You are on page 1of 11

PHYSICAL REVIEW B 102, 094101 (2020)

Worm-algorithm-type simulation of the quantum transverse-field Ising model

Chun-Jiong Huang ,1,2,3 Longxiang Liu,1,2,3 Yi Jiang,4,* and Youjin Deng 1,2,3,†
1
Shanghai Branch, National Laboratory for Physical Sciences at Microscale and Department of Modern Physics,
University of Science and Technology of China, Shanghai, 201315, China
2
CAS Center for Excellence and Synergetic Innovation Center in Quantum Information and Quantum Physics,
University of Science and Technology of China, Hefei, Anhui 230026, China
3
CAS-Alibaba Quantum Computing Laboratory, Shanghai, 201315, China
4
Department of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China

(Received 24 June 2020; accepted 18 August 2020; published 2 September 2020)

We apply a worm algorithm to simulate the quantum transverse-field Ising model in a path-integral
representation of which the expansion basis is taken as the spin component along the external-field direction. In
such a representation, a configuration can be regarded as a set of nonintersecting loops constructed by “kinks” for
pairwise interactions and spin-down (or -up) imaginary-time segments. The wrapping probability for spin-down
loops, a dimensionless quantity characterizing the loop topology on a torus, is observed to exhibit small
finite-size corrections and yields a high-precision critical point in two dimensions (2D) as hc = 3.044 330(6),
significantly improving over the existing results and nearly excluding the central value of the previous result
hc = 3.044 38(2). At criticality, the fractal dimensions of the loops are estimated as d↓ (1D) = 1.37(1) ≈ 118
and d↓ (2D) = 1.75(3), consistent with those for the classical 2D and 3D O(1) loop model, respectively. An
interesting feature is that in one dimension (1D), both the spin-down and -up loops display the critical behavior
in the whole disordered phase (0  h < hc ), having a fractal dimension d = 1.750(7) that is consistent with the
hull dimension dH = 74 for critical 2D percolation clusters. The current worm algorithm can be applied to simulate
other quantum systems like hard-core boson models with pairing interactions.

DOI: 10.1103/PhysRevB.102.094101

I. INTRODUCTION the worm algorithm exhibits efficiency comparable to cluster


schemes. A high-precision estimate of the square-lattice criti-
The quantum transverse-field Ising model (QTFI) is a
cal point is obtained as 3.044 330(6), significantly improving
textbook model in quantum many-body physics and plays
the existing results and nearly excluding the central value of
an important role in quantum phase transition [1] and quan-
3.044 38(2) [5]. In the path-integral representation for the cur-
tum information science [2]. The one-dimensional QTFI can
rent worm algorithm, a configuration can be regarded as a set
be solved exactly [3], and it has been widely used to test
of nonintersecting loops constructed by “kinks” for pairwise
theoretical or numerical methods [4,5] and to study novel
interactions and spin-down (or -up) imaginary-time segments.
quantities like entanglement entropy [6,7] and quantum fi-
Rich geometric properties are observed for these loops. In par-
delity susceptibility [8]. In higher dimensions, analytical re-
ticular, it is found that in 1D, the loops over a wide parameter
sults are scarce, and one has to rely on numerical or ap-
range exhibit scaling laws that are in the universality class of
proximate methods. Many methods have been developed,
the classical 2D percolation. Deep theoretical understanding
including transfer-matrix method [9], series expansion [8,10],
is desired. Further, a variety of physical quantities, including
continuous-time Monte Carlo approach [5,11–15], tensor
the magnetic and the fidelity susceptibilities, are examined.
renormalization group method [16], density matrix renormal-
The rest of the paper is organized as following. Section II
ization group [17], projected entangled-pair states [18], and
explains the current path-integral representation for the QTFI
machine learning method [19,20], etc. Nevertheless, to obtain
and the formulation of the worm algorithm. The numerical
a high-precision critical point still remains to be a challenging
results are presented in Sec. III. A brief summary is given in
task. To our knowledge, the best estimates of the critical point
Sec. IV.
for the 2D QTFI are 3.044 2(4) in Ref. [8] and 3.044 38(2) in
Ref. [5], achieved by stochastic series expansion (SSE) and
continuous-time Wolff cluster methods, respectively.
In this work, we apply a worm algorithm to simulate QTFI II. WORM ALGORITHM
in 1D and 2D. It is shown that as pointed out in Refs. [21,22], The Hamiltonian of the QTFI on the d-dimensional cubic
lattice is
 
*
jiangyi@ustc.edu.cn H = −t σiz σ jz − h σix , (1)

yjdeng@ustc.edu.cn i j i

2469-9950/2020/102(9)/094101(11) 094101-1 ©2020 American Physical Society


HUANG, LIU, JIANG, AND DENG PHYSICAL REVIEW B 102, 094101 (2020)

(a) (b) (c) shall refer to K1 and K2 as the hopping and the pairing term,
I
M
respectively.
With Eq. (3), the partition function of Hamiltonian (2) can
τ be formulated in the Feynman’s path-integral representation
τ τ x
(also called the world-line representation) as

x x Z = Tr[e−βH ] = α0 |e−βH |α0 
α0
FIG. 1. Illustration of path-integral configurations. (a) For 
Eq. (1). Red (gray) segments represent the spin-up (-down) state. = lim α0 |e−Hdτ |αn−1  . . . α1 |e−Hdτ |α0 
β
dτ = n
(b) A G configuration for Eq. (8) in the rotated basis, having an open n→∞ {αi }
path with the two ends marked as I and M. Blue (green) lines are for ∞  β
 β  β N

pairing (hopping) kink. (c) A sketch of Z configuration on the 1D = ... dτk F (t, t  , h) (5)
torus, with a nonlocal loop of winding numbers (Wτ = 4, Wx = α0 N=0 0 τ1 τN−1 k=1
1). For simplicity, the lattice structure is not shown.
with the integrand function
  β 
where σiα (α = x,z) are Pauli matrices, i j represents nearest- F (t, t  , h) = t Nh t 
Np
exp − U (τ )dτ , (6)
neighboring sites, t > 0 is the ferromagnetic interaction 0
strength, and h is the transverse field. Taking the σ z spin where Nh and Np are, respectively, the number of hopping
component as the expansion basis for the path-integral repre- and pairing kinks (N = Nh + Np ), |αi  = |σ1z , σ2z , . . . , σNz 
sentation, one can map a d-dimensional QTFI onto a (d +1)- is an eigenstate in the σ z basis (N is the total number of
dimensional classical system, for which each lattice site has a lattice sites). Moreover, Eq. (5) can be graphically viewed
continuous line of spin segments; see Fig. 1(a) for an example. as the summation/integration over configurations in the (d +
This continuous dimension is called the imaginary-time (τ ) 1)-dimensional space-time {i, τ }, of which the statistical
direction, along which the spin state σiz can be flipped by the weight is
σix operator but must satisfy the periodic condition. The length
of the τ dimension is the inverse temperature β = 1/kB T (the N

