Sunteți pe pagina 1din 6

Spin-supersolid vs phase separation in frustrated antiferromagnets

A. Fabricio Albuquerque,1, 2 Nicolas Laorencie,3 Jean-David Picon,4, 5 and Fr d ric Mila5 e e


Laboratoire de Physique Th orique, CNRS and Universit de Toulouse, F-31062 Toulouse, France e e 2 School of Physics, The University of New South Wales, Sydney, NSW 2052, Australia 3 Laboratoire de Physique des Solides, Universit Paris-Sud, UMR-8502 CNRS, 91405 Orsay, France e 4 Theoretische Physik, ETH Z rich, 8093 Z rich, Switzerland u u 5 Institute of Theoretical Physics, Ecole Polytechnique F d rale de Lausanne, Switzerland e e We investigate the possibility to stabilize spin-supersolidity in a frustrated S = 1/2 Heisenberg model of weakly coupled dimers, with special emphasis on its competition with phase separation. This study is based on effective models derived with the contractor renormalization (CORE) algorithm and studied with a selfconsistent cluster mean-eld theory, complemented by an investigation of phase separation based on a leapfrog mechanism. A region of stable supersolidity is identied at intermediate frustration. However, at strong frustration, supersolidity is found to be unstable towards phase separation, in contradiction with recent results based on a tensor-product algorithm. The reliability of the present approach is demonstrated in the context of a nonfrustrated, anisotropic version of the model for which extensive Quantum Monte Carlo simulations have been performed. Finally, to guide the search for experimenral realizations of supersolidity in Mott insulators, the full phase diagram as a function of frustation and magnetic eld is derived.
PACS numbers: 75.10.Jm, 03.75.Nt, 05.30.Jp
1

arXiv:1005.0638v1 [cond-mat.str-el] 4 May 2010

Introduction Dimer-based antiferromagnets (DAFs) are promising candidates for displaying new phases of bosonic matter [1]. Magnetic excitations in such systems, termed triplons, are well described by lattice models of interacting bosons, whose density can be nely tuned by the application of magnetic eld [24]. Experimentally, eld-induced Bose-Einstein condensation (BEC) of triplons has been observed in a number of DAFs (see the review [1]) and, remarkably, exotic quantum criticality has been detected in the spin-dimer compound BaCuSi2 O6 [5, 6]. The presence of magnetic frustration further adds to the rich phenomenology of these systems by enhancing repulsive interactions between triplons, something that may eventually stabilize incompressible phases that break the lattice translational symmetry [3, 7]. Such crystalline phases are for instance realized in the Shastry-Sutherland material SrCu2 (BO3 )2 [8, 9], where they are signaled by a series of magnetization plateaux at unconventional llings stabilized by complex triplon interactions [10, 11]. The occurrence of both BEC and solid phases in the phase diagram of DAFs under magnetic eld suggests that the magnetic equivalent of the phase simultaneously displaying diagonal and off-diagonal order known as supersolid (SS) [12 14] may be realized in these systems. Indeed, insofar as more exotic possibilities are excluded [15], according to the Ginzburg-Landau-Wilson paradigm a continuous transition between phases breaking different symmetries (as it is the case with BEC and crystalline phases) is precluded and we are therefore left with two possibilities: (i) a rst-order transition or (ii) the appearance of an intermediate phase, such as a SS, where both order parameters coexist. The latter possibility has been recently explored in numerical investigations of Heisenberg-like Hamiltonians on dimerized lattices that have indeed been conrmed to display spin-supersolid (spin-SS) phases in the vicinity of magnetization plateaux, characterized by nite values for both spin-stiffness and checkerboard-solid (CBS) order parameter [1619].

