n eel ordered honeycomb subsystem, coexisting with ...it is in the quantum magnetism arena,...

14
Correlated partial disorder in a weakly frustrated quantum antiferromagnet M. G. Gonzalez, F. T. Lisandrini, G. G. Blesio, A. E. Trumper, C. J. Gazza, and L. O. Manuel Instituto de F´ ısica Rosario (CONICET) and Universidad Nacional de Rosario, Boulevard 27 de Febrero 210 bis, (2000) Rosario, Argentina Partial disorder –the microscopic coexistence of long-range magnetic order and disorder– is a rare phenomenon, that has been experimental and theoretically reported in some Ising- or easy plane- spin systems, driven by entropic effects at finite temperatures. Here, we present an analytical and numerical analysis of the S =1/2 Heisenberg antiferromagnet on the 3 × 3-distorted triangular lattice, which shows that its quantum ground state has partial disorder in the weakly frustrated regime. This state has a 180 eel ordered honeycomb subsystem, coexisting with disordered spins at the hexagon center sites. These central spins are ferromagnetically aligned at short distances, as a consequence of a Casimir-like effect originated by the zero-point quantum fluctuations of the honeycomb lattice. Introduction- Zero-point quantum fluctuations in con- densed systems are responsible for a wide variety of in- teresting phenomena, ranging from the existence of liq- uid helium near zero temperature to magnetically disor- dered Mott insulators 1,2 . It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum fluctuations. Among these factors, space dimensionality, lattice coor- dination number, spin value S, and frustrating exchange interactions are the most relevant 3,4 . While folk wisdom visualizes zero-point quantum fluc- tuations like a uniform foam resulting from an almost random sum of states, in some cases these fluctuations contribute to the existence of very unique phenomena. These phenomena include semiclassical orders 3 , order by disorder 5 , effective dimensionality reduction 6,7 , and topological orders associated with quantum spin liquid states 1 , among others. Another role for quantum fluc- tuations is to allow the emergence of complex degrees of freedom from the original spins, like weakly coupled clusters or active spin sublattices decoupled from orphan spins. The latter has been proposed to explain the spin liquid behavior of the LiZn 2 Mo 3 O 8 8 . Here the system is described by a triangular spin- 1 2 Heisenberg antiferro- magnet which is deformed into an emergent honeycomb lattice weakly coupled to the central spins. Besides spin liquids, the presence of weakly coupled magnetic subsystems can lead to partial disorder, that is, the microscopic coexistence of long-range magnetic order and disorder. This rare phenomenon has been experimental and theoretically reported in different lo- calized or itinerant Ising- or XY-spin highly frustrated systems 9–25 and it is driven by entropic effects at finite temperature. In general, it is believed that some amount of spin anisotropy is needed to get partial disorder, and that the disordered subsystem behaves as a perfect para- magnet, with its decoupled spins justifying then the cal- ificative of orphan spins. In this work, we present an isotropic frustrated mag- netic system whose ground state exhibits partial disorder, originated by zero-point quantum effects, in contrast to the entropic origin of the so-far known cases. Specifically, we compute the ground state of the S =1/2 antiferro- magnetic Heisenberg model in the 3 × 3-distorted tri- angular lattice (Fig. 1), by means of the linear spin wave theory (LSWT) and the numerically exact density ma- trix renormalization group (DMRG). For the weakly frus- trated 0 J 0 /J . 0.18 range, we find a novel partial dis- order state, without semiclassical analog, that consists in the coexistence of a N´ eel order in the honeycomb sublat- tice and disordered central spins. In addition, the spins of the disordered sublattice are ferromagnetically aligned at short distances, a correlated behavior induced, as we will show, by a Casimir-like effect due to the “vacuum” quantum fluctuations inherent in the quantum N´ eel order of the honeycomb lattice. A B C J J' ϕ B ϕ A FIG. 1: (color online) 3 × 3-distorted triangular lattice, with two different exchange interactions J and J 0 . The arrows correspond to the spin directions of the semiclassical magnetic order. At the heart of the decoupling mechanism is the com- petition between the exchange energy favored by larger coordination numbers and the zero-point quantum fluc- tuations. This can be roughly illustrated by simple toy models. For example, we can resort to the triangle and the hexagon with a central spin of Fig. 2, that present a ground state energy level crossing at J 0 /J = 1 and J 0 /J =0.63, respectively. For small J 0 , the strongly connected spins form singlets, leaving the central spin completely decoupled. Model and methods- We study the S = 1 2 Heisenberg model on the 3 × 3-distorted triangular lattice. Under this distortion, the original triangular lattice is split into arXiv:1804.06720v2 [cond-mat.str-el] 11 Dec 2018

Upload: others

Post on 25-Jun-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

Correlated partial disorder in a weakly frustrated quantum antiferromagnet

M. G. Gonzalez, F. T. Lisandrini, G. G. Blesio, A. E. Trumper, C. J. Gazza, and L. O. ManuelInstituto de Fısica Rosario (CONICET) and Universidad Nacional de Rosario,

Boulevard 27 de Febrero 210 bis, (2000) Rosario, Argentina

Partial disorder –the microscopic coexistence of long-range magnetic order and disorder– is a rarephenomenon, that has been experimental and theoretically reported in some Ising- or easy plane-spin systems, driven by entropic effects at finite temperatures. Here, we present an analytical andnumerical analysis of the S = 1/2 Heisenberg antiferromagnet on the

√3×√

3-distorted triangularlattice, which shows that its quantum ground state has partial disorder in the weakly frustratedregime. This state has a 180◦ Neel ordered honeycomb subsystem, coexisting with disordered spinsat the hexagon center sites. These central spins are ferromagnetically aligned at short distances,as a consequence of a Casimir-like effect originated by the zero-point quantum fluctuations of thehoneycomb lattice.

Introduction- Zero-point quantum fluctuations in con-densed systems are responsible for a wide variety of in-teresting phenomena, ranging from the existence of liq-uid helium near zero temperature to magnetically disor-dered Mott insulators1,2. It is in the quantum magnetismarena, precisely, where a plethora of control factors areavailable for tuning the amount of quantum fluctuations.Among these factors, space dimensionality, lattice coor-dination number, spin value S, and frustrating exchangeinteractions are the most relevant3,4.

While folk wisdom visualizes zero-point quantum fluc-tuations like a uniform foam resulting from an almostrandom sum of states, in some cases these fluctuationscontribute to the existence of very unique phenomena.These phenomena include semiclassical orders3, orderby disorder5, effective dimensionality reduction6,7, andtopological orders associated with quantum spin liquidstates1, among others. Another role for quantum fluc-tuations is to allow the emergence of complex degreesof freedom from the original spins, like weakly coupledclusters or active spin sublattices decoupled from orphanspins. The latter has been proposed to explain the spinliquid behavior of the LiZn2Mo3O8

8. Here the systemis described by a triangular spin- 12 Heisenberg antiferro-magnet which is deformed into an emergent honeycomblattice weakly coupled to the central spins.

