) ∈G (A11)

We also have PxyPxPxyP−1y = PyPxP−1

y P−1x = 1. Thus

GPxyPxyGPx

PxGPxyPxy(GPy

Py)−1 ∈GGPyPyGPxPx(GPyPy)−1(GPxPx)−1 ∈G (A12)

which implies

GPxy(i)GPx

(Pxy(i))GPxy(PxyPx(i))G−1

Py(i) ∈G

GPy(i)GPx

(Py(i))G−1Py

(Px(i))G−1Px

(i) ∈G (A13)

The fact T 2 = 1 leads to condition

G2T (i) ∈ G (A14)

and P 2x = P 2

y = P 2xy = 1 leads to

GPx(i)GPx(Px(i)) ∈GGPy (i)GPy (Py(i)) ∈G

GPxy (i)GPxy (Pxy(i)) ∈G (A15)

The above conditions completely determine the PSG’s.The solutions of the above equations for G = Z2, U(1),and SU(2) allow us to obtain PSG’s for Z2, U(1), andSU(2) spin liquids. However, we would like to point outthat the above conditions define the so called algebraicPSG’s, which are somewhat different from the invariantPSG defined in section IV A. More precisely, an algebraicPSG is defined for a given IGG and a given symmetrygroup SG. It is a group equipped with a projection P andsatisfies the following conditions

IGG ⊂PSG, P (PSG) = SG (A16)P (gu) =P (u), for any u ∈ PSG and g ∈ IGG

It is clear that an invariant PSG is always an algebraicPSG. However, some algebraic PSG’s are not invariantPSG’s. This is because a generic ansatz uij that are in-variant under an algebraic PSG may be invariant undera larger invariant PSG. If we limit ourselves to spin liq-uids constructed using uij , then an algebraic PSG char-acterizes a mean-field spin liquid only when it is also aninvariant PSG at the same time.

We would like to remark that the definition of invari-ant PSG can be generalized. In section IVA, the in-variant PSG is defined as a collection of transformationsthat leave an ansatz uij invariant. More generally, aspin liquid is not only characterized by the two-pointcorrelation (Uij)αβ =

⟨ψαiψ

†βj

⟩but also by many-point

correlations such as (Uijmn)αβγλ =⟨ψαiψβjψ

†γmψ†λn

⟩.

We may define the generalized invariant PSG as a collec-tion of transformations that leave many-point correlationinvariant. It would be very interesting to see if the gen-eralized invariant PSG coincide with the algebraic PSG.

Page 35: Quantum Orders and Symmetric Spin Liquids - Massachusetts

35

[1] L. D. Landau and E. M. Lifschitz, Satistical Physics -Course of Theoretical Physics Vol 5, Pergamon, London,1958.

[2] V. L. Ginzburg and L. D. Landau, J. exp. theor. Phys.20, 1064 (1950).

[3] D. C. Tsui, H. L. Stormer, and A. C. Gossard, Phys.Rev. Lett. 48, 1559 (1982).

[4] R. B. Laughlin, Phys. Rev. Lett. 50, 1395 (1983).[5] V. Kalmeyer and R. B. Laughlin, Phys. Rev. Lett. 59,

2095 (1987).[6] X.-G. Wen, F. Wilczek, and A. Zee, Phys. Rev. B 39,

11413 (1989).[7] X.-G. Wen, Phys. Rev. B 40, 7387 (1989).[8] X.-G. Wen and Q. Niu, Phys. Rev. B 41, 9377 (1990).[9] X.-G. Wen, Int. J. Mod. Phys. B 4, 239 (1990).

[10] X.-G. Wen, Advances in Physics 44, 405 (1995).[11] X.-G. Wen, cond-mat/0110397.[12] H. K. Onnes, Comm. Phys. Lab. Univ. Leiden, Nos 119

120, 122 (1911).[13] X.-G. Wen, Int. J. Mod. Phys. B 5, 1641 (1991).[14] P. W. Anderson, Science 235, 1196 (1987).[15] G. Baskaran, Z. Zou, and P. W. Anderson, Solid State

Comm. 63, 973 (1987).[16] G. Baskaran and P. W. Anderson, Phys. Rev. B 37, 580

(1988).[17] D. S. Marshall et al., Phys. Rev. Lett. 76, 4841 (1996).[18] H. Ding et al., Nature 382, 51 (1996).[19] A. J. Millis, L. Ioffe, and H. Monien, J. Phys. Chem. Sol.