The existence of a spin-SS phase for the strongly anisotropic spin-half model investigated in [16, 17] is to some extent surprising. Indeed, one may naively expect this system to map onto the simplest model of hard-core bosons on the square lattice, including only nearest-neighbor (NN) hopping amplitude t1 and interaction V1 (the so-called t V model), that has been shown to be unstable against phase separation (PS) and to display a rst-order transition between condensate and CBS phases [20]. It is thus natural to conjecture that extra terms not included in the simplest t V model, such as correlated hoppings [21], must account for SS behavior in this system. In a previous work [22], we have employed the Contractor Renormalization (CORE) algorithm [23, 24] in order to derive effective Hamiltonians for triplons in the model of [16, 17]. We have shown that the correlated hopping processes with amplitude s1 and s2 depicted in Fig. 1(b-c) allow for holes (singlets, absence of triplons) introduced into a CBS background to delocalize [Fig. 1(d)] and thus suppress phase separation, in what we have termed leapfrog mechanism for spin-SS [22]. Although the model studied in [16, 17, 22] is interesting from a theoretical perspective, since the absence of a signproblem allows for extensive quantum Monte Carlo (QMC) simulations, its strong anisotropic character makes it unrealistic for Mott insulators. More promising in this respect are SU(2) invariant frustrated systems where magnetization plateaux are stabilized by frustrating interactions and where supersolidity may emerge for realistic couplings in the Hamiltonian [18, 19]. In this Letter, we investigate the possibility to stabilize supersolidity in an isotropic spin-1/2 frustrated DAF. We pay special attention to the competition with phase separation and provide strong evidence that, although supersolidity is indeed unstable towards PS for very strong frustration, there exists an intermediate frustration regime in which supersolidity appears to be a stable phase between a superuid and a CBS. Since our results are in contradiction with those obtained re-

2
(a) (b) ~ s1 ~ s
2

J J J
(c)

(d)

(e)

FIG. 1: (Color online) (a) Antiferromagnetic bilayer investigated in this paper [Eq. (1)], with couplings: J (thick vertical lines), J (thiner in-layer lines) and J (dashed lines). A magnetic eld h promotes singlets (vertical pairs of lled circles) to triplets (pairs of open circles); a CBS conguration at half-lling is depicted. (b-c) Correlated hoppings behind the leapfrog mechanism for supersolidity [holes hop in between red and light-blue sites only if the dark circles are occupied by holes in (b), at least one of the sites must be occupied; if both are, the amplitude is 21 ], that allow extra holes s to delocalize in a CBS background by leapfrogging on the other sublattice (d). (e) N = 2 2 cluster for SCMFT: interactions (thick black lines) involving only in-cluster sites (dark-lled circles) are treated exactly while couplings to the environment (grey lines) in a MF way. Only NN bonds are depicted but the effective model from CORE also includes longer-ranged terms.

cently on the same model with the help of a tensor-product algorithm by Chen et al. [19], who reported supersolidity in a parameter range where we nd PS, we have tested the ability of our method to properly account for the competition between supersolidity and PS by comparing its predictions for the unfrustrated, anisotropic model of Ref. [16, 17, 22] with QMC simulations, with very convincing results. Finally, the full phase diagram as a function a frustration and magnetic eld is derived. 1 Effective model We consider the frustrated spin- 2 Hamiltonian dened on a bilayer [Fig. 1 (a)] H=
i,j =1,2