Besides spin liquids, the presence of weakly coupledmagnetic subsystems can lead to partial disorder, thatis, the microscopic coexistence of long-range magneticorder and disorder. This rare phenomenon has beenexperimental and theoretically reported in different lo-calized or itinerant Ising- or XY-spin highly frustratedsystems9–25 and it is driven by entropic effects at finitetemperature. In general, it is believed that some amountof spin anisotropy is needed to get partial disorder, andthat the disordered subsystem behaves as a perfect para-magnet, with its decoupled spins justifying then the cal-ificative of orphan spins.

In this work, we present an isotropic frustrated mag-netic system whose ground state exhibits partial disorder,originated by zero-point quantum effects, in contrast tothe entropic origin of the so-far known cases. Specifically,we compute the ground state of the S = 1/2 antiferro-

magnetic Heisenberg model in the√

3×√

3-distorted tri-angular lattice (Fig. 1), by means of the linear spin wavetheory (LSWT) and the numerically exact density ma-trix renormalization group (DMRG). For the weakly frus-trated 0 ≤ J ′/J . 0.18 range, we find a novel partial dis-order state, without semiclassical analog, that consists inthe coexistence of a Neel order in the honeycomb sublat-tice and disordered central spins. In addition, the spinsof the disordered sublattice are ferromagnetically alignedat short distances, a correlated behavior induced, as wewill show, by a Casimir-like effect due to the “vacuum”quantum fluctuations inherent in the quantum Neel orderof the honeycomb lattice.

ABCJJ'

ϕBϕA

FIG. 1: (color online)√

3 ×√

3-distorted triangular lattice,with two different exchange interactions J and J ′. The arrowscorrespond to the spin directions of the semiclassical magneticorder.

At the heart of the decoupling mechanism is the com-petition between the exchange energy favored by largercoordination numbers and the zero-point quantum fluc-tuations. This can be roughly illustrated by simple toymodels. For example, we can resort to the triangle andthe hexagon with a central spin of Fig. 2, that presenta ground state energy level crossing at J ′/J = 1 andJ ′/J = 0.63, respectively. For small J ′, the stronglyconnected spins form singlets, leaving the central spincompletely decoupled.

Model and methods- We study the S = 12 Heisenberg

model on the√

3×√

3-distorted triangular lattice. Underthis distortion, the original triangular lattice is split into

arX

iv:1

804.

0672

0v2

[co

nd-m

at.s

tr-e

l] 1

1 D

ec 2

018

Page 2: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

2

FIG. 2: Hexagon and triangle toy models with one spin con-nected to the remaining ones by a different exchange interac-tion J ′ (dashed lines).

a honeycomb subsystem and a sublattice of spins at thecenter of each hexagon (see Fig. 1). Naturally, two dif-ferent nearest neighbor exchange interactions arise, andthe Heisenberg Hamiltonian turns out

H = J∑〈ij〉

Si · Sj + J ′∑[ik]

Si · Sk, (1)

where 〈ij〉 runs over nearest neighbor spins belonging tothe honeycomb lattice (with equivalent sublattices A andB), while [ik] links the honeycomb and central spins C,that interact with energy J ′. Throughout this work, wetake J = 1 as the energy unit, while J ′ is the only varyingparameter, which we consider in the range [0, 1].

Almost two decades ago, this model was proposedin the context of the honeycomb reconstruction ofthe metallic surface of Pb/Ge(111)26. On the otherhand, the complementary range J ′ ≥ 1 has beenconsidered27,28, and very recently the uniform magneti-zation in the regime J ′ ≤ 1 has been computed by exactdiagonalization29.

Related stacked triangular lattice XY-antiferromagnets have been extensively studied inthe context of the magnetic properties of hexagonalABX3 compounds with space group P63cm

12,19,30.This model has two very well known limits: (i) for

J ′ = 1, we recover the Heisenberg model on the isotropictriangular lattice, with its three equivalent sublatticesand a 120◦ Neel ordered ground state31,32; while (ii) forJ ′ = 0, we have a honeycomb Heisenberg model withits 180◦ Neel ordered ground state33, and orphan (com-pletely decoupled) spins at the centers of the hexagons.

The classical ground state of (1) is a simple three-sublattice order34, as depicted in Fig. 1, characterizedby the magnetic wave vector Q = 0 and by the anglesφA = −φB = − arccos(−J ′/2) that the spin directions onsublattices A and B make with the spin direction in sub-lattice C. This ground state evolves continuously fromthe honeycomb (plus orphan C spins) to the isotropic tri-angular classical ground states, and it is a ferrimagnet for0 < J ′ < 1. The Lacorre parameter35, whose departurefrom the unity quantifies the degree of magnetic frustra-tion, is (2+J ′2)/(2+4J ′), so the maximal frustrated casecorresponds to the isotropic triangular lattice.

In this work, we solve (1) by means of complementaryanalytical and numerical techniques –the semiclassical

linear spin wave theory34 and the density matrix renor-malization group36–, in order to highlight the quantumbehavior without classical counterpart of the model. TheDMRG calculations were performed on ladders of dimen-sion Lx × Ly37, with Ly = 6 and Lx up to 15, imposingcylindrical boundary conditions (periodic along the y di-rection). We use up to 3000 DMRG states34 in the mostunfavorable case to ensure a truncation error below 10−6

in our results.Linear spin wave results- LSWT yields the same

ground state magnetic structure as the classical one (seeFig. 1), with a semiclassically renormalized local mag-netization mα (α = A,B,C) for each sublattice34, dis-played in the inset of Fig. 3. As the A and B sub-lattices are equivalent, their order parameters coincide,while they are different from the central spin local magne-tization mC . For J ′ = 0, the central spins are decoupledfrom the honeycomb lattice and, consequently, mC cantake any value from 0 to 1/2. As soon as J ′ is turned on,the sublattice C takes a large magnetization value, morethan 80% of its classical value, while mA varies continu-ously.

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.2 0.4 0.6 0.8 1.0

m

J'

mAmC

0.00

0.25

0.50

0.0 0.2 0.4 0.6 0.8 1.0

m

J'

FIG. 3: (color online) Local magnetization of sublattices Aand C as a function of J ′, calculated with DMRG (main panel,where lines are merely a guide to the eye) and with LSWT(inset). The DMRG results correspond to the Ly = 6, Lx =12 cluster.

Another interesting feature that can be seen is that theincrease of the frustrating interaction J ′ leads to an en-hancement of the local magnetization in the honeycomblattice, up to a broad maximum around J ′ = 0.35 (seethe darker curve in the inset of Fig. 3). This (appar-ent) paradoxical result can be explained by the increaseof the effective coordination number induced by J ′, thatdrives the system closer to its classical behavior. Al-ternatively, it can be thought that, as J ′ is turned on,the honeycomb spins feel the C subsystem as an uni-

Page 3: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

3

form Weiss magnetic field B = mCJ′ that, through the

suppression of quantum fluctuations, contributes to theincrease of the local magnetization mA (= mB), as it wasfound in other frustrated systems38. It is worth to no-tice that, for any J ′, the larger order parameter belongto the sublattice C, which can be considered to be thesublattice with the smaller effective coordination num-ber, zCeff ≈ 6J ′/J ≤ zAeff ≈ 3 + 3J ′/J. This is in agree-ment with the fact that in lattices with inequivalent sitesor bonds, the order parameter is lower in the sites withlarger coordination numbers39.