56, 1641 (1995).[20] N. P. Ong, Z. Z. Wang, J. C. J. M. Tarascon, L. H.

Greene, and W. R. McKinnon, Phys. Rev. B 35, 8807(1987).

[21] W. N. Hardy, D. A. Bonn, D. C. Morgan, R. Liang, andK. Zhang, Phys. Rev. Lett. 70, 3999 (1993).

[22] Y. J. Uemura et al., Phys. Rev. Lett. 62, 2317 (1989).[23] Y. J. Uemura et al., Phys. Rev. Lett. 66, 2665 (1991).[24] S. A. Kivelson and V. J. Emery, Nature 374, 434 (1995).[25] A. G. Loeser et al., Science 273, 325 (1996).[26] J. M. Harris, Z.-X. Shen, P. J. White, D. S. Marshall,

and M. C. Schabel, Phys. Rev. B 54, 665 (1996).[27] M. Takigawa et al., Phys. Rev. B 43, 247 (1991).[28] M. Matsuda et al., Phys. Rev. B 62, 9148 (2000).[29] P. Bourges et al., Science 288, 1234 (2000).[30] P. Dai, H. A. Mook, R. D. Hunt, and F. Doan, Phys.

Rev. B 63, 54525 (2001).[31] I. Affleck and J. B. Marston, Phys. Rev. B 37, 3774

(1988).[32] G. Kotliar and J. Liu, Phys. Rev. B 38, 5142 (1988).[33] X.-G. Wen and P. A. Lee, Phys. Rev. Lett. 76, 503

(1996).[34] W. Rantner and X.-G. Wen, Phys. Rev. Lett. 86, 3871

(2001).[35] Y. Suzumura, Y. Hasegawa, and H. Fukuyama, J. Phys.

Soc. Jpn. 57, 2768 (1988).[36] I. Affleck, Z. Zou, T. Hsu, and P. W. Anderson, Phys.

Rev. B 38, 745 (1988).[37] E. Dagotto, E. Fradkin, and A. Moreo, Phys. Rev. B 38,

2926 (1988).[38] X.-G. Wen, Phys. Rev. B 44, 2664 (1991).[39] P. A. Lee and N. Nagaosa, Phys. Rev. B 45, 5621 (1992).[40] C. Mudry and E. Fradkin, Phys. Rev. B 49, 5200 (1994).

[41] D. P. Arovas and A. Auerbach, Phys. Rev. B 38, 316(1988).

[42] N. Read and S. Sachdev, Phys. Rev. Lett. 62, 1694(1989).

[43] N. Read and S. Sachdev, Phys. Rev. Lett. 66, 1773(1991).

[44] S. Sachdev, Phys. Rev. B 45, 12377 (1992).[45] Z. Y. Weng, D. N. Sheng, Y. C. Chen, and C. S. Ting,

Phys. Rev. B 55, 3894 (1997).[46] S. Sachdev and K. Park, cond-mat/0108214.[47] S. A. Kivelson, D. S. Rokhsar, and J. P. Sethna, Phys.

Rev. B 35, 8865 (1987).[48] D. S. Rokhsar and S. A. Kivelson, Phys. Rev. Lett. 61,

2376 (1988).[49] L. B. Ioffe and A. I. Larkin, Phys. Rev. B 40, 6941

(1989).[50] E. Fradkin and S. Kivelson, Mod. Phys. Lett. B 4, 225

(1990).[51] R. Moessner and S. L. Sondhi, Phys. Rev. Lett. 86, 1881

(2001).[52] G. J. Chen, R. Joynt, F. C. Zhang, and C. Gros, Phys.

Rev. B 42, 2662 (1990).[53] G. Misguich, C. Lhuillier, B. Bernu, and C. Waldtmann,

Phys. Rev. B 60, 1064 (1999).[54] L. Capriotti, F. Becca, A. Parola, and S. Sorella, cond-

mat/0107204.[55] T. Kashima and M. Imada, cond-mat/0104348.[56] R. Coldea, D. Tennant, A. Tsvelik, and Z. Tylczynski,

Phys. Rev. Lett. 86, 1335 (2001).[57] G. Khaliullin and S. Maekawa, Phys. Rev. Lett. 85, 3950

(2000).[58] X.-G. Wen, hep-th/01090120.[59] X.-G. Wen, to appear .[60] L. Ioffe and A. Larkin, Phys. Rev. B 39, 8988 (1989).[61] X.-G. Wen, Phys. Rev. B 60, 8827 (1999).[62] N. Read and B. Chakraborty, Phys. Rev. B 40, 7133

(1989).[63] L. Balents, M. P. A. Fisher, and C. Nayak, Int. J. Mod.