is invariant under the transformation J J , with the consequence that the models phase diagram is symmetric about the line J = J . In studying the model of Eq. (1), we adopt an approach similar to the one employed in our preceding work [22] and derive an effective bosonic model by relying on the CORE algorithm [23, 24]. We consider spin-dimers connected by J [Fig. 1(a)] as elementary blocks and select the singlet 1 |s = 2 [| | ] and the S z = +1 triplet |t+ = | as the block states in the CORE expansion. For all parameters in Eq. (1) considered in the present work, J , J [0, 0.5], this choice is justied by the large reduced density-matrix weights associated to such block states and by the rapid convergence of effective couplings for increasing range in the expansion [25]. Effective couplings are derived by diagonalizing clusters of coupled spin-dimers and by projecting a matching number of low-lying cluster eigenstates onto the basis formed by tensor products of the retained block states, |s and |t+ [26]. The effective bosonic Hamiltonian obtained in this way is essentially identical to the one derived for the model studied in [22], only the magnitudes for each coupling being different. Similarly, the effective model obtained here is not invariant under particle-hole transformation and, in particular, amplitudes for leapfrog processes are non-zero only when holes are involved [22]. From this last observation we expect that only hole-doped SS phases can be stabilized in the spin-dimer model Eq. (1) and conclude that the effective Hamiltonian is more conveniently expressed in terms of hole operators ( n = n is the occupation number n bn b for holes at the dimer-lattice site rn ). We therefore adopt the same notation as in our previous work [22], to which the reader is referred for a complete list of single- and multi-hole interactions and hopping processes [see Eqs. (8, B1-B5) in [22]]. We also remark that the effective Hamiltonian preserves the symmetry of the original model Eq. (1) and remains invariant under J J . Mean-eld analysis Effective Hamiltonians resulting from CORE are often complex and different strategies may be pursued in trying to extract physically sound results from them. One possibility is the mean-eld (MF) theory of Ref. [22], which reproduces semi-quantitatively the results of quantum Monte Carlo (QMC) simulations [16, 17] for the anisotropic spin-dimer model. However, MF calculations are known to overestimate the extent of SS phases and it would be desirable to include, at least partially, effects due to quantum uctuations. From this perspective, the self-consistent cluster mean-eld theory (SCMFT) [27] which takes local quantum uctuations into account and that has been recently applied to the t V model for hard-core bosons on the triangular lattice [28] seems particularly well suited for our purposes. Indeed, the extent of the SS phase in the ground-state phase diagram obtained by applying SCMFT to this model [28] compares considerably better with results from QMC simulations [29] than what is found from a more conventional MF approach [30]. SCMFT is applied by diagonalizing the effective CORE Hamiltonian on the N = 2 2 cluster depicted in Fig. 1(e).

J Si, Sj, + J (Si,1 Sj,2 + Si,1 Sj,2 )


z Si, =1,2

+
i

J Si,1 Si,2 h

. (1)

i, j denotes NN sites in each square layer of the frustrated bilayer depicted in Fig. 1(a). J couples spins in different layers to build the basic dimers of the model (we henceforth set J = 1). The applied magnetic eld h acts as a chemical potential, promoting spin-dimers from a singlet (hole) to a triplet (triplon) state. Effective interactions appear as the result of inlayer J and frustrating J antiferromagnetic couplings. We remark that the lattice depicted in Fig. 1(a) remains invariant if every other spin-dimer is rotated by and therefore Eq. (1)

3
0.5 0.4 0.3 0.2 0.1 0 0.4 0.6 0.8 0.5 0.4 0.3 0.2 0.1 0 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 h 2 2.2 2.4 2.6
(b)
J|| = 0.38 J = 0.21

1.6
M0

S (, ) 0 m
z

BEC

M1

1.4 1.2 hCBS, hSS


J|| = 0.38 J = 0.15

hCBS - CORE-SCMFT hSS - CORE-SCMFT hCBS - Chen et al. hSS - Chen et al.

0.3
CORE-SCMFT Chen et al.

0.25 0.2 0.15 0.1 0.05

(a)

0.8 0.6 0.4

1.2 1.4 1.6 1.8 h


SS CBS

2.2 2.4 2.6

M0

BEC

BEC

M1

0.2 0 0

(a)
0.2 0.4 0.6 0.8 J / J|| 1 0

(b)

J|| = 0.38
0 1

0.2 0.4 0.6 0.8 J / J||

FIG. 2: (Color online) SCMFT results for the condensate density 0 [squares, Eq. (2)], CBS structure factor S(, ) [circles, Eq. (3)] and magnetization density [triangles, Eq. (4)] for the effective CORE Hamiltonian for Eq. (1) with couplings (J, J ) considered in [19]: (a) (0.38, 0.15) and (b) (0.38, 0.21). Successive phases for increasing magnetic eld h are labeled as: spin-gapped (M0 ), condensate (BEC), supersolid (SS), checkerboard solid (CBS) and fully polarized (M1 ).

FIG. 3: (Color online) J = 0.38. (a) Extent of CBS (hCBS ) and SS (hSS ) phases [maximum minus minimum value of the eld h leading to the corresponding phase for given parameters (J, J )]. (b) Value of the structure factor [Eq. (3)] at the CBS plateau. In (a) and (b), symbols indicate results by Chen et al. [19] and lines the here obtained results.