Boltzmann constant is set as kB = 1 from now). WZ (t, t  , h) = dτk F (t, t  , h). (7)
To formulate a worm algorithm that is effective for config- k=1
urations of closed loops, we choose the external-field direction In such a representation, each lattice site has a line of spin
as the expansion basis and rewrite Hamiltonian (1) as segments, and at imaginary time τk (k = 1, 2, . . . , N), a pair
  of neighboring spins is simultaneously flipped either by a
H ≡ K + U = −t σix σ jx − h σiz . (2) hopping term K1 or by a pairing term K2 . We shall call them
i j i the hopping or the pairing kink, respectively. Starting from an
As a result, U and K are, respectively, the diagonal arbitrary space-time point (i, τ ), one would construct a closed
andthe nondiagonal terms. The pairwise interactions K = loop by following spin-up (-down) segments and kinks. Thus,
−t i j σix σ jx can be further expressed in terms of the raising a configuration effectively consists of closed loops.
An important ingredient of the worm algorithm is then to
and lowering spin operators σ ± = (σ x ± iσ y )/2, as
extend the configuration space Z for Eq. (5) by including two
K ≡ K1 + K2 defects. For the QTFI, the extended configuration space G is
  for the spin-spin correlation function of the Pauli matrix σ x :
= −t (σi+ σ j− + H.c.) − t (σi+ σ j+ + H.c.). (3) 

i j i j
G(xI , xM , τI , τM ) = Tr Tτ σIx (τI )σMx (τM )e−βH , (8)
where Tτ is the τ -ordering operator. In addition to closed
The term K1 flips a pair of opposite spins and thus the total
loops, a path-integral configuration in the G space contains an
magnetization is conserved along the τ direction, while K2
open path with two ending points; see an example in Fig. 1(b).
flips a pair of spins of the same sign. We note that with
We shall refer to the ending points as “Ira” (I) and “Masha”
the Holstein-Primakoff transformation, bi (b†i ) = σi− (σi+ ) and
(M), and denote their coordinates in the space-time as (xI , τI )
thus ni ≡ b†i bi = (σiz + 1)/2, the QTFI can be mapped onto a and (xM , τM ). The statistical weight of the G configuration can
hard-core Bose-Hubbard (BH) model with Hamiltonian be written as
  
H = −t (b†i b j + H.c.) − t  (b†i b†j + H.c.) − μ ni , N
dτI dτM 
i j i j WG = dτk F (t, t  , h), (9)
i
ωG k=1
(4)
where F is given by Eq. (6) and ωG is an arbitrary positive
where t  = t, the particle number ni = 0, 1, and the chemical constant. When I coincides with M, the open path forms a
potential μ = 2h. In the language of the hard-core BH model, closed loop, and the G space is reduced to the Z space.
K1 accounts for the hopping of a particle, and K2 , which The full configuration space for the worm-type simulation
simultaneously creates/deletes a pair of particles, represents corresponds to the combination of the G and the Z spaces.
the pairing of two neighboring bosons. For convenience, we For ergodicity, the simulation must be able to change the

094101-2
WORM-ALGORITHM-TYPE SIMULATION OF THE QUANTUM … PHYSICAL REVIEW B 102, 094101 (2020)

Making use of Eqs. (7) and (9), the acceptance probabilities


I M for the Metropolis filter can be calculated as

βN Fnew
(a) Create/Annihilate defects (I, M) Pcrea = min 1, Aa τa ,
ωG Fold
(11)
1 1 ωG Fnew
I M I M Panni = min 1, ,
Aa τa βN Fold
(b) Move imaginary time of defect M where Fnew and Fold , given by Eq. (6), is respectively for the
M configuration after and before the corresponding operation.
Note that the statistical-weight change Fnew /Fold is mainly de-
I M I termined by the random displacement |δ|  τa /2. As a result,
the acceptance probabilities in Eq. (11) can be optimized by
(c) Insert/Delete a kink tuning τa .
A natural choice for the relative weight is ωG = βN since
FIG. 2. The three updates. the acceptance probabilities in Eq. (11) then hardly depend
on L and β. Physically, this is because the spin-spin corre-
lation function G(xI , xM , τI , τM ) has the space-time translation
kink number, the space-time location of any kink as well invariance so that the statistical weight of a G configuration
as of defects (I, M), and to switch configurations back and should be normalized by the factor 1/ωG = 1/(βN ). Further,
forward between the Z and G spaces. We adopt the following with this choice, the number of Monte Carlo steps between
three updates: (a) create/annihilate defects (I, M), (b) move two adjacent creations of (I, M), called the worm-return time,
imaginary time of defect M, and (c) insert/delete a kink. The measures the ratio of the G space over the Z space, and
first operation switches configurations between the Z and exactly gives the dynamic magnetic susceptibility of the QTFI
the G spaces by creating or annihilating a pair of defects which is stated explicitly in Sec. III C.
(I, M). The second updates the τM value, and the third changes (b) Move imaginary time of defect M. The update, reverse
the xM value by inserting or deleting a kink. Except “create to itself, is chosen with a probability Ab in the G space.
defects (I, M),” all updates only apply in the G space, and each One randomly selects a time displacement δ ∈ [−τb /2, τb /2)
of them is chosen with an a priori probability given before and δ = 0 assigns τM = mod(τM + δ, β ) for the new temporal
simulation. location of defect M, and flips the spin states in-between; see
(a) Create/annihilate defects (I, M). If the current configu- Fig. 2(b). The types of in-between kinks are also interchanged.
ration is in the Z space, “create defects (I, M)” is the only pos- The acceptance probability is
sible update. One randomly picks up a point (xI , τI ) from the Pmove = min {1, Fnew /Fold }. (12)
whole space-time volume βN = βL d , draws a uniformly dis-
tributed imaginary-time displacement δ ∈ [−τa /2, τa /2) and (c) Insert/delete a kink. Each operation is chosen with a
δ = 0 with the range τa ∼ O(1/h) [23], assigns xM = xI and probability Ac in the G space. In “insert a kink,” one randomly
τM = mod(τI + δ, β ), and flips the spin state between defects chooses one of the zd = 2d neighboring world lines of xM , say
I and M. The β periodicity is taken into account by the modular xM , and updates the spatial location of M as (xM , τM ) → (xM , τM ).
function. As illustrated in Fig. 2(a), the types (hopping or pair- Meanwhile, one inserts a kink k between world lines xM and
ing) of kinks between defects I and M, if any, are interchanged xM at imaginary time τk = mod(τM + δ, β ), with a random
during this operation. displacement δ ∈ [−τc /2, τc /2) and δ = 0. Further, the spin
The update “annihilate defects (I, M),” the reverse oper- states between τM and τk , on both xM and xM , are flipped which
ation of “create defects (I, M),” is chosen with an a priori causes the types of in-between kinks, linking xM and xM , to
probability Aa in the G space. It changes a G configuration stay the same. However, the types of in-between kinks are
into a Z one by annihilating defects (I, M) and flipping the interchanged if they link some other world lines to xM or xM .
spin state in-between. This is possible only if I and M are An example is illustrated in Fig. 2(c).
on the same world line xI = xM and their imaginary-time In the reverse operation, “delete a kink,” one also picks
displacement min {|τI − τM |, β − |τI − τM |}  τa /2. up a random neighboring world line xM of xM and moves M
Accordingly, the detailed-balance condition reads as as (xM , τM ) → (xM , τM ). Further, one counts the number nk of
kinks that connect world lines xM and xM in the imaginary-time
domain [τM − τc /2, τM + τc /2). If no kink exists nk = 0, the
dτI dτM operation is rejected. Otherwise, one randomly picks up one
WZ Pcrea = AaWG Panni , (10)
βN τa of the nk kinks and deletes it, and meanwhile flips the spin
states on both world lines between τM and the imaginary time
where Pcrea (Panni ) is the acceptance ratio for the “create of the deleted kink. Besides types of kinks linking xM or xM to
defects” (“annihilate defects”) operation, WZ (WG ) is the other world lines are interchanged as well.
statistical weight for the configuration before (after) the cre- The detailed balance condition of this pair of operations
ation of defects, and dτI /(βN ) and dτM /τa account for the reads as
probability of choosing the space-time location for I and M, 1 dτk 1 1
respectively. Ac W Pinse = Ac W+ Pdele , (13)
zd τc zd nk