DMRG results- For all the considered range 0 ≤ J ′ ≤ 1,the computed spin correlations 〈Si · Sj〉 exhibit a threesublattice pattern, in full agreement with the semiclassi-cal approach. Thus, if a given sublattice is ordered, allits spins will point out in the same direction (ferromag-netic order) and its local magnetization can be evaluatedusing the expression40

m2α =

1

Nα(Nα − 1)

∑i,j∈αi 6=j

〈Si · Sj〉 , (2)

where α denotes the sublattice (A, B, or C), and Nα is itsnumber of sites. The calculated mA and mC are shown inthe main panel of Fig. 3 for the Ns = 12× 6 cluster (forother cluster sizes, we have obtained a similar result34).

The most eye-catching difference between the DMRGand the semiclassical local magnetizations appears in theweakly frustrated parameter region, close to the honey-comb phase, 0 ≤ J ′ . 0.18. There, DMRG shows avanishing order parameter mC for the C sublattice alongwith an almost constant honeycomb lattice local magne-tization mA. This corresponds to a partially disorderedphase, driven solely by quantum fluctuations (in com-petition with frustration), as we are working at T = 0.Notice that, in general, partially disordered phases areassociated with entropic effects, and they appear at in-termediate temperatures, between the lower and higherenergy scales of the system11,20,22.

There is a critical value J ′c ' 0.18 where the C sublat-tice gets suddenly ordered, as it happens at J ′ = 0+ inLSWT (inset of Fig. 3). Furthermore, beyond this crit-ical value, the DMRG calculations show a higher localmagnetization in the central spin sublattice, in agree-ment with the semiclassical expectation39.

The local magnetization mA (= mB) of the honeycombspins decreases when J ′ increases due to the frustrationintroduced by the coupling with the central spins, untilit reaches its minimum value in the isotropic triangularlattice, corresponding to the most frustrated case41. Incontrast with the spin wave results, mA does not exhibit aclear maximum for intermediate values of J ′, but showsan almost constant region ranging from J ′ = 0 to 0.2.This feature signals a negligible effect of the central spinson the honeycomb ones.

As we have mentioned above, the semiclassical groundstate is ferrimagnetic for 0 < J ′ < 1. So, in order tofurther characterize the DMRG ground state magnetic

structure, we calculate the lowest eigenenergy in the dif-ferent Sz subspaces. In the case of J ′ = 0, the spins C aretotally disconnected from the honeycomb subsystem and,thus, do not contribute to the total energy. This resultsin a perfect paramagnetic behavior of the central spinsublattice, with a high degeneracy of the ground state,as E(Sz = 0) = E(Sz = ±1) = ... = E(Sz = Smaxz =±1/2×Ns/3), where Smaxz denotes the maximum valueof Sz whose subspace belongs to the ground state man-ifold, and Ns is the number of sites of the cluster. ForJ ′ 6= 0 the situation changes: for J ′ ≤ J ′c there are nomore orphan spins and Smaxz = 0 (see Fig. S2(a)34),indicating that the partially disordered phase is a (cor-related) singlet. On the other hand, at the critical valueJ ′c, S

maxz jumps to a finite value, signaling a first-order

transition from the singlet to the ferrimagnetic state (seeFig. S2(b)34). With further increase of J ′, Smaxz de-creases until it vanishes for the isotropic triangular pointJ ′ = 1, whose ground state again is a singlet. Therefore,the DMRG magnetic order has a ferrimagnetic charac-ter for J ′c < J ′ < 1. This behavior shows little finitesize effects34 and it quantitatively agrees with the exactdiagonalization predictions29.

Next, we calculate the DMRG angles between the lo-cal magnetization in different sublattices. Previously, wehave checked that the classical angles φα are not renor-malized in LSWT, in contrast to what happens in otherfrustrated models, like the triangular anisotropic Heisen-berg model, where the pitch of its spiral magnetic orderis sizable renormalized by quantum fluctuations42. In or-der to estimate the angles, we take into account that thelarge values of the DMRG local magnetizations enable usto use a semiclassical picture of the spins. Let us thinkof a three-spin unit cell built by spins A, B, and C oflengths mA, mB , and mC , respectively, as seen in Fig. 1.We can assume that the Smaxz subspace corresponds tothe C spin pointing out in the z-direction, while the Aand B spins make angles φA = −φB with it. Then, weget the equation

Ns3

(mC + 2mA cosφB) = Smaxz ,

for φB26, which finally leads to the angle θ between A

and B spins

θ = 2 arccos

[1

2mA

(mC −

3Smaxz

Ns

)]. (3)

In Fig. 4 the angle θ is plot as a function of J ′. It canbe seen that the honeycomb 180◦ Neel order persists allalong the partially disordered phase where mc vanishes.This result is a simple consequence of the singlet char-acter of the ground state for 0 < J ′ < J ′c (Smaxz = 0 inEq. 3). As a consequence, the canting behavior observedsemiclassically for any finite J ′ moves to the region aboveJ ′c in the strong quantum limit S = 1/2 calculated withDMRG. That is to say that, when J ′ is small, C spinsare disordered because the system gains zero-point quan-tum energy from that disorder. For larger values of J ′,

Page 4: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

4

120

140

160

180

0.0 0.2 0.4 0.6 0.8 1.0

θ

J'

DMRGLSWT

FIG. 4: (color online) DMRG and LSWT angles θ betweenspins in the A and B sublattices as a function of J ′.

the system chooses to gain (frustrated) exchange energyover zero-point quantum fluctuation, and the C sublat-tice gets ordered, canting simultaneously the A and Bspins.

It is worth to emphasize that, even when the groundstate seems to undergo a first-order transition at J ′c(Smaxz changes abruptly and the local magnetization mC

sharply rises), the angle θ between the A and B spinsvaries continuously from its 180◦ value in the partiallydisordered phase. This is similar to the spin wave behav-ior around J ′ = 0, where the sublattice C is disordered,but as soon as J ′ rises, mC suddenly grows over mA

without any abrupt change in the magnetic order.The quantitative agreement between the DMRG and

LSWT angles for J ′ & J ′c, displayed in Fig. 4, is a clearevidence that, beyond the partially disordered phasepresent in the weakly frustrated regime, the quantumground state of the model is very well described semi-classically.