In setting the clusters Hamiltonian, in-cluster interactions are treated exactly while couplings to the environment in a selfconsistent way: for instance, a given interaction connecting sites rm and rn contributes a term proportional to nm nn for each in-cluster bond [thick black lines in Fig. 1(e)] and with mean-eld terms of the form [ m nn + nm nn ] for n bonds connecting the cluster to its environment [grey lines in Fig. 1(e)]. At each step, the ground-state for the cluster Hamiltonian is calculated and expectation values nm , m at every site rm computed; these are then used in setb ting the mean elds for the next iteration, until convergence is achieved (see [28] for details). In this way, we compute the condensate density 1 0 = N
2

(1)
n

rn

n b

(2)

the CBS structure factor (normalized per site) S(, ) = 1 N2 (1)rm rn nm nn ,


m,n

(3)

and the magnetization along the eld direction mz = 1 2N (1 nn ) .


n

(4)

In Fig. 2 we plot these quantities as a function of the magnetic eld h for couplings considered in [19], (J , J ) =

(0.38, 0.15) and (0.38, 0.21). We rst notice that the overall agreement between our results and the data presented in [19] is remarkably good [31]. For the least frustrated case of (J , J ) = (0.38, 0.15) [Fig. 2(a)], the system rst undergoes a quantum transition from a spin-gapped (equivalent to a bosonic Mott insulator with zero-lling for triplons, M0 in Fig. 2) to a BEC phase at the lower critical eld hc1, and then from the BEC to a fully polarized phase (Mott insulator with unitary triplon lling, M1 ) at the upper critical eld hc2. More interestingly, additional CBS and SS phases are stabilized for the more frustrated case of (J , J ) = (0.38, 0.21) [Fig. 2(b)]. The existence of a SS phase at the low-eld boundary of the CBS plateau, with nite values for both 0 and S(, ) is at least partially due to the presence of correlated hoppings for holes in the effective CORE Hamiltonian. Indeed, no SS phase is observed for an effective model obtained by setting s1 = s2 = 0 [Fig. 1(b-c)] while keeping all the other effective couplings unchanged (see also the discussion below). This situation is to be contrasted with the rst-order transition from CBS to BEC at higher elds, explained by the vanishing amplitudes for correlated hoppings for triplons for all values J , J [0, 0.5]. We further explore the phase diagram of the system by varying the frustrating coupling J while xing J = 0.38. In Fig. 3(a) we plot the extent of the SS, CBS phase, respecmax min tively hSS,CBS = hmax hmin SS,CBS SS,CBS [hSS,CBS (hSS,CBS ) denotes the upper (lower) boundary of the SS, CBS phase] as a function of J /J . In order to gauge the accuracy of the here employed CORE-SCMFT approach in Fig. 3(a) our results for hSS and hCBS are compared against those from [19] and in Fig. 3(b) we plot both our results and those from [19] for the the structure factor S(, ) at the CBS plateau. Excellent agreement is found in both cases and we further remark that our results for S(, ) in the CBS phase in Fig. 3(b) conrm that quantum uctuations are indeed partially taken into account by SCMFT: in contrast to what

S ( , )

J|| = 0.38

4
0.5

hSS
> 0.25
0.20 0.15

0.4 0.3 0.2

0.4

0.10 0.3 0.05


0.1

< 0.25
0.2

(b) - J|| = 0.38 (c) - J|| = 0.38

0.00

0 12 8

(a)
0 0 0.1 0.2 0.3

> 0.25
0.4 0.5
0 0.5

4 0 1

J / J||

FIG. 4: (Color online) (a) SCMFT results for the extent of the SS phase hSS (see main text) for the frustrated DAF Eq. (1). The symmetry J J has been explored in obtaining the data. The continuous red curve indicate couplings for which = 0.25 and the vertical dashed line highlights J = 0.38. (b) versus J /J for J = 0.38. The horizontal red line indicates = 0.25. (c) Ratio between NN repulsion V1 and hopping t1 versus J /J for J = 0.38. In all panels, circles indicate couplings investigated by Chen et al. [19].