094101-3
HUANG, LIU, JIANG, AND DENG PHYSICAL REVIEW B 102, 094101 (2020)

where Pinse (Pdele ) is for the acceptance ratio for “insert a locate the phase transition hc , we make use of the topological
kink” (“delete a kink”). The statistical weights W and W+ , properties of the nonintersecting loops on the torus, instead of
given by Eq. (9), are respectively for the configuration before the scaling behaviors of physical quantities like the magnetic
and after inserting a kink. The infinitesimal dτk on the left- susceptibility.
hand side is canceled by W+ , which has one more kink. The
acceptance probabilities are then A. Critical point

τc Fnew Given a Z configuration, we record how many times
Pinse = min 1, ,
nk +1 Fold Wi  0 each loop  winds along the ith direction (i =

nk Fnew
(14) 2, . . . , d ), and calculate the total winding number Wi =
1,
Pdele = min 1, ,  Wi from all the loops; see Fig. 1(c) for an illustration.
τc Fold A path-integral configuration is said to wrap along the ith
direction as long as Wi > 0. This is indicated as Ri = 1;
where nk denotes the number of in-between kinks for the
current configuration. The denominator nk + 1 in Pinse reflects  Ri = 0. The average wrapping probability R =
otherwise,
(1/d ) i Ri  is then calculated, with · representing the
an extra kink in the updated configuration.
ensemble average. In the sub-percolating phase, the loops
The worm algorithm is then formulated as in Algorithm 1,
are too small to percolate, and the R value quickly drops to
in which a priori probabilities satisfy Aa + Ab + 2Ac = 1.
0 as L becomes larger. In the superpercolating phase, there
It is mentioned again that the acceptance probabilities in the
is at least one giant loop with large Wi , and the R value
updates can be optimized by tuning the ranges of random τ
rapidly converges to 1. At the percolation threshold, the R
displacement, τa , τb , and τc . As its analog for the classical
values for different system sizes L have an asymptotically
Ising model which carries out a weighted random walk over
common intersection with a nontrivial value between 0 and
the lattice, the defect M in this quantum Monte Carlo method
1. In short, the wrapping probability R is a dimensionless
effectively performs a random walk in the space-time and
quantity characterizing the topological feature of loops on
simultaneously updates the spin states it passes by.
torus. In many cases, such wrapping probabilities are found
For the conventional BH model which does not have the
to exhibit small finite-size corrections, and have been widely
pairing term, the interchange between the hopping and pairing
used for locating critical points [24–28].
kink cannot be allowed. For “create/annihilate defects” and
“move imaginary time of M,” the above illegal updates can
be avoided when performing these operations only within a Algorithm 1 Worm algorithm
larger spin segment. In “insert/delete a kink,” the simplest
remedy is that the proposed update is rejected as long as BEGIN: Given a Z configuration.
it leads to an illegal configuration, giving a price that the loop
if it is a Z configuration
acceptance probabilities are decreased by a factor of O(1/h).
choose the “create defects (I, M)” operation
As a more sophisticated remedy, one can reformulate the
else
operation in a way such that no illegal configuration would be
choose an operation with its a priori probability except
introduced. “create defects (I, M)”
Finally, for the computational efficiency, it is important to end if
implement hash tables such that each operation is done within calculate the acceptance probability P and carry out the
O(1) CPU time. operation with the probability P
end loop
III. NUMERICAL RESULTS
In the absence of the external field (h = 0), the spin-up and For the QTFI, two types of nonintersecting loops, spin-up
-down states are fully balanced in the QTFI (2). As h turns on, (↑) or -down (↓) loops, can be constructed. The Monte Carlo
the system evolves into a disordered phase with the spin-down (MC) results in Fig. 3 show that irrespective of the spatial
state being suppressed and the spin-up order still not formed dimension (1D or 2D), both the spin-up and -down loops
(h < hc ). It enters into the spin-up ordered phase (h > hc ) display critical behaviors near the quantum critical point hc .
through a second-order quantum phase transition. The critical In 1D, particularly rich behaviors are observed. In the whole
point is exactly known as hc /t = 1 in 1D and numerically disordered phase (0  h < hc ), both R↑ and R↓ have nontriv-
determined as hc /t ≈ 3.044 in 2D (square lattice). Without ial values, 0 < R↑ (R↓ ) < 1, indicating the fractal structures
loss of generality, the pairwise interaction is set as t = 1 from of these loops. At the transition point h = hc , the R↑ and
now unless stated explicitly. R↓ values have a sharp drop, which becomes infinitely sharp
Using the worm algorithm, we simulate the 1D and 2D for L → ∞. For h > hc , the R↓ value drops to 0, meaning
QTFIs with linear lattice size L and inverse temperature β = that the spin-down loops are too small to percolate, while
L; the choice of β = L is due to the dynamic critical exponent the R↑ value converges to nontrivial value if h is not too
z = 1 for the QTFI. Periodic boundary conditions are applied large, suggesting that the spin-up loops still exhibit fractal
in each spatial direction, so that the lattice is essentially a properties. Nevertheless, as h is further increased, the R↑ value
torus. The linear size is taken up to L = 512 in 1D and L = also gradually approaches to 0. This is understandable because
128 in 2D, and no severe critical slowing down is observed. A the number of kinks decreases when h increases, and thus
variety of geometric and physical quantities are sampled. To the spin loops are less likely to percolate. In 2D, the R↑ and

094101-4
WORM-ALGORITHM-TYPE SIMULATION OF THE QUANTUM … PHYSICAL REVIEW B 102, 094101 (2020)