Phys. B 12, 1033 (1998).[64] T. Senthil and M. P. A. Fisher, Phys. Rev. B 62, 7850

(2000).[65] T. Senthil and M. P. A. Fisher, Phys. Rev. Lett. 86, 292

(2001).[66] X.-G. Wen, Quantum Orders and Symmetric Spin Liq-

uids (the original version),http://dao.mit.edu/˜wen/pub/qorder.html (2001).

[67] X.-G. Wen, Phys. Rev. Lett. 84, 3950 (2000).[68] W. Rantner and X.-G. Wen, to appear.[69] H. Yoshizawa, S. Mitsuda, H. Kitazawa, and K. Kat-

sumata, J. Phys. Soc. Jpn. 57, 3686 (1988).[70] R. J. Birgeneau et al., Phys. Rev. B 39, 2868 (1989).[71] S.-W. Cheong et al., Phys. Rev. Lett. 67, 1791 (1991).[72] H. F. Fong, B. Keimer, D. L. Milius, and I. A. Aksay,

Phys. Rev. Lett. 78, 713 (1997).[73] P. Bourges et al., Phys. Rev. B 56, 11439 (1997).[74] K. Yamada et al., Physical Review B 57, 6165 (1998).[75] H. Fong et al., Phys. Rev. B 61, 14773 (2000).[76] S. Wakimoto et al., Phys. Rev. B 61, 3699 (2000).[77] A. M. Polyakov, Nucl. Phys. B 120, 429 (1977).[78] T. Appelquist and D. Nash, Phys. Rev. Lett. 64, 721

(1990).

Page 36: Quantum Orders and Symmetric Spin Liquids - Massachusetts

36

[79] T. Appelquist, A. G. Cohen, and M. Schmaltz, Phys.Rev. D 60, 45003 (1999).

[80] D. H. Kim and P. A. Lee, Annals of Physics 272, 130(1999).

[81] A. V. Chubukov and S. Sachdev, Phys. Rev. Lett. 71,169 (1993).

[82] A. Sokol and D. Pines, Phys. Rev. Lett. 71, 2813 (1993).[83] G. Kotliar, Phys. Rev. B 37, 3664 (1988).[84] Z.-X. Shen et al., Phys. Rev. Lett. 70, 1553 (1993).[85] D. A. Brawner and H. R. Ott, Phys. Rev. B 50, 6530

(1994).[86] J. R. Kirtley et al., Nature 373, 225 (1995).[87] D. J. Scalapino, E. Loh, and J. E. Hirsch, Phys. Rev. B

34, 8190 (1986).[88] D. J. Scalapino, E. Loh, and J. E. Hirsch, Phys. Rev. B

35, 6694 (1987).[89] X.-G. Wen and Y.-S. Wu, Phys. Rev. Lett. 70, 1501

(1993).[90] W. Chen, M. P. A. Fisher, and Y.-S. Wu, Phys. Rev. B

48, 13749 (1993).[91] N. Read and D. Green, Phys. Rev. B 61, 10267 (2000).[92] P. Wiegmann, Princeton preprint IASSNS-HEP-91/29

(unpublished) (1991).

[93] A more precise definition of quantum order is given inRef. [11].

[94] The Landau’s theory may not even be able to describeall the classical orders. Some classical phase transitions,such as Kosterliz-Thouless transition, do not change anysymmetries.

[95] In this paper we will distinguish the invariance of anansatz and the symmetry of an ansatz. We say an ansatzhas a translation invariance when the ansatz itself doesnot change under translation. We say an ansatz has atranslation symmetry when the physical spin wave func-tion obtained from the ansatz has a translation symme-try.

[96] In his unpublished study of quantum antiferromagnetismwith a symmetry group of large rank, Wiegmenn[92] con-structed a gauge theory which realizes a double valuedmagnetic space group. The double valued magnetic spacegroup extends the space group and is a special case ofprojective symmetry group.

[97] We need to integrate out the phase of the φ field to geta gauge invariant result.