happens with the semi-classical MF approach employed in [22], here the value of S(, ) at the plateau is somewhat reduced from its classical value Sclassical (, ) = 1/4. Phase diagram and phase separation We turn now our attention to the obtention of a global J - J phase diagram that may guide the experimental search for realizations of spin-supersolidity and therefore extend our analysis by varying J in Eq. (1). In Fig. 4(a) we plot hSS as a function of J , J [0, 0.5]. These results suggest that, far from being a rare occurrence, spin-supersolidity is widespread throughout the parameter space and can extend over fairly wider ranges of h than it is observed for the value J = 0.38 [highlighted by the vertical dashed line in Fig. 4(a)] considered in [19]. However, some caution is required in drawing conclusions from the results presented in Fig. 4(a). Indeed, the emergence of a SS in a system of hard-core bosons on the square lattice depends on the interplay between correlated hoppings that favor supersolidity [21, 22] and the instability of the system toward PS [32]. In addressing this point, we rely on the strong coupling analysis by Sengupta et al. [32] that shows that the energetic gain behind domain wall formation in the hard-core t V model close to half lling is proportional to the NN hopping amplitude, ct1 (1 < c < 2). As discussed in [22], it is thus useful to introduce the leapfrog ratio 2|1 | + |2 | s s = . |t 1 | (5)

0.1

interacting regime, for the ratio between the energetic gains associated to the leapfrog mechanism, that stabilizes the SS, and to domain wall formation, that precludes its existence. Following this argument, we expect that the approximate condition c/4 (1 < c < 2) is to be fullled for supersolidity to emerge in models of hard-core bosons on the square lattice. is readily obtained from the CORE effective couplings and values of J , J leading to = 0.25, the threshold value for a SS phase to appear (assuming c = 1), are indicated by the continuous red curve in Fig. 4(a). We notice that not all values of J , J yielding a spin-SS phase within our CORE-SCMFT approach fulll 0.25 and, in particular, the couplings considered in [19] for which a SS phase is observed (circles in Fig. 4) have an associated leapfrog ratio < 0.25 [see Fig. 4(b)]. Although our simple estimate for the threshold value for can not be expected to be accurate (however, see below), we remark that a SS phase is obtained from our CORE-SCMFT approach even for couplings J J , where our strong coupling analysis applies [the ratio between NN repulsion V1 and hopping 1 diverges toward the line J = J ; see Fig. 4(c), amplitude t where V1 /t1 is plotted for J = 0.38] and PS is expected to occur instead [33]. This inconsistency indicates that SCMFT is insensitive to the instability toward PS in systems of hard-core bosons on the square lattice [33]. In circumventing this limitation of SCMFT we assume that two conditions must be simultaneously fullled for SS phases to exist: (i) a SS must be observed within CORE-SCMFT and (ii) the leapfrog ratio must satisfy > 0.25. Despite of the fact that the condition > 0.25 considerably decreases the size of the region expected to support SS phases from a pure SCMFT analysis, supersolidity is still observed for a wide range of couplings in Eq. (1) [Fig. 4(a)]. QMC simulations At this stage, it would naturally be desirable to gauge the validity of the criterion based on the leapfrog ratio for the stability of the SS phase against PS. To do this, we explore the aforementioned similarity between the physics displayed by the present frustrated model [Eq. (1)] and the anisotropic model analyzed in [16, 17] (see also our previous work [22]), that reads Hxxz = J
i,j =1,2 y y x x z z Si, Sj, + Si, Sj, + Si, Sj,

V1 / t1

+
i

J Si,1 Si,2 h

z Si, =1,2

. (6)

4 (2|1 | + |2 |) is the ground-state energy of a single hole s s doped into a frozen CBS, hopping via the correlated processes with amplitude s1 and s2 [Fig. 1(b-c)]. Therefore, 4 /c is a simple estimate, expected to hold in the strongly

The absence of frustration in Eq. (6), for which a CBS plateau is instead stabilized by the anisotropy , allows one to perform large scale QMC simulations [16, 17]. In [16, 17], couplings J /J = 0.29 and = 3.3 were analyzed; here, we extend the analysis by varying J in the range 0.14 < J /J < 0.4. QMC results for the longitudinal magnetization mz obtained for bilayers of N = 2 16 16 spins at the inverse temperature = 64J are shown in