0.8 0.52 0.02


0.6 0.7 fit
0.6 R↑ 0.5 R↓ 0.5
0.00 R̃↓
0.4 0.12 0.4
0.99 1.00 1.01 64 0.3
0.99 1.00 1.01 0.51 vs.
0.09
128
-0.02
(h − hc )Lyt
0.2 256 1D
0.06 -0.10 0.00 0.10
1.00 1.10 1.20
0.0
(a) (b) R↓ 0.50
0.0 0.4 0.8 1.2 1.6 0.0 0.4 0.8 1.2 1.6 2.0 0.4995
h h
1.0 0.49
16 1D
0.8 32
64 64
0.6 256
R↑ R↓ 2D 0.48 512
0.4
0.9998 0.9999 1.0000 1.0001 1.0002
0.2
(c) (d) h
0.0
0.0 2.0 4.0 0.0 2.0 4.0 6.0 FIG. 4. Wrapping probability R↓ in 1D. The gray band indicates
h h
an interval of 2σ above and below the estimate hc = 1.000 001(5).
FIG. 3. Wrapping probabilities R↑ and R↓ versus the transverse The yellow band indicates an interval of 2σ above and below the
field h. The error bars are much smaller than the size of points. The estimate Rc = 0.4995(3). The inset shows R̃↓ = R↓ − Rc − bi L yi
vertical black lines indicate the critical point. (a), (b) are for 1D, and versus (h − hc )L yt , with a1 = −0.1625(9), a2 = −0.34(13), bi =
(c), (d) are for 2D. The inset plots of (a) and (b) show the curve near 4.3(1.5), yi = −2, and yt = 1.
hc = 1.
In 2D, an eye-view fitting of the R↓ data in Fig. 5 already
gives the critical point approximately as hc ≈ 3.044 33, with
R↓ values converge to 1 in the disordered phase 0  h < hc , uncertainty at the fifth decimal place. We find that it is
suggesting a superpercolating phase both for the spin-up and sufficient to describe these data by Eq. (15) with a2 = 0 which
-down loops. For h > hc , the R↓ value quickly reaches 0, but means the fit is linear and for Lmin = 16, b2 can also be set to
R↑ seems to converge to 1 for L → ∞. At h = hc , the R↓ value zero. The fit gives R↓,c = 0.528 1(14) and hc = 3.044 330(6).
has a sharp drop, and the derivative of R↑ with respect to h To test the reliability of the value and the error bar of hc ,
probably also develops a singularity as L increases. we plot in Fig. 6 the R↓ data against L at some fixed h near
Extensive simulations are then carried out at h = hc = 1 in hc . It can be seen that at h = hc = 3.044 330, the wrapping
1D and h = 3.04435 in 2D, and the data nearby are obtained probability R↓ converges to a constant value within the 2σ
by the standard reweighting technique [29]. The system size shadow area in Fig. 6. In contrast, as L increases, the R↓ data
is taken as L = 16, 64, 256, 512 in 1D, with at least 4 × 108
samples for each L, and L = 8, 16, 32, 64, 128 in 2D, with at
least 1 × 108 samples for each L.
According to the least-squares criterion, the R↓ data, partly 0.01 fit
shown in Figs. 4 and 5, are fitted by 0.54 0.00 R̃↓
vs.
-0.01

2 (h − hc )Lyt
R↓ = R↓,c + ak (h − hc )k L kyt + bi L yi + b2 L y2 . (15)
-0.1 0.0 0.1
k=1 0.53
R↓
0.5281
The thermal renormalization exponents are fixed at the known
values in the classical (d + 1) Ising universality, i.e., yt (1D) = 2D
1 and yt (2D) = 1.5868 [30]. The term with bi comes from 0.52
the leading irrelevant thermal field, which has the exponent 16
32
yi (1D) = −2 and yi (2D) = −0.821 [28,30]. The subleading 64
correction exponents are set as y2 (1D) = −3 and y2 (2D) = 128
−2. As a precaution, we gradually increase Lmin and exclude 0.51
3.04415 3.04425 3.04435 3.04445 3.04455
the L < Lmin data from the fit to see how the ratio of the
h
residual χ 2 to the degree of freedom changes with Lmin .
In 1D, it is found that the MC data for Lmin = 64 can be FIG. 5. Wrapping probability R↓ in 2D. The gray band indicates
well described by Eq. (15) without the correction-to-scaling an interval of 2σ above and below the estimate hc = 3.044 330(6).
term (b2 = 0). The fit yields hc = 1.000 001(5), in excellent The yellow band indicates an interval of 2σ above and below the es-
agreement with the exact quantum critical point hc = 1. Also, timation Rc = 0.528 1(14). The inset displays R̃↓ = R↓ − Rc − bi L yi
we have R↓,c = 0.4995(3), which suggests that it might ex- versus (h − hc )L yt , with a1 = −0.0855(8), bi = −0.031(8), yi =
actly be 21 ; see the inset of Fig. 3(b). −0.821, and yt = 1.568.

094101-5
HUANG, LIU, JIANG, AND DENG PHYSICAL REVIEW B 102, 094101 (2020)

0.535 104 0.4 0.4 104


2D 0.6 0.6
0.8 0.8
103
1.0 1.0
S1↑ 103 S1↓
1.2
102
0.530
2 1D 1D
0.5281 10
101
16 32 64 128 256 16 32 64 128 256
R↓ L L

0.525 FIG. 7. The largest-loop size S1↑ and S1↓ versus L at h =


0.4, 0.6, 0.8, 1.0, 1.2 for 1D. For S1↑ , the straight lines have a slope
7
3.044300 4
, irrespective of the h value. For S1↓ , the lines have slope 47 for
3.044330 h < hc and 118 for h = hc .
0.520 3.044360
3.044380
20 40 60 80 100 120 zNk = 0.30(3) for 2D. The efficiency of the worm algorithm is
comparable to that of the Wolff-type cluster method [5]. Note
L
that as the spatial dimension increases, all the values of zO de-
FIG. 6. Plot of R↓ versus L for various values of h for the 2D crease. From the numerical results of the worm algorithm for
QTFI. The yellow strip indicates an interval of 2σ above and below the classical Ising model [21,34], we expect zO = 0 (without
the estimate R↓,c = 0.528 1(14). The solid lines are plotted according critical slowing down) for d  dc , where dc = 3 is the upper
to the fitting result. critical dimensionality for the QTFI.

for h = 3.044 300 and 3.044 360 bend upward and downward, B. Geometric properties of loops
respectively, suggesting that they are clearly away from the We have determined the quantum critical point hc with a
thermodynamic critical point. For h = 3.044 38, which is high precision by locating the percolation threshold of the
the estimated central value of the critical point in Ref. [5], the loop configurations. Hereby, we shall further explore other
downward bending is stronger, meaning that it cannot be the geometric properties of the spin-up and -down loops at and
critical point. Table I gives a (incomplete) list of the existing away from hc . In the Z space, we measure the average
results for hc in 2D. It is clear that our estimate has the highest length S1 of the largest loop and the probability distribution
precision. that a randomly chosen loop is of size s, i.e., P(s, L) ≡
We now briefly discuss the efficiency of the current worm (1/N (L))∂N (s, L)/∂s, where N (L) ∼ βL d is the total num-
algorithm, which is already reflected by the precision of the ber of loops and N (s, L) is the number of loops of size in
estimated critical point hc . For a quantitative evaluation, we range (s, s + ds).
calculate at criticality the integrated autocorrelation times τint In 1D, one knows from Fig. 3 that for the whole region
for the energy E, magnetization M, and kink number Nk , 0  h  hc , both the spin-up and -down loops exhibit critical
in the unit of MC sweeps. A MC sweep is defined such scaling behaviors. For the spin-up loops, such fractal prop-
that on average, each imaginary-time spin line is updated by erties further survive in the ordered phase h > hc . In these
βt times. From the least-squares fitting τ ∝ L zO , we obtain cases, we expect that the largest-loop size scales as S1 ∝ L d ,
the dynamical exponent as zE = 0.38(3), zM = 0.35(3), and where d < (1 + 1) is the loop fractal dimension, and that the
zNk = 0.41(3) for 1D, and zE = 0.28(3), zM = 0.23(4), and loop-size distribution behaves as [35,36]
P(s, L) ∼ s−τ f (s/L d ), (16)
TABLE I. Estimated critical point hc on the square lattice. CMC: where τ is called the Fisher exponent. Moreover, the expo-
cluster Monte Carlo method; SSE: stochastic series expansion; S-W:
nents τ and d are related by the hyperscaling relation τ =
Swendsen-Wang in continuous time; HOSVD: tensor renormaliza-
1 + (d + 1)/d . The function f (x ≡ s/L d ) is universal and
tion group method based on the higher-order singular value decom-
describes the finite-size cutoff of s near S1 ∼ L d .
position; iPEPS: infinity projected entangled-pair state; MERA: mul-
tiscale entanglement renormalization ansatz; CTM: corner transfer
We simulate at h = 0.4, 0.6, 0.8, 1.0, 1.2 and the results
matrix. are shown in Fig. 7. The straight lines in the log-log plot
suggest that indeed, the largest loop has a fractal structure.
Method hc For the spin-up loops, irrespective of the h value, the straight
lines have the same slope approximately as 47 . Further, the
This work 3.044 330(6) amplitude of the power law S1↑ ∼ L 7/4 increases as a function
CMC [5] 3.044 38(2) of h, at least in the range of 0.4  h  1.2. In contrast, as h
SSE [8] 3.044 2(4)
increases, the largest-loop size S1↓ decreases and then drops
S-W [11] 3.044(1)
to a significantly smaller value at h = hc . Further, while the
HOSVD(D=14) [16] 3.043 9
lines for h < hc still have a slope near 47 , the line for h = hc
iPEPS [31] 3.04
MERA [32] 3.075 has a smaller slope which is about 11 8
. This suggests that
CTM [33] 3.14 the spin-down loops start with a dense and critical phase for
h < hc , experience a critical state at h = hc , and then enter