Up to now, we have characterized the region betweenJ ′ = 0 and 0.18 as a singlet partially disordered phase.To deepen the understanding of such disorder, in Fig. 5the average nearest (nn) and next-nearest (nnn) neighborspin correlations43 between the central spins is shown asa function of J ′. It can be seen that, even though the sumover all the correlations in the C sublattice is zero in theregion of its null local magnetization (see Eq. 2), the nncorrelation has an almost constant positive value, closeto 1/8, while the nnn correlation is close to zero. Thismeans that central spins exhibit (very) short range fer-romagnetic correlations between them, suggesting thatthe partial disordered phase may be thought as a sort ofa resonating spin-triplet valence bond state44. In othermagnetic systems which exhibit partial disorder, mostlywith Ising or XY-like spins, the disordered subsystem is

a perfect paramagnet of orphan spins, with zero correla-tion between them11,12. As there is no explicit exchangeinteraction between the C spins, their correlation shouldbe mediated by the coexistent honeycomb Neel order.

0.0

0.1

0.2

0.3

0.0 0.2 0.4 0.6 0.8 1.0

<S i

S j>

C

J'

nn Cnnn C

FIG. 5: Average DMRG nearest neighbor (nn) and next-nearest neighbor (nnn) correlations between central spins asa function of J ′.

In order to build up a qualitative argument about theorigin of the correlated character of the partially disor-dered phase below J ′c, we appeal to a Weiss molecularfield approach for the simplest toy model (see Supplemen-tal Material34 for details). We consider a 4-spin cluster,composed of two nearest neighbor honeycomb spins (1and 2) interacting with the two closest C spins (3 and4), with a Hamiltonian

H = JS1 · S2 + J ′ (S1 + S2) · (S3 + S4) .

After fixing the state of honeycomb spins 1 and 2 as a“classical Neel order”, plus zero-point quantum fluctua-tions quantified by a parameter r, we arrive at an effectiveHamiltonian for the central spins 3 and 4, that consistsof a Zeeman term associated with an effective uniformmagnetic field B ∝ rJ ′, perpendicular to the honeycombNeel order. Hence, this toy model helps us to under-stand how the ferromagnetic correlations between near-est neighbor central spins are built up under an effectiveinteraction between them, driven by the vacuum fluctua-tions of the honeycomb Neel order; that is, the correlationbetween central spins can be considered a Casimir-like ef-fect. This argument is further supported by the presenceof nearest-neighbor antiferromagnetic spin-spin correla-tions between the orphan spins and the honeycomb ones(see Ref.34). This treatment is valid whenever this Neelorder is unaffected by the feedback of the spins 3 and4. This seems to be the case in the DMRG calculationsfor the lattice, as the local magnetization mA changes

Page 5: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

5

only slightly by the coupling of the C spins in the par-tially disordered phase (see Fig. 3). Also, the toy modelexplains the almost constant nearest neighbor correla-tion between central spins that can be seen, below J ′c, inFig. 5 (except for the non-monotonic behavior very closeto J ′ = 0, probably due to numerical inaccuracies). Itshould be mentioned that, due to the singlet characterof the DMRG ground state, the correlations between thecentral spins are isotropic, and not perpendicular to agiven direction like in the toy model.

The absence of long-range ferromagnetic order of theC subsystem below J ′c

45 can be roughly explained as fol-lows: in the previous Weiss mean-field argument, thezero-point quantum fluctuations of the honeycomb sub-system act as a magnetic field for the C spins. The direc-tion of this effective molecular magnetic field is ”random”as it depends on the phase fluctuations of the depar-ture from the Neel state. Therefore, while the nearest-neighbor C spins are ferromagnetically correlated, theoverall subsystem remains disordered due to the ”ran-domness” of the zero point quantum fluctuations.

Summary- We have studied a S = 1/2 Heisenberg an-tiferromagnet with inequivalent exchange interactions ona distorted triangular lattice which, in the weakly frus-trated regime 0 ≤ J ′/J . 0.18, exhibits a novel corre-lated partial disordered phase, driven by the competi-tion between zero-point quantum fluctuations and frus-tration. Even if partial disordered phases were alreadyknown9–25–in anisotropic systems at finite temperaturesand with disordered subsystems that behave as perfectparamagnets–, here we uncover, for the first time, a par-tially disordered phase as the ground state of a simpleisotropic Heisenberg antiferromagnet. This phase ex-hibits the coexistence of two magnetic subsystems: oneantiferromagnetically ordered, and the other, disorderedwith its spins ferromagnetically correlated at short dis-tances due to the zero-point quantum fluctuations of theordered subsystem via a Casimir-like effect.

Acknowledgements- We thank Dr. H. Nakano for call-ing our attention to Ref.29. This work was supported byCONICET under grant Nro. 364 (PIP2015).

1 L. Balents, Spin liquids in frustrated magnets, Nature 464,199 (2010).

2 Y. Zhou, K. Kanoda, and T.-K. Ng, Quantum spin liquidstates, Rev. Mod. Phys. 89, 025003 (2017).

3 G. Misguich and C. L’huillier, in Frustrated spin systems,edited by H. T. Diep (World Scientific, New York, 2005).

4 C. Lacroix, P. Mendels, and F. Mila, Introduction toFrustrated Magnetism: Materials, Experiments, Theory(Springer, Berlin, Heidelberg, 2011).

5 C. L. Henley, Ordering due to disorder in a frustrated vectorantiferromagnet, Phys. Rev. Lett. 62, 2056 (1989).

6 Y. Hayashi and M. Ogata, Possibility of gapless spin liq-uid state by one-dimensionalization, J. Phys. Soc. Jpn. 76,053705 (2007).

7 M. G. Gonzalez, E. A. Ghioldi, C. J. Gazza, L. O. Manuel,and A. E. Trumper, One dimensionalization in the spin-1Heisenberg model on the anisotropic triangular lattice, Phys.Rev. B 96, 174423 (2017).

8 R. Flint and P. A. Lee, Emergent Honeycomb Lattice inLiZn2Mo3O8, Phys. Rev. Lett. 111, 217201 (2013).

9 M. Mekata, Antiferro-Ferrimagnetic Transition in Triangu-lar Ising Lattice, J. Phys. Soc. Jpn. 42, 76 (1977).

10 P. Azaria, H. T. Diep, and H. Giacomini, Coexistence oforder and disorder and reentrance in an exactly solvablemodel, Phys. Rev. Lett. 59, 1629 (1987).

11 H. T. Diep, M. Debauche, and H. Giacomini, Exact so-lution of an anisotropic centered honeycomb lattice: Reen-trance and partial disorder, Phys. Rev. B 43, 8759 (1991).

12 M. L. Plumer, A. Caille, and H. Kawamura, Chiral tetra-critical behavior in a partially frustrated triangular antifer-romagnet: Application to ABX3 compounds, Phys. Rev. B44, 4461 (1991).

13 R. Quartu and H. Diep, Partial order in frustrated quan-tum spin systems, Phys. Rev. B 55, 2975 (1997).

14 C. Santamaria and H. T. Diep, Evidence of partial disorderin a frustrated Heisenberg spin system, J. Appl. Phys. 81,5276 (1997).