5
. J = 2.5 . J = 3 . J = 3.5 . J = 4 . J = 4.5 . J = 5 . J = 6 . J = 7
0.4

magnetization mz

0.25

Eq. (6) from our previous work [22] and is plotted in the inset of Fig. 5. Surprisingly, the aforementioned analysis that predicts PS if < 0.25 and a SS phase for > 0.25 exactly coincides with the QMC data in Fig. 5, validating the criterion discussed in the previous paragraph in the case of Eq. (6). Discussion and summary Having checked, by performing QMC simulations, the validity of the criterion > 0.25 for the unfrustrated model Eq. (6), we can comment further on the frustrated model Eq. (1) and compare our results to those otained by Chen et al. [19]. In Fig. 6 we plot the phase diagram h J for J = 0.38 [vertical dashed line in Fig. 4(a)], as obtained from CORE-SCMFT. Motivated by the excellent agreement between our strong coupling analysis and the QMC data for Eq. (6), we assume that the obtention of a SS phase for < 0.25 (above the horizontal dashed line in Fig. 6, that indicates = 0.25) is an artifact of SCMFT and that PS is instead likely to take place under these circumstances. Intriguingly, this assumption contradicts Chen et al. [19], who have observed a SS phase for couplings leading to < 0.25 [see Fig. 4(b)]. While we cannot exclude the possibility that our criterion > 0.25 is too stringent for the frustrated model Eq. (1), it would be important to further test the ability of the algorithm employed in [19] to detect PS in bosonic lattice models. It would for instance be very interesting to see how it compares with QMC for the unfrustrated, anisotropic model regarding the competition between supersolidity and phase separation. Summarizing, we have studied a spin-half frustrated bilayer model by combining CORE and SCMFT. Our results reveal the presence of a spin-supersolid phase under an applied magnetic eld, which appears at the edge of a half-saturated magnetization plateau and is stabilized by a leapfrog mechanism. The interplay between supersolidity and instability toward phase separation has been characterized through a criterion involving a quantity we term leapfrog ratio and whose validity is supported by QMC results for an alternative unfrustrated model. We obtain a global phase diagram for frustrated spin-dimer antiferromagnets where a supersolid phase occur for couplings realizable in real magnets and expect that our results may guide the experimental search for systems exhibiting spin-supersolidity.
Leapfrog Ratio

0.3

0.2

0.1

10

magnetic eld h
FIG. 5: (Color online) QMC results for the magnetization mz of the anisotropic bilayer XXZ model [Eq. (6)] for N = 2 16 16 spins and inverse temperature = 64J for the various couplings J indicated on the plot. Inset: Leapfrog ratio versus J , as obtained from CORE.
1 0.9

CBS
0.8

< 0.25
PS

J / J||

0.7 0.6 0.5 0.4 0.3 1

SS

= 0.25
BEC
1.5 1.75 2 2.25

J|| = 0.38
1.25

> 0.25
2.5

FIG. 6: (Color online) Phase diagram obtained from CORE-SCMFT for effective Hamiltonian for Eq. (1) with J = 0.38. Phases are labelled as in Fig. 2. Large circles on the right indicate couplings analyzed in [19] and the horizontal dashed line indicate the value of J for which = 0.25.

Fig. 5. The occurrence of PS, signaled by a discontinuous jump in the magnetization curve, is clear for J 5. On the other hand, for J 3 the half-saturated plateau is absent and no SS or CBS phases are observed. Supersolidity is achieved in the intermediate range of 3 J 5, as evidenced by simultaneously nite values for the spin-stiffness and CBS structure factor (not plotted here). The leapfrog ratio is readily obtained from the CORE expansion for

Acknowledgments

We acknowledge fruitful discussions with C. D. Batista and M. Troyer, as well as funding from the French ANR program ANR-08-JCJC-0056-01, from ARC (Australia), from the SNF and from MaNEP.

[1] T. Giamarchi, C. R egg, and O. Tchernyshyov, Nat. Phys. 4, u 198 (2008). [2] K. Totsuka, Phys. Rev. B 57, 3454 (1998).