094101-6
WORM-ALGORITHM-TYPE SIMULATION OF THE QUANTUM … PHYSICAL REVIEW B 102, 094101 (2020)

TABLE II. Estimates of d↑ and d↓ at different h in 1D. spin-up loops spin-down loops
−1
h 0.4 0.6 0.8 1.0(hc ) 1.2 10 64 64
128 128
−3
10 256 256
d↑ 1.754(6) 1.750(5) 1.747(5) 1.750(6) 1.751(3)
d↓ 1.751(5) 1.749(7) 1.750(7) 1.37(1) 10−5

10−7

10−9 (a) (a’)


into a sparse phase containing enormous small loops. The S1
−1
data for both the spin-up and -down loops are fitted by 10

10−3
S1 = L d (a0 + b1 L y1 ), (17)

P (s, L)
10−5
with different choices of the correction exponent y1 = 10−7
−0.5, −1.0, or −1.5. We find that the fits are rather stable,
and the results are shown in Table II. 10−9 (b) (b’)
We notice that the configuration of the classical O(n) 10 −1

loop model on the honeycomb lattice also consists of non-


10−3
intersecting loops [37–39]. Moreover, the O(n) loop model
with n = 1 corresponds to the 2D Ising model, and has a 10−5
hull/loop dimension as dhull = 11 at the critical point xc = 10−7
 √ √ 8
1/ 2 + 2 − n = 1/ 3 and dhull = 74 in the dense phase 10−9 (c) (c’)
x > xc , where x is the statistical weight for each loop unit
100 101 102 103 104 100 101 102 103 104
[37–39]. These behaviors are very similar to those of the s s
spin-down loops for the 1D QTFI. Accordingly, we conjecture
that in 1D, the fractal dimensions d↓ (h = hc ) = 1.37(1) and FIG. 8. Loop-size distribution P(s, L) for different h values in
d↓ (h < hc ) = 1.750(6) are exactly identical to 11 8
and 47 , 1D. The figures in the left (right) panel are for the spin-up (-
respectively. We also conjecture that the fractal dimension down) loops, and the first, second, and third rows correspond to
d↑ = 1.747(5), which is independent of the h value, is also h = 0.8, 1.0, 1.2, respectively. The straight lines, with slope − 157 or
− 27 , are a guide for the eye.
exactly equivalent to 47 . Further, it is noted that by the duality 11

relation, the loops on the honeycomb lattice can be mapped


onto the boundaries of the spin domains for the Ising model on spin-up loops spin-down loops
the triangular lattice. In the dense phase x > xc , these domains 16
(a) 64 (a’) 64
are simply critical site-percolation clusters. Therefore, we ex- 128 128
12 2.0
pect that the domains, enclosed by the spin-up or -down loops, 256 256
are also fractal and have a fractal dimension corresponding 8
to that for critical Ising spin domains or percolated clusters 1.0
in 2D. 4
To further demonstrate the fractal structure of the spin-up
0 0.0
and -down loops in 1D, we display in Fig. 8 the MC data 16 (b) (b’) 1.2
for the loop-size distribution P(s, L). Indeed, one observes
algebraically decaying behaviors s−τ for the spin-up loops 12
P (s, L)sτ

0.8
with h = 0.8, 1.0, 1.2 and for the spin-down loops with h = 8
0.8, 1.0. The cutoff size of s for the power-law scaling, due
0.4
to finite-size effects, increases as the system size L. The 4
hyperscaling relation τ = 1 + (d + 1)/d is well confirmed 0 0.0
by the fact that the data for different L collapse onto the (c) (c’)
straight lines with slope − 15 7
or − 27
11
. For the spin-down loops 12
0.6
in the ordered phase h = 1.2, the P(s, L) data for different L
drop quickly, illustrating that the loop sizes are finite even in 8

the thermodynamic limit L → ∞. We further plot sτ P(s, L) 0.3


4
versus s/L d in Fig. 9. With the values of (d , τ ) as ( 47 , 157
) or
( 11 ,
8 11
27
), the data for different L collapse well onto a single 0 0.0
0 0.4 0.8 1.2 1.6 0 0.4 0.8 1.2
curve, illustrating the universal feature of the cutoff function s/Ld ↑ s/Ld ↓
f (x). It is interesting to see that for the spin-down loops at h =
hc , function f (x) displays a two-peak structure [Fig. 9(b )]. FIG. 9. P(s, L)sτ versus s/L d in 1D. The plots in the left (right)
We regard that the first peak at the smaller value of x reflects panel are for the spin-up (-down) loops, and the first, second, and
the residual effect of the spin-down loops in the disordered third rows correspond to h = 0.8, 1.0, 1.2, respectively. The values
phase h < hc . Meanwhile, it is observed that for the spin-up of d↑ and d↓ are listed in Table II, and the τ value is calculated
loops with h = 1.2, function f (x) exhibits a shoulder feature from the hyperscaling relation τ = 1 + (d + 1)/d .

094101-7
HUANG, LIU, JIANG, AND DENG PHYSICAL REVIEW B 102, 094101 (2020)

104 1D 2D
40
104 1D 2D 3

8 30
16
2 24 12

χ /L
3 16
Tω 10 32
103 48 24 20
64
1
10

102 102
24 25 26 27 28 24 25 26 0 0
0.90 0.95 1.00 1.05 1.10 0.31 0.32 0.33 0.34 0.35
L L
t t