15 X. G. Zheng, H. Kubozono, K. Nishiyama, W. Higemoto,T. Kawae, A. Koda, and C. N. Xu, Coexistence of long-range order and spin fluctuation in geometrically frustratedclinoatacamite, Phys. Rev. Lett. 95, 057201 (2005).

16 X. G. Zheng, T. Kawae, H. Yamada, K. Nishiwama, andC. N. Xu, Coexisting ferromagnetic order and disorder inuniform system of hydroxyhalide Co2(OH)3Cl, Phys. Rev.Lett. 97, 247204 (2006).

17 S. Agrestini, C. Mazzoli, A. Bombardi, and M. R. Lees,Incommensurate magnetic ground state revealed by resonantx-ray scattering in the frustrated spin system Ca3Co2O6,Phys. Rev. B 77, 140403 (R) (2008).

18 Gia-Wei Chern, R. Moessner, and O. Tchernyshyov, Par-tial order from disorder in a classical pyrochlore antiferro-magnet, Phys. Rev. B 78, 144418 (2008).

19 Y. Nishiwaki, T. Nakamura, A. Oosawa, K. Kakurai, N.Todoroki, N. Igawa, Y. Ishi, and T. Kato, Partial disor-dered phase of Ising spin system in distorted triangular lat-tice RbCoBr3, J. Phys. Soc. Jpn. 77, 104703 (2008).

20 Y. Nishiwaki, A. Oosawa, K. Kakurai, K. Kaneko, M.Tokunaga, and T. Kato, Neutron diffraction study of succes-sive magnetic phase transitions of distorted triangular lat-tice antiferromagnet RbFeBr3, J. Phys. Soc. Jpn. 80, 084711(2011).

21 Q. N. Chen, M. P. Qin, J. Chen, Z. C. Wei, H. H. Zhao,B. Normand, and T. Xiang, Partial Order and Finite-Temperature Phase Transitions in Potts Models on IrregularLattices, Phys. Rev. Lett. 107, 165701 (2011).

22 H. Ishizuka and Y. Motome, Partial disorder in an Ising-spin Kondo Lattice model on a triangular lattice, Phys. Rev.Lett. 108, 257205 (2012).

23 K. Tomiyasu, H. Yasui, and Y. Yamaguchi, Observationof Partial-Disorder-Type Spin Fluctuations in FrustratedMn3Pt, J. Phys. Soc. Jpn. 81, 114724 (2012).

24 H. Ishizuka and Y. Motome, Thermally induced phases inan Ising Kondo lattice model on a triangular lattice: Partialdisorder and Kosterlitz-Thouless state, Phys. Rev. B 87,

Page 6: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

6

155156 (2013).25 B. Javanparast, Z. Hao, M. Enjalran, and M. Gingras,

Fluctuation-driven selection at criticality in a frustratedmagnetic system: The case of multiple-k partial order onthe pyrochlore lattice, Phys. Rev. Lett. 114, 130601 (2015).

26 J.P. Rodriguez and E. Artacho, Metallic surface recon-struction driven by frustrated antiferromagnetism, Phys.Rev. B 59, R705 (1999).

27 H. Nakano and T. Sakai, Ferrimagnetism in the spin-1/2Heisenberg antiferromagnet on a distorted triangular lattice,J. Phys. Soc. Jpn. 86, 063702 (2017).

28 A. Shimada, H. Nakano, T. Sakai, and K. Yoshimura, Spin1/2 triangular lattice Heisenberg antiferromagnet with

√3×√

3-type distortion -behavior around the boundaries of theintermediate phase, J. Phys. Soc. Jpn. 87, 03706 (2018).

29 A. Shimada, T. Sakai, H. Nakano, and K. Yoshimura,Spontaneous magnetization of the spin-1/2 Heisenberg an-tiferromagnet on the triangular lattice with a distortion, J.Phys.: Conf. Ser. 969, 012126 (2018).

30 M. F. Collins and O. A. Petrenko, Triangular antiferro-magnets, Can. J. Phys. 75, 605, 1997.

31 L. Capriotti, A. E. Trumper, and S. Sorella, Long-rangeNeel order in the triangular Heisenberg model, Phys. Rev.Lett. 82, 3899 (1999).

32 S. R. White and A. L. Chernyshev, Neel order in squareand triangular lattice Heisenberg models, Phys. Rev. Lett.99, 127004 (2007).

33 J. Oitmaa, C. J. Hamer, and W. Zheng, Quantum magnetson the honeycomb and triangular lattices at T = 0, Phys.Rev. B 45, 9834 (1992).

34 See Supplemental material for a derivation of the classi-cal ground state of the Hamiltonian, the development ofthe linear spin wave theory, the presentation of a general-ized Weiss molecular field used to interpret the partiallydisordered phase, and several discussion about the DMRGresults.

35 P. Lacorre, The constraint functions: an attempt to evalu-

ate the constraint rate inside structures that undergo orderedmagnetic frustration, J. Phys. C: Solid State Phys. 20, L775(1997).

36 S. R. White, Density matrix formulation for quantumrenormalization groups, Phys. Rev. Lett. 69, 2863 (1992).

37 A. Weichselbaum and S. R. White, Incommensurate cor-relations in the anisotropic triangular Heisenberg lattice,Phys. Rev. B 84, 245130 (2011).

38 B. Schmidt, M. Siahatgar, and P. Thalmeier, Stabilizationof Neel order in frustrated magnets with increasing magneticfield, EPJ Web of Conferences 40, 04001 (2013).

39 A. Jagannathan, R. Moessner, and S. Wessel, Inhomo-geneous quantum antiferromagnetism on periodic lattices,Phys. Rev. B 74, 184410 (2006).

40 H. Nakano, S. Todo, and T. Sakai, Long-range order of thethree-sublattice structure in the S = 1 Heisenberg antiferro-magnet on a spatially anisotropic triangular lattice, J. Phys.Soc. Jpn. 82, 043715 (2013).

41 As we are working on finite size clusters, the local mag-netizations are somewhat higher than the expected onesin the thermodynamic limit: m ' 0.20 for the triangularlattice31,32, and m ' 0.26 for the honeycomb case33.

42 L. O. Manuel and H. A. Ceccatto, Magnetic and quantumdisordered phases in triangular-lattice Heisenberg antiferro-magnets, Phys. Rev. B 60, 9489 (1999).

43 We compute this correlation averaging over all the firstneighbors of the cluster, except those closest to the openedges in order to avoid border effects.

44 R. Shindou and T. Momoi, SU(2) slave-boson formulationof spin nematic states in S= 1/2 frustrated ferromagnets,Phys. Rev. B 80, 064410 (2009).

45 Another evidence of the absence of long-range ferromag-netic order in the central spin subsystem, below J ′c, is givenby its zero magnetization even under the application of asmall uniform magnetic field h to this subsystem. See34.