[3] F. Mila, Eur. Phys. J. B 6, 201 (1998). [4] T. Giamarchi and A. M. Tsvelik, Phys. Rev. B 59, 11398 (1999). [5] S. E. Sebastian, N. Harrison, C. D. Batista, L. Balicas,

6
M. Jaime, P. A. Sharma, N. Kawashima, and I. R. Fisher, Nature 441, 617 (2006). N. Laorencie and F. Mila, Phys. Rev. Lett. 102, 060602 (2009). M. Takigawa and F. Mila, in Introduction to Frustrated Magnetism, edited by C. Lacroix, P. Mendels, and F. Mila (Springer, 2010). H. Kageyama, K. Yoshimura, R. Stern, N. V. Mushnikov, K. Onizuka, M. Kato, K. Kosuge, C. P. Slichter, T. Goto, and Y. Ueda, Phys. Rev. Lett. 82, 3168 (1999). K. Kodama, M. Takigawa, M. Horvati , C. Berthier, c H. Kageyama, Y. Ueda, S. Miyahara, F. Becca, and F. Mila, Science 298, 395 (2002). J. Dorier, K. P. Schmidt, and F. Mila, Phys. Rev. Lett. 101, 250402 (2008). A. Abendschein and S. Capponi, Phys. Rev. Lett. 101, 227201 (2008). A. F. Andreev and I. M. Lifshitz, Sov. Phys. JETP 29, 1107 (1969). E. Kim and M. H. W. Chan, Nature 427, 225 (2004). E. Kim and M. H. W. Chan, Science 305, 1941 (2004). T. Senthil, A. Vishwanath, L. Balents, S. Sachdev, and M. P. A. Fisher, Science 303, 1490 (2004). K.-K. Ng and T. K. Lee, Phys. Rev. Lett. 97, 127204 (2006). N. Laorencie and F. Mila, Phys. Rev. Lett 99, 027202 (2007). P. Sengupta and C. D. Batista, Phys. Rev. Lett. 98, 227201 (2007). P. Chen, C.-Y. Lai, and M.-F. Yang, Phys. Rev. B 81, 020409 (2010). G. G. Batrouni and R. T. Scalettar, Phys. Rev. Lett. 84, 1599 (2000). [21] K. P. Schmidt, J. Dorier, A. M. L uchli, and F. Mila, Phys. Rev. a Lett. 100, 090401 (2008). [22] J.-D. Picon, A. F. Albuquerque, K. P. Schmidt, N. Laorencie, M. Troyer, and F. Mila, Phys. Rev. B 78, 184418 (2008). [23] C. J. Morningstar and M. Weinstein, Phys. Rev. Lett. 73, 1873 (1994). [24] C. J. Morningstar and M. Weinstein, Phys. Rev. D 54, 4131 (1996). [25] A. Abendschein and S. Capponi, Phys. Rev. B 76, 064413 (2007). [26] In this way, and by imposing that each clusters low-energy spectrum is exactly reproduced, effective couplings of up to range-2 (see Fig. 5 in [22]) are computed. [27] E. Zhao and A. Paramekanti, Phys. Rev. B 76, 195101 (2007). [28] S. R. Hassan, L. de Medici, and A.-M. S. Tremblay, Phys. Rev. B 76, 144420 (2007). [29] S. Wessel and M. Troyer, Phys. Rev. Lett. 95, 127205 (2005). [30] G. Murthy, D. Arovas, and A. Auerbach, Phys. Rev. B 55, 3104 (1997). [31] Similarly good agreement is found for the other couplings considered in [19], (J, J ) = (0.38, 0.23) and (0.38, 0.27). [32] P. Sengupta, L. P. Pryadko, F. Alet, M. Troyer, and G. Schmid, Phys. Rev. Lett. 94, 207202 (2005). [33] A SS phase is obtained from the CORE-SCMFT procedure for arbitrarily small amplitudes for the leapfrog processes depicted in Fig. 1(b-c). First order transitions are only observed when correlated processes vanish, as it is the case for the higher-eld transition out from the CBS plateau in Fig. 2(b) (correlated hoppings for triplons vanish for all J, J [0, 0.5]).

[6] [7]

[8]

[9]

[10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20]

S-ar putea să vă placă și