FIG. 10. Worm-return time at criticality hc . The straight lines,


FIG. 11. Fidelity susceptibility χF (t ) versus t, with (a) for 1D
with slope 2yh − (d + 1), are a guide for the eye.
and (b) for 2D. The black solid lines indicate the critical points
tc (1D) = 1 and tc (2D) = 1/3.044 330. The insets show χF /L 2yt ver-
on the smaller-x side. We expect that as h increases, such sus (t − tc )L yt which indicate the universality of the function fχF (x).
a shoulder feature would become more pronounced and its
location would move toward the value x = 0. This is because D. Fidelity susceptibility
that as h is enhanced, the number of kinks will be gradually
suppressed and the sizes of the spin-up loops will eventually It is well known that many systems can undergo quantum
start to decrease. In the limiting case h → ∞, all the spin- phase transitions without spontaneous symmetry breaking
up loops will become individual imaginary-time lines with and thus without a good definition of local order parame-
length β. ter. These phase transitions are beyond the Ginzburg-Landau
In 2D, the spin-down loops are fractal only at h = hc , and paradigm, and are difficult to be detected by conventional
the spin-up loops are always in a superpercolating phase. thermodynamic observables. Fidelity susceptibility, a quantity
The fit of the S1↓ data by Eq. (17) gives d↓ = 1.75(3). proposed in the quantum information science [41], has been
Again, this is in excellent agreement with the loop dimension shown to be useful for such a purpose [8,42–45]. Consider
dhull = 1.734(4) for the classical O(n = 1) loop model on the a quantum phase transition driven by some given parameter
3D hydrogen-peroxide lattice [40], on which the loops are λ and let |φ(λ) represent the corresponding wave function,
also nonintersecting. As expected, for the spin-down loops at the fidelity F (λ, ) of the system is defined as the overlap
h = hc = 3.044 330, the loop-size distribution P(s, L) follows between the wave functions with different values of λ, i.e.,
Eq. (16). F (λ, ) = |φ(λ)|φ(λ + )|. Accordingly, the fidelity sus-
ceptibility χF (λ) is calculated as

C. Worm-return time ∂ 2 ln F(λ, ) 
χF (λ) = −  . (20)
The worm-return time Tw is the average update steps ∂ 2 =0
between two adjacent Z configurations in the markov chain For the QTFI, we hereby choose the driving parameter λ to be
MC simulation. Mathematically, it can be expressed as the the pairwise interaction t, which is conjugate to the number of
integral of spin-spin correlation function (8) over the lattice kinks Nk . Given a Z configuration, let Nk,1 and Nk,2 denote
and the imaginary time as the total number of kinks in the first-half imaginary-time
⎡ ⎤ domain 0  τ < β/2 and the second-half one β/2  τ < β,
 β β 
1 respectively. It can be shown by following Ref. [43] that the
Tw = Tr⎣Tτ dτI dτM σIx (τI )σMx (τM )e−βH ⎦ fidelity susceptibility χF (t ) is proportional to the covariance
ωG Z 0 0 xI ,xM of Nk,1 and Nk,2 , and can be written as
⎡ 2 ⎤
 β  Nk,1 Nk,2  − Nk,1 Nk,2 
1 χF (t ) = ,
= Tr⎣ dτ σix (τ ) e−βH ⎦ 2t 2
(21)
ωG Z 0 i
where the external field h is now set to be 1. The MC data
 2 
1 β of χF (t ) for the 1D and 2D QTFIs are shown in Fig. 11.
−βH
= Tr x
dτ M (τ ) e . (18) As expected, the χF (t ) data for each L display a peak near
ωG Z 0 the critical point tc . As system size L increases, the peak
With the choice of ωG = βN, the worm-return time Tw is location tL , called the pseudocritical point, moves toward the
precisely equal to the dynamic magnetic susceptibility χ xx = thermodynamic critical point tc , and the peak itself becomes
β sharper with a smaller width. Following the standard finite-
[ 0 dτ M x (τ )]2 /βN. Figure 10 shows the Tw data at h = hc size scaling analysis, we expect that near the critical point tc ,
for both 1D and 2D, which are fitted by the fidelity susceptibility χF scales as
Tw = L 2yh −(d+1) (a0 + bi L yi ). (19) χF (t, L) = L 2yt fχF (L yt (t − tc )). (22)
In 1D, the fit with yi = −2 gives yh = 1.876(2), in excellent Indeed, making use of yt (1D) = 1 and yt (2D) = 1.5868, we
agreement with the exact value 15
8
. In 2D, we set yi = −0.821 obtain a good collapse when plotting χF /L 2yt versus (t −
[30] and obtain yh = 2.484(4), which is again well consistent tc )L yt , as shown in the insets of Fig. 11.
with the result yh = 2.4816(1) for the classical 3D Ising For the fidelity F (λ, ), we can also choose the driving
model [30]. parameter to be the external field h for the QTFI, which is

094101-8
WORM-ALGORITHM-TYPE SIMULATION OF THE QUANTUM … PHYSICAL REVIEW B 102, 094101 (2020)

conjugate to the σ z -component magnetization M. In this case, ∞  β


 β  β N

we should consider the magnetization M1 for 0  τ < β/2 = C ... dτi
and M2 for β/2  τ < β, and the fidelity susceptibility χF (h) {α0 } N=0 0 τ1 τN−1 k=1

would be proportional to the covariance of M1 and M2 . At the Np


critical point, we expect χF ∝ L 2yh . ×nN↓ t Nh t  (e−2h )S↓ , (23)
While Fig. 11 illustrates the applicability of the fidelity where C = e , N↓ specifies the number of spin-down
hβN
susceptibility χF as a tool for studying the quantum phase loops, and S↓ is the total length of spin-down loops. We
transition, it is worth mentioning that by calculating the expect that for t = t  , the phase transition of such a “quantum
covariance of two quantities of the same kind but in sepa- O(n) loop” model in d dimensions will belong to the same
rated spatial/imaginary-time domains, χF normally has large universality class as that for the classical O(n) loop model in
fluctuations. Thus, to achieve a good statistics for χF would (d + 1) dimensions. In 1D, we further expect that the exact
require extensive simulations. value of the quantum critical point hc (n) can be obtained for
the “quantum O(n) loop” model, and that the spin-down loops
would exhibit rich geometric properties both at criticality
hc (n) and in the disordered phase h < hc . In particular, for
(d = 1, n = 2), the phase transition would be of the cele-
IV. DISCUSSION
brated Berezinskii-Kosterlitz-Thouless topological transition.
We formulate a worm-type algorithm and study the QTFI All these expectations can be explored by the current worm-
in a path-integral representation in which configurations are type algorithm, and remain to be a future work.
sets of nonintersecting loops. By locating the percolation The efficiency of the current worm algorithm implies its
threshold of loop configurations via the so-called wrapping broad applications in a variety of spin and hard-core systems.
probability, we obtain a high-precision quantum critical point A straightforward application is to simulate the QTFI on other
hc = 3.044 330(6) for the QTFI on the square lattice. These lattices regardless of dimensionality. For the high dimension
nonintersecting loops are further observed to exhibit rich ge- d  3, one expects very minor or absent critical slowing
ometric properties, particularly in 1D, where both the spin-up down, and thus interesting logarithmic corrections can be
and -down loops have fractal structures over a wide parameter examined. It can be also of significant relevance in solid-state
range. By examining the similarity of the scaling behaviors experiments since the pairing terms σi+ σ j+ and σi− σ j− are
for the d-dimensional QTFI and for the (d + 1)-dimensional found to occur in frustrated quantum materials due to the
classical O(n = 1) model, we conjecture that in 1D the two dipolar-octupolar doublets [46–60]. Further, in addition to the
fractal dimensions are d↓ (hc ) = 11 and d↓ (h < hc ) = 74 , and external field h, one can introduce pairing interaction σiz σ jz
8
that in 2D, d↓ (hc ) = 1.75(3) is equal to the hull dimension along the σ z direction, which can be either ferromagnetic
dhull = 1.734(4) for the classical 3D loop model. The finite- or antiferromagnetic. This allows the worm-type study of
size scalings of magnetic and fidelity susceptibilities are also quantum spin systems with geometric frustration with respect
examined. It is confirmed that the fidelity susceptibility can be to the σ z component. In combination with the so-called clock
used to probe quantum phase transitions. Monte Carlo method [61], one can even study spin systems
Motivated by the fact that the classical O(1) loop model with long-range σiz σ jz interaction without heavy computa-
is a specific case of the O(n) loop model with n = 1, we can tional overhead. Finally, we mention that a similar worm
generalize the loop path-integral representation of the QTFI algorithm has recently been used in the SSE representation of
by giving each spin-down loop a statistical weight n. As a the hard-core bosonic Hubbard model with pairing terms [62].
consequence, the partition function (5) is generalized to be
ACKNOWLEDGMENTS
∞  β
 β  β N