Page 7: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

1

Supplemental material for:Correlated partial disorder in a weakly frustrated quantum antiferromagnet

M. G. Gonzalez, F. T. Lisandrini, G. J. Blesio, A. E. Trumper, C. J. Gazza, and L. O. Manuel

Instituto de Fısica Rosario (CONICET) and Universidad Nacional de Rosario, Boulevard 27 de Febrero 210 bis,(2000) Rosario, Argentina

We sketch the linear spin wave calculation for the Heisenberg model on lattices with complex unit cells, as it corre-sponds to the

√3×√

3-distorted triangular lattice. Also, we present the Weiss mean field-like treatment of a 4-spincluster model in order to understand the role of quantum fluctuations of the honeycomb Neel order in the establish-ment of ferromagnetic correlations between central spins in the partially disordered phase. Finally, we present someDMRG results related to the convergence with the number of states kept, finite size effects, the determination ofSmaxz , and the spin-spin correlation between honeycomb and C spins.

I. LINEAR SPIN WAVE CALCULATION

In this section, we present the linear spin wave theory used to solve semiclassically the Heisenberg model. TheHamiltonian (1) can be rewritten in a generic way as

H =1

2

∑ijαβ

Jαβ(Ri −Rj)Siα · Sjβ , (S1)

where i, j denote unit cells with vector positions Ri, Rj , respectively, while α, β represent the spins inside the unitcell. Jαβ(Ri −Rj) is the exchange interaction between the spin α in the unit cell Ri and the spin β in the unit cell

Rj . In our case, the√

3×√

3-distorted triangular lattice is a non-Bravais lattice that can be described as a triangular

lattice with basis vectors a1 =(

3a2 ,√3a2

)and a2 = (0,

√3a) (a is the nearest neighbor distance), and a unit cell with

the three sites A, B, and C (see Fig. 1). Jαβ ’s only take non-zero values for nearest neighbors:

JAB(0) = JAB(a1) = JAB(a1 − a2) = J,

JAC(0) = JBC(0) = JAC(a2) = JBC(a2) = JAC(a1) = JBC(a2 − a1) = J ′,

and the remaining ones can be obtained through the relation Jαβ(Ri −Rj) = Jβα(Rj −Ri).In order to perform the spin wave analysis, we first need the classical ground state of the Heisenberg model. For

this purpose, we consider classical spin vectors in (S1) and we minimize the corresponding energyS1

Eclas =∑αβq

Jαβ(q)Sqα · S−qβ , (S2)

where Jαβ(q) =∑

R Jαβ(R)eiq·R is the Fourier transform of the exchange interaction matrix J, and Sqα =1√N

∑i Siαe

iq·Ri are classical vectors subject to the normalization condition S2iα = 1. The sum over R runs over all

the N unit cells in the cluster. For the√

3×√

3-distorted triangular lattice, the minimization of the classical energy(S2) gives the magnetic wave vector Q = 0, implying that the same absolute magnetic pattern is repeated in eachunit cell. If φα is the angle between the spins in the α sublattice and the z direction, after a little algebra we findφA = −φB = − arccos(−J ′/2J), and φC = 0. This three sublattice (coplanar) structure is shown in Fig. 1.

Once we know the classical ground state magnetic structure, we make a rotation to local axes in (S1). That is, ateach site we define a local frame in such a way that the classical order points along the local z axis. If, for simplicity,we consider a planar classical magnetic structure (like in our case), lying in the xz plane, then the spin operators inthe local axes are

Siα = Ry(Θiα)Siα,

where Θiα = Q ·Ri + φα is the angle between the classical Siα spin and the global z direction, and Ry(Θiα) is the

matrix rotation of angle Θiα around the y direction. In terms of the local spin operators S, the Hamiltonian (S1)turns into

H =1

2

∑ijαβ

[cos(Θiα−Θjβ)

(SxiαS

xjβ + SziαS

zjβ

)+ sin(Θiα−Θjβ)

(SziαS

xjβ − SxiαSzjβ

)+ SyiαS

yjβ

].

Page 8: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

2

In the local axis frame, the classical magnetic structure is a ferromagnetic one along the z direction. Now, we can

represent the local spins by bosonic operators aiα, a†iα, using the typical Holstein-Primakoff transformationS2:

Sziα = S − a†iαaiα,

S+iα =

√2S − a†iαaiα aiα, (S3)

S−iα = a†iα

√2S − a†iαaiα.

In the semiclassical approximation, valid for S → ∞, the (operator) square roots that appear in the Holstein-

Primakoff representation (S3) are replaced by the scalar√

2S, yielding the linear spin wave approximation. Afterperforming this approximation and a Fourier transformation to momentum space, the Hamiltonian (S1) is written as

HLSW =

(1 +

1

S

)Eclas +

S

2

∑αβq

a†−qα · Dαβ(q) · aqβ , (S4)

where the spinor aqβ = (aqβ , a†−qβ)T , the bosonic dynamical matrix is

Dq =

(Aαβ(q) −Bαβ(q)−Bαβ(q) Aαβ(q)

),

and

Aαβ(q) =∑R

Jαβ(R) cos(q ·R) cos2(Q ·R + φα − φβ

2

)− δαβ

∑γR

Jαγ(R) cos (Q ·R + φα − φγ) ,

Bαβ(q) =∑R

Jαβ(R) cos(q ·R) sin2

(Q ·R + φα − φβ

2

).

The quadratic Hamiltonian (S4) can be diagonalized by means of a paraunitary Bogoliubov transformationS3. After

this procedure, the order parameter at each site can be evaluated using miα = | < Sziα > |.

II. WEISS MEAN FIELD-LIKE APPROXIMATION FOR A 4-SPIN CLUSTER

In order to understand the origin of the correlated character of the partial disordered phase below J ′c ' 0.18, weresort to a molecular Weiss-like mean field approach for a simple toy model. We consider a 4-spin cluster as it isshown in Fig. S1, composed of two central spins C (3 and 4) and two honeycomb spins, one A (1) and another B (2).Its Hamiltonian is

H = JS1 · S2 + J ′ (S1 + S2) · (S3 + S4) , (S5)

and we are interested in the case J ′ � J .

1

J'J

2

3

4

FIG. S1: 4-spin cluster model. The solid line corresponds to an interaction J , while the dashed ones correspond to J ′.

The exact ground state of Hamiltonian (S5) can be easily found, and it consists, for J ′/J ≤ 1/2,S4 in a singlet statebetween the 1 and 2 spins, while the remaining spins 3 and 4 can have any spin projection σ, σ′:

|GS >=

(|1 ↑ 2 ↓> −|1 ↓ 2 ↑>√

2

)⊗ |3σ4σ′ > .