Z(t, t  , h, n) = ... dτi We acknowledge N. Prokofiev for initializing the project
{α0 } N=0 0 τ1 τN−1 k=1 and X.-P. Yao for technical help in preparing Fig. 1(c). This
β work is supported by the National Natural Science Foundation
Np
×nN↓ t Nh t  e− 0 U (τ )dτ
of China under Grant No. 11625522.

[1] Sachdev Subir, Quantum Phase Transitions (Cambridge Uni- [4] G. Vidal, Classical Simulation of Infinite-Size Quantum Lattice
versity Press, Cambridge, 2011). Systems in One Spatial Dimension, Phys. Rev. Lett. 98, 070201
[2] A. Dutta, G. Aeppli, B. K. Chakrabarti, U. Divakaran, T. F. (2007).
Rosenbaum, and D. Sen, Quantum Phase Transitions in Trans- [5] H. W. J. Blöte and Y. Deng, Cluster monte carlo simula-
verse Field Spin Models: From Statistical Physics to Quantum tion of the transverse ising model, Phys. Rev. E 66, 066110
Information (Cambridge University Press, 2015). (2002).
[3] P. Pfeuty, One-dimensional ising model with a transverse [6] A. Kitaev and J. Preskill, Topological Entanglement Entropy,
field,Ann. Phys. (NY) 57, 79 (1970). Phys. Rev. Lett. 96, 110404 (2006).

094101-9
HUANG, LIU, JIANG, AND DENG PHYSICAL REVIEW B 102, 094101 (2020)

[7] M. Levin and X.-G. Wen, Detecting Topological Order in A [28] P. Hou, S. Fang, J. Wang, H. Hu, and Y. Deng, Geometric
Ground State Wave Function, Phys. Rev. Lett. 96, 110405 properties of the fortuin-kasteleyn representation of the ising
(2006). model, Phys. Rev. E 99, 042150 (2019).
[8] A. F. Albuquerque, F. Alet, C. Sire, and S. Capponi, Quan- [29] D. P. Landau and K. Binder, A Guide to Monte Carlo Simu-
tum critical scaling of fidelity susceptibility, Phys. Rev. B 81, lations in Statistical Physics, 4th. ed. (Cambridge University
064418 (2010). Press, Cambridge, 2014), Chap. 7.
[9] T. Kaya, One-dimensional quantum transverse-field ising [30] Y. Deng and H. W. J. Blöte, Simultaneous analysis of several
model: A semiclassical transfer matrix approach, Int. J. Mod. models in the three-dimensional ising universality class, Phys.
Phys. B 24, 5457 (2010). Rev. E 68, 036125 (2003).
[10] O. F. Syljuåsen and A. W. Sandvik, Quantum monte carlo with [31] R. Orus and G. Vidal, Simulation of two-dimensional quantum
directed loops, Phys. Rev. E 66, 046701 (2002). systems on an infinite lattice revisited: Corner transfer matrix
[11] H. Rieger and N. Kawashima, Application of a continuous time for tensor contraction, Phys. Rev. B 80, 094403 (2009).
cluster algorithm to the two-dimensional random quantum ising [32] G. Evenbly and G. Vidal, Entanglement Renormalization in
ferromagnet, Eur. Phys. J. B 9, 233 (1999). Two Spatial Dimensions, Phys. Rev. Lett. 102, 180406 (2009).
[12] N. V. Prokof’ev, B. V. Svistunov, and I. S. Tupitsyn, “Worm” [33] R. Orus, Exploring corner transfer matrices and corner tensors
algorithm in quantum Monte Carlo simulations, Phys. Lett. A for the classical simulation of quantum lattice systems, Phys.
238, 253 (1998). Rev. B 85, 205117 (2012).
[13] T. Ikegami, S. Miyashita, and H. Rieger, Griffiths-mccoy sin- [34] E. M. Elçi, J. Grimm, L. Ding, A. Nasrawi, T. M. Garoni, and
gularities in the transverse field ising model on the randomly Y. Deng, Lifted worm algorithm for the ising model, Phys. Rev.
diluted square lattice, J. Phys. Soc. Jpn. 67, 2671 (1998). E 97, 042126 (2018).
[14] S. V. Isakov and R. Moessner, Interplay of quantum and thermal [35] K. Binder, “Clusters” in the ising model, metastable states and
fluctuations in a frustrated magnet, Phys. Rev. B 68, 104409 essential singularity, Ann. Phys. (NY) 98, 390 (1976).
(2003). [36] M. E. Fisher, The theory of condensation and the critical point,
[15] V. I. Iglovikov, R. T. Scalettar, R. R. P. Singh, and J. Oitmaa, Phys. Phys. Fiz. 3, 255 (1967).
Disorder line and incommensurate floating phases in the quan- [37] Q. Liu, Y. Deng, and T. M. Garoni, Worm monte carlo study
tum ising model on an anisotropic triangular lattice, Phys. Rev. of the honeycomb-lattice loop model, Nucl. Phys. B 846, 283
B 87, 214415 (2013). (2011).
[16] Z. Y. Xie, J. Chen, M. P. Qin, J. W. Zhu, L. P. Yang, and T. [38] H. Saleur and B. Duplantier, Exact Determination of the Perco-
Xiang, Coarse-graining renormalization by higher-order singu- lation Hull Exponent in Two Dimensions, Phys. Rev. Lett. 58,
lar value decomposition, Phys. Rev. B 86, 045139 (2012). 2325 (1987).
[17] S. R. White, Density Matrix Formulation for Quantum Renor- [39] A. Coniglio, Fractal Structure of Ising and Potts Clusters: Exact
malization Groups, Phys. Rev. Lett. 69, 2863 (1992); Density- Results, Phys. Rev. Lett. 62, 3054 (1989).
Matrix Algorithms for Quantum Renormalization Groups, [40] Q. Liu, Y. Deng, T. M. Garoni, and H. W.J. Blöte, The o(n) loop
Phys. Rev. B 48, 10345 (1993). model on a three-dimensional lattice, Nucl. Phys. B 859, 107
[18] F. Verstraete and J. I. Cirac, Renormalization algorithms for (2012).
quantum-many body systems in two and higher dimensions, [41] M. Nielsen and I. Chuang, Quantum Computation and Quantum
arXiv:cond-mat/0407066. Information (Cambridge University Press, Cambridge, 2010).
[19] G. Carleo and M. Troyer, Solving the quantum many-body [42] P. Zanardi and N. Paunković, Ground state overlap and quantum
problem with artificial neural networks, Science 355, 602 phase transitions, Phys. Rev. E 74, 031123 (2006).
(2017). [43] L. Wang, Y.-H. Liu, J. Imriška, P. N. Ma, and M. Troyer, Fi-
[20] J. Carrasquilla and R. G. Melko, Machine learning phases of delity Susceptibility Made Simple: A Unified Quantum Monte
matter, Nat. Phys. 13, 431 (2017). Carlo Approach, Phys. Rev. X 5, 031007 (2015).
[21] N. Prokof’ev and B. Svistunov, Worm Algorithms for [44] S.-J. Gu and H.-Q. Lin, Scaling dimension of fidelity suscepti-
Classical Statistical Models, Phys. Rev. Lett. 87, 160601 bility in quantum phase transitions, Europhys. Lett. 87, 10003
(2001). (2009).
[22] Y. Deng, T. M. Garoni, and A. D. Sokal, Dynamic Critical [45] L. Campos Venuti and P. Zanardi, Quantum Critical Scal-
Behavior of the Worm Algorithm for the Ising Model, Phys. ing of the Geometric Tensors, Phys. Rev. Lett. 99, 095701
Rev. Lett. 99, 110601 (2007). (2007).
[23] Under this condition, τa is almost constant for different βL d . [46] Y.-P. Huang, G. Chen, and M. Hermele, Quantum Spin Ices
Besides, the acceptance ratio will not be too small or too large. and Topological Phases from Dipolar-Octupolar Doublets on
[24] M. E. J. Newman and R. M. Ziff, Efficient Monte Carlo Al- the Pyrochlore Lattice, Phys. Rev. Lett. 112, 167203 (2014).
gorithm and High-Precision Results for Percolation, Phys. Rev. [47] M. C. Hatnean, M. R. Lees, O. A. Petrenko, D. S. Keeble,
Lett. 85, 4104 (2000). G. Balakrishnan, M. J. Gutmann, V. V. Klekovkina, and
[25] P. H. L. Martins and J. A. Plascak, Percolation on two- and B. Z. Malkin, Structural and magnetic investigations of single-
three-dimensional lattices, Phys. Rev. E 67, 046119 (2003). crystalline neodymium zirconate pyrochlore Nd2 Zr 2 O7 , Phys.
[26] X. Feng, Y. Deng, and H. W. J. Blöte, Percolation transitions in Rev. B 91, 174416 (2015).
two dimensions, Phys. Rev. E 78, 031136 (2008). [48] A. Bertin, P. Dalmas de Réotier, B. Fåk, C. Marin, A. Yaouanc,
[27] J. Wang, Z. Zhou, W. Zhang, T. M. Garoni, and Y. Deng, A. Forget, D. Sheptyakov, B. Frick, C. Ritter, A. Amato,
Bond and site percolation in three dimensions, Phys. Rev. E 87, C. Baines, and P. J. C. King, Nd2 Sn2 O7 : An all-in–all-out py-
052107 (2013). rochlore magnet with no divergence-free field and anomalously