Page 9: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

3

Naturally, the average spin correlation between 3 and 4 is zero, and this model corresponds to another example, likethe ones presented in Fig. 2, where the less connected spins are completely decoupled. However, we are not interestedin the exact solution of (S5) –with its singlet state between spins 1 and 2–, but in an approximated treatment that

mimics what happens in the partially disordered phase of the√

3×√

3-distorted triangular Heisenberg model.As DMRG numerical results indicate that the honeycomb Neel order remains almost unaltered for 0 ≤ J ′ . 0.18,

we propose to fix the spins 1 and 2 in the normalized state

|ψ(1, 2) >=1√

1 + 2r2[|1 ↑ 2 ↓> +r (|1 ↑ 2 ↑> +1 ↓ 2 ↓>)] , (S6)

that represent a “classical Neel order” (spin 1 ↑ and spin 2 ↓) plus quantum fluctuations quantified by the realparameter r. r can be related to the local magnetization of the “honeycomb spins” as

m0 =< ψ(1, 2)|Sz1 |ψ(1, 2) >=1

2(1 + 2r2),

so we get the relation r '√

12 −m0 � 1 for m0 close to its classical value.

In the spirit of the Weiss molecular field theory, we approximate (S5) by an effective Hamiltonian for spins 3 and 4,

H34 = J < ψ(1, 2)|S1 · S2|ψ(1, 2) > +J ′ < ψ(1, 2)| (S1 + S2) |ψ(1, 2) > · (S3 + S4) , (S7)

which results in the expression

H34 = 4m0rJ′ (Sx3 + Sx4 ) + const. (S8)

That is, the quantum fluctuations of the classical Neel order of the spins 1 and 2 act as an effective uniform magneticfield in the x direction for the central spins 3 and 4. For any r > 0 the ground state of (S8) is |3 ↓ 4 ↓>x, where xrefers to the quantization axis.

Of course, the x and y directions are equivalent, the same as the spin projection. To show this, we consider a moregeneral “Neel state” than (S6), giving the possibility that the zero-point quantum fluctuation states |1 ↑ 2 ↑> and|1 ↓ 2 ↓> are out of phase, that is

|ψ(1, 2) >=1√

1 + 2r2

[|1 ↑ 2 ↓> +r

(eiθ↑ |1 ↑ 2 ↑> +eiθ↓ |1 ↓ 2 ↓>

)]. (S9)

Inserting this state in (S7), we get the effective Hamiltonian

H34 = 2m0rJ′ [(cos θ↑ + cos θ↓) (Sx3 + Sx4 ) + (sin θ↓ − sin θ↑) (Sy3 + Sy4 )] ,

that is, a Zeeman term for spins 3 and 4 with an effective magnetic field ~B = 2m0rJ′ (cos θ↑ + cos θ↓, sin θ↓ − sin θ↑, 0) ,

perpendicular to the classical Neel order in the z direction. Again, in the ground state |ψ(3, 4) > of this Hamiltonian

the spins 3 and 4 are aligned along the ~B direction and, straightforwardly, it results in the classical ferromagneticcorrelation

< ψ(3, 4)|S3 · S4|ψ(3, 4) >=1

4

between them, independently of the value of J ′. It is worth to notice that in the two extreme cases, when the stateof spins 1 and 2 is taken as a classical Neel one or as a singlet, there is no correlation between spins 3 and 4. So, aquantum Neel order is necessary for the spins 3 and 4 to be correlated.

III. DMRG RESULTS

A. Ground state energy as a function of Sz: determination of Smaxz

The left and right panels of Fig. S2 show the dependence of the ground state energy with the quantum number Szfor the partially disordered phase (J ′ = 0.12 < J ′c) and for the ferrimagnetic phase (J ′ = 0.25 > J ′c), respectively. Inorder to illustrate a technical point, in Fig. S2, we display the ground state energy for different m number of stateskept in the DMRG computations.

Page 10: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

4

-25.91

-25.90

-25.89

0 1 2 3 4 5 6 7 8 9 10

E

Sz

m=800m=1500m=3000

-25.9

-25.8

0 2 4 6 8 10 12

E

Sz

-26.39

-26.38

-26.37

-26.36

-26.35

0 1 2 3 4 5 6 7 8 9 10

E

Sz

m=800m=1500m=3000

-26.4

-26.2

-26.0

0 2 4 6 8 10 12

E

Sz

(a) (b)

FIG. S2: Energy ground state as a function of Sz for different number of states m kept in DMRG: (a) for J ′ = 0.12 < J ′c inthe PD phase, (b) for J ′ = 0.25 > J ′c in the ferrimagnetic phase. The cluster size is Ns = 12× 6.

For J ′ < J ′c, in the PD phase, the ground state has Sz = 0, corresponding to a singlet state. There are nearlyquasi-degenerate states, corresponding to small Sz values; however, there is an appreciable energy difference (of theorder of J ′) between the ground state and the fully saturated ferromagnetic C subsystem, as it is illustrated in theinset of Fig. S2 (a).

Within the ferrimagnetic phase J ′ > J ′c (right panel), there is a degeneracy among states with different Sz, up toa maximum Smax

z = 8 . For smaller m, it seems that the ground state has a Sz quantum number different from zero,however, due to the SU(2) symmetry, this cannot be the case. This non-monotonic behavior of the ground stateenergy as a function of Sz is related to the fact that for larger Sz the Hilbert space is smaller, and DMRG performsbetter, yielding a lower ground state energy. Note that, as we increase the number of states m, the ground stateenergy clearly shows a degeneracy for Sz between 0 and Smax

z , signaling the presence of a ferrimagnetic ground state.In this way, we determine the value of Smax

z for different J ′.

At J ′ = J ′c, Smaxz jumps from zero to a finite value, lower than that corresponding to a fully polarized C subsystem,

± 12Ns/3. This can be explained considering that, although the local magnetization at the center of the hexagons is

very large (see Fig. 3 of the main text), however, it never reaches the classical value 1/2 (mC < 0.45). So, even atthe onset of the magnetic ordering of the C spins, quantum fluctuations prohibit that Smax

z has its maximum possiblevalue. From Eq. (3) of the main text, we can see that close to J ′c, where θ ' π, Smax

z has a value of the order ofmCNs/3.

B. Stability of the partial disordered phase under a magnetic field

In the above subsection, we have found that the ground state of the partially disordered phase has Sz = 0, withother Sz states close in energy. In order to check the robustness of the vanishing of the mC local magnetization inthe PD phase, we have applied a uniform small magnetic field h = J ′/10 only to the C central spins. In this way, wewould favor any tendency of the central spins to align ferromagnetically, while the honeycomb sublattice would remainunperturbated. However, as it is shown in the left panel of Fig. S3, mC = 0 even in the presence of the magnetic fieldh, departing its behavior from the zero-field case only for J ′ very close to J ′c. On the other hand, mA(= mB) (rightpanel) does not change at all in the PD phase. Note that in the previous known partially disordered phases (presentin certain Ising and XY-like models), the orphan spins are uncorrelated and, as a consequence, the application of asmall magnetic field will be enough to order them ferromagnetically. In our model, the correlated singlet nature ofthe PD phase makes the system robust against the magnetic field perturbation.

Close to the isotropic point (J ′ = J), the effect of the magnetic field is appreciable because h becomes an im-portant fraction of J, reducing quantum fluctuations in both magnetic sublattices, and, as a consequence, the localmagnetizations mA and mC increase with h.