094101-10
WORM-ALGORITHM-TYPE SIMULATION OF THE QUANTUM … PHYSICAL REVIEW B 102, 094101 (2020)

slow paramagnetic spin dynamics, Phys. Rev. B 92, 144423 [56] R. Sibille, E. Lhotel, V. Pomjakushin, C. Baines, T. Fennell, and
(2015). M. Kenzelmann, Candidate Quantum Spin Liquid in the Ce3+
[49] E. Lhotel, S. Petit, S. Guitteny, O. Florea, M. Ciomaga Hatnean, Pyrochlore Stannate Ce2 Sn2 O7 , Phys. Rev. Lett. 115, 097202
C. Colin, E. Ressouche, M. R. Lees, and G. Balakrishnan, (2015).
Fluctuations and All-in–All-Out Ordering in Dipole-Octupole [57] Y.-D. Li and G. Chen, Symmetry enriched U (1) topological
Nd2 Zr 2 O7 , Phys. Rev. Lett. 115, 197202 (2015). orders for dipole-octupole doublets on a pyrochlore lattice,
[50] J. Xu, V. K. Anand, A. K. Bera, M. Frontzek, D. L. Abernathy, Phys. Rev. B 95, 041106(R) (2017).
N. Casati, K. Siemensmeyer, and B. Lake, Magnetic struc- [58] J. Gaudet, E. M. Smith, J. Dudemaine, J. Beare, C. R. C.
ture and crystal-field states of the pyrochlore antiferromagnet Buhariwalla, N. P. Butch, M. B. Stone, A. I. Kolesnikov, G. Xu,
Nd2 Zr 2 O7 , Phys. Rev. B 92, 224430 (2015). D. R. Yahne, K. A. Ross, C. A. Marjerrison, J. D. Garrett, G. M.
[51] V. K. Anand, A. K. Bera, J. Xu, T. Herrmannsdörfer, C. Ritter, Luke, A. D. Bianchi, and B. D. Gaulin, Quantum Spin Ice Dy-
and B. Lake, Observation of long-range magnetic ordering in namics in the Dipole-Octupole Pyrochlore Magnet Ce2 Zr 2 O7 ,
pyrohafnate Nd2 Hf 2 O7 : A neutron diffraction study, Phys. Rev. Phys. Rev. Lett. 122, 187201 (2019).
B 92, 184418 (2015). [59] B. Gao, T. Chen, D. W. Tam, C.-L. Huang, K. Sasmal, D. T.
[52] O. Benton, Quantum origins of moment fragmentation in Adroja, F. Ye, H. Cao, G. Sala, M. B. Stone, C. Baines, J. A. T.
Nd2 Zr 2 O7 , Phys. Rev. B 94, 104430 (2016). Verezhak, H. Hu, J.-H. Chung, X. Xu, S.-W. Cheong, M.
[53] J. Xu, C. Balz, C. Baines, H. Luetkens, and B. Lake, Spin Nallaiyan, S. Spagna, M. Brian Maple, A. H. Nevidomskyy
dynamics of the ordered dipolar-octupolar pseudospin- 21 py- et al., Experimental signatures of a three-dimensional quantum
rochlore Nd2 Zr 2 O7 probed by muon spin relaxation, Phys. Rev. spin liquid in effective spin-1/2 Ce2 Zr 2 O7 pyrochlore, Nat.
B 94, 064425 (2016). Phys. 15, 1052 (2019).
[54] V. K. Anand, D. L. Abernathy, D. T. Adroja, A. D. Hillier, [60] Y. D. Li and G. Chen, Non-spin-ice pyrochlore U (1) quantum
P. K. Biswas, and B. Lake, Muon spin relaxation and inelastic spin liquid: Manifesting mixed symmetry enrichments, Phys.
neutron scattering investigations of the all-in/all-out antiferro- Rev. Research 2, 013334 (2020).
magnet Nd2 Hf 2 O7 , Phys. Rev. B 95, 224420 (2017). [61] M. Michel, X. Tan, and Y. Deng, Clock Monte Carlo methods,
[55] P. Dalmas de Réotier, A. Yaouanc, A. Maisuradze, A. Bertin, Phys. Rev. E 99, 010105(R) (2019).
P. J. Baker, A. D. Hillier, and A. Forget, Slow spin tunneling in [62] A. J. R. Heng, W. Guo, A. W. Sandvik, and P. Sengupta, Pair
the paramagnetic phase of the pyrochlore Nd2 Zr 2 O7 , Phys. Rev. hopping in systems of strongly interacting hard-core bosons,
B 95, 134420 (2017). Phys. Rev. B 100, 104433 (2019).

094101-11

You might also like