Page 11: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

5

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.2 0.4 0.6 0.8 1.0

mC

J'

h=0.0 J'h=0.1 J'

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.2 0.4 0.6 0.8 1.0

mA

J'

h=0.0 J'h=0.1 J'

FIG. S3: Local magnetization of the central spin sublattice mC (left panel) and the honeycomb sublattice mA(= mB) (rightpanel) for the 12 × 6 triangular cluster, with and without the application of a uniform magnetic field h = J ′/10 in the Csublattice only.

C. Convergence with the number of states kept and finite size effects

To analyze the convergence of the DMRG predictions with the number m of states kept, we show in Fig. S4 thebehavior of the local magnetizations for different m, using the cluster with Lx = 12 × Ly = 6 sites. It can be seenan excellent quantitative agreement for the three values of m presented, for almost all J ′. Only for m = 1000, mC

has some noisy behavior close to J ′c, that dissappears for larger m. It is worth to emphasize that the numericalresults in our model converge for a relatively small value of m in comparison to what happens in highly frustratedantiferromagnets, like the kagome Heisenberg model. This fact would signal that in the PD phase for J ′ 6= 0 there isa low effective degeneracy, albeit the system has an extensive degeneracy for J ′ = 0.

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.2 0.4 0.6 0.8 1.0

mC

J'

m=1000m=1500m=3000

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.2 0.4 0.6 0.8 1.0

mA

J'

m=1000m=1500m=3000

FIG. S4: Local magnetization mC (left) and mA(= mB) (right) as a function of J ′ for the triangular cluster 12×6, for differentnumber m of kept states.

Once we are confident that for m & 1500 the DMRG predictions are well converged, we focus on the finite sizeeffects. We present in Fig. S5 the local magnetizations mA and mC as a function of J ′ for clusters with Ly = 6 legsand different Lx, using m = 1500 states. Clearly, the finite size effects are, in general, not negligible, although theoverall behavior of both magnetizations does not change with Lx. Most relevant for our findings is the fact that, for0 ≤ J ′ . 0.5, mC exhibits rather small finite size effects, resulting, as a consequence, that the partially disorderedphase survives in the thermodynamic limit, with a reliable critical value J ′c ' 0.18. On the other hand, the honeycombsublattice has appreciable finite size effects for all J ′, the same happens for the central spins close to the isotropic

Page 12: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

6

point J ′ = J . In these cases, an appropriate thermodynamic limit extrapolationS5 would yield reduced, but finite,local magnetizations (mA = mC ' 0.205 for the isotropic pointS5, mA ' 0.268 for the honeycomb lattice).

With respect to the stability of our results for clusters with different Ly, in Fig. S6 we present the local mag-netizations for the Lx = 12, Ly = 4 triangular cluster. It can be seen that for this smaller cluster, there is also atransition to a partially disordered phase below J ′c ' 0.18 and there are similar behaviors of the local magnetizationof both subsystems. We have checked that the C spins are also ferromagnetically correlated (with an almost constantcorrelation below J ′c) for this cluster. Besides, we have found that close to the isotropic triangular point (J ′ ' 0.7),the magnetization curve for the C subsystem has a shallow dip, probably due to the quasi-one-dimensional characterof this cluster (as it was discussed for the isotropic triangular Heisenberg modelS5).

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.2 0.4 0.6 0.8 1.0

mC

J'

9x612x615x6

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.2 0.4 0.6 0.8 1.0

mA

J'

9x612x615x6

FIG. S5: Local magnetization mC (left) and mA(= mB) (right) as a function of J ′ for triangular clusters with Ly = 6 anddifferent Lx.

0.0

0.1

0.2

0.3

0.4

0.5

0.0 0.2 0.4 0.6 0.8 1.0

m

J'

mAmC

FIG. S6: Local magnetization of the honeycomb sublattice mA(= mB) and of the C sublattice mC for the 12 × 4 triangularcluster.

Regarding the finite size effects in the ferrimagnetic state, Figs. S7 display the behavior of the ground state energyfor J ′ = 0.25 as a function of Sz for cluster sizes corresponding to Ly = 6 and different Lx’s. If we normalize Sz withthe number of C sites, NC (right panel of Fig. S7), the value of the normalized Smax

z is almost independent of Lx.This gives us confidence about the relevance of our results in the thermodynamic limit.

Page 13: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

7

0.000

0.004

0.008

0.012

0.016

0 2 4 6 8 10 12 14

(E-E

0)/|

E 0|

Sz

Lx=9Lx=12Lx=15

0.004

0.008

0.012

0.016

0 0.2 0.4 0.6 0.8 1

(E-E

0)/|

E 0|

2Sz/NC

Lx=9Lx=12Lx=15

(a) (b)

0.000

FIG. S7: DMRG ground state energy as function of (a) the quantum number Sz and (b) the quantum number Sz normalizedby the number NC of C spins, for different cluster sizes with Ly = 6.

D. Spin-spin correlations between the magnetic subsystems in the partially disordered phase

Fig. S8 shows the (averaged) nearest neighbor spin-spin correlation between the C spins and the A (or B) honeycombspins as a function of J ′. It can be seen that, except for J ′ = 0, there exists a finite antiferromagnetic correlationbetween the two subsystems, even in the partially disordered phase, 0 ≤ J ′ ≤ 0.18. This result reinforces thecorrelated character of the partial disorder phase, that is, the orphan spins do not behave as a perfect paramagnetlike in previous known PD phases. Furthermore, the finite correlation between subsystems supports our argumentsabout the Casimir-like effect that C spins are ferromagnetically aligned at short distance via the zero-point quantumfluctuations of the magnetically ordered honeycomb subsystem.

-0.20

-0.15

-0.10

-0.05

0.00

0.0 0.2 0.4 0.6 0.8 1.0

<S i

Sj>

C,A

B

J'

FIG. S8: Nearest neighbor spin-spin correlation between the honeycomb subsystem A(B) spins and the center of hexagon Cspins.

[S1] T. Nagamiya, in Solid State Physics Vol. 20, edited by F. Seitz and D. Turnbull (Academic Press, New York, 1967).[S2] T. Holstein and H. Primakoff, Field Dependence of the Intrinsic Domain Magnetization of a Ferromagnet, Phys. Rev. 58,

1098 (1940).[S3] J. H. P. Colpa, Diagonalization of the quadratic boson Hamiltonian, Physica A 93, 327 (1978).

Page 14: N eel ordered honeycomb subsystem, coexisting with ...It is in the quantum magnetism arena, precisely, where a plethora of control factors are available for tuning the amount of quantum

8

[S4] At J ′/J = 1/2, there is transition to a plaquette-singlet state, where < S3 · S4 >= 1/4.[S5] S. R. White and A. L. Chernyshev, Neel order in square and triangular lattice Heisenberg models, Phys. Rev. Lett. 99,

127004 (2007).