You are on page 1of 9

PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

Peratic phase transition by bulk-to-surface response

Xingze Qiu ,1,2 Hai Wang ,1,2 Wei Xia,1 and Xiaopeng Li1,2,3,*
1
State Key Laboratory of Surface Physics, Institute of Nanoelectronics and Quantum Computing,
and Department of Physics, Fudan University, Shanghai 200438, China
2
Shanghai Qi Zhi Institute, Shanghai 200030, China
3
Shanghai Research Center for Quantum Sciences, Shanghai 201315, China

(Received 29 September 2021; accepted 15 September 2022; published 5 October 2022)

The study of dynamical phase transitions has been attracting considerable research efforts in the last decade.
One theme of present interest is to search for exotic scenarios beyond the framework of equilibrium phase
transitions. Here, we establish a duality between many-body dynamics and static Hamiltonian ground states
for both classical and quantum systems. We construct frustration-free Hamiltonians whose ground-state phase
transitions have rigorous duality to chaotic transitions in dynamical systems. By this duality, we show that
the corresponding ground-state phase transitions are characterized by a bulk-to-surface response; these phase
transitions are then dubbed “peratic,” meaning that they are defined by their response to the boundary. For
the classical system, we show how the timelike dimension emerges in the static ground states. For the quantum
system, the ground state is a superposition of geometrical lines on a two-dimensional array; these lines encode the
dynamical Floquet evolution history of one-dimensional disordered spin chains. Our prediction of a peratic phase
transition has direct consequences in quantum simulation platforms such as Rydberg atoms and superconducting
qubits, as well as anisotropic spin glass materials. The discovery would shed light on the unification of dynamical
phase transitions with equilibrium systems.

DOI: 10.1103/PhysRevResearch.4.043009

I. INTRODUCTION Here, we construct frustration-free Hamiltonians whose


ground states have rigorous duality to dynamical systems.
The characterization of different phases of matter is
Through this theoretical construction, we find that a phase
fundamental to our understanding of nature. With rapid
transition mechanism—bulk-to-surface (BTS) response—
developments in controlling many-body states away from
defines a peratic phase transition in Hamiltonian ground
equilibrium in condensed matter and quantum information
states. This mechanism is established by building exact du-
experiments, the study of nonequilibrium phase transitions
ality to order-to-chaos and MBL transitions for classical
has attracted much present research interest. Nonequilib-
and quantum systems, respectively. The Hamiltonian ground
rium many-body physics previously unreachable can now be
states have two distinctive phases as determined by whether
probed in experimental systems such as light-driven electronic
the bulk is rigid against manipulations at the surface. The
matter [1,2] and highly controllable quantum simulation [3–7]
phase transition is characterized by a BTS response
or quantum computing platforms [8–11]. Quantum many-
body localization (MBL) happens for a disordered system χBTS = Var(Obulk )|surf. man. (1)
in its dynamical properties at infinite temperature [12,13].
that is zero or finite, with Obulk  being an observable in
Anomalous protected edge modes at zero Chern number,
the bulk and Var(Obulk ) being the variance of the bulk ob-
which are unexpected for equilibrium systems, could appear
servable as we manipulate the surface. The BTS response
in periodically driven Floquet quantum dynamics [14–16].
quantifies the stability of the bulk against surface manip-
Spontaneous translation symmetry breaking is generalized
ulations, which vanishes if the bulk is stable and acquires
to the temporal domain, giving rise to crystallization in
a finite value otherwise. The phase transition in the classi-
time [7,8,17–20]. There is mounting evidence suggesting that
cal ground state has an exact duality to the order-to-chaos
nonequilibrium phases may contain exotic scenarios beyond
transition in classical nonlinear dynamical systems [21]. The
equilibrium setups.
quantum ground-state phase transition constructed by forming
a superposition of fluctuating line configurations has a rigor-
ous duality with the transition from quantum ergodic [22–24]
*
xiaopeng_li@fudan.edu.cn to MBL [12,13] in quantum many-body dynamics. Unlike
the standard phase transitions described either by spon-
Published by the American Physical Society under the terms of the taneous symmetry breaking [25,26] or by quantum state
Creative Commons Attribution 4.0 International license. Further topology [27–29], there is no change in symmetry or topology
distribution of this work must maintain attribution to the author(s) across the peratic phase transition. Our theory implies that
and the published article’s title, journal citation, and DOI. the study of unconventional phases in nonequilibrium systems

2643-1564/2022/4(4)/043009(9) 043009-1 Published by the American Physical Society


QIU, WANG, XIA, AND LI PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

which can all be satisfied. The Hamiltonian is thus frustration-


free. The binary degrees of freedom in the Hamiltonian
ground state are given by zigj . We then observe that the j
dependence of the binary variables in the static ground state
is timelike, as the jth layer is completely determined by the
( j − 1)th layer. A timelike dimension thus arises in the static
ground state.
By treating the j axis as a time evolution direction, Eq. (3)
maps onto classical nonlinear dynamics having an order-to-
chaos dynamical phase transition [21]. In the ordered phase
(σ < σc ), the dynamical state is pinned by the local field
FIG. 1. Classical peratic phase transition. (a) The BTS response u j : The difference between two initial states as measured by
χBTS across the phase transition for I = 500. It vanishes at small Hamming distance is quickly washed away in the dynamical
coupling variance σ and becomes finite across the phase transition. evolution. In the chaotic phase (σ > σc ), the difference in the
The BTS response is averaged over random surface terms hsurf , with initial states would either remain or gain amplification by the
p being the probability of each surface term hsurf,i taking a positive nonlinear dynamics. Consequently, the ground state of the
value. The inset in (a) shows its first derivative for I = 100, 300, Hamiltonian in Eq. (2) also has two phases. For σ < σc , the
500 with p = 0.5. This derivative develops a peak near the transition system is in a rigid phase where the bulk is stable and immune
point, which sharpens up as we increase the system size. This implies to surface manipulations with different hsurf . For σ > σc , we
a divergent second derivative in the thermodynamic limit. (b) The de- have a volatile phase with the bulk sensitive to surface ma-
pendence of the local BTS response χ j on the j index. This quantity
nipulations. Quantitatively, the BTS response for this specific
has an exponential decay at small σ when χBTS vanishes. The decay
system is defined as
becomes a power-law decay at the critical point σc . Here, we choose  
J = 5I and M = 2. The results are calculated by averaging over 104 1 
samples, and the statistical data error is smaller than the symbol size. χBTS = lim Var(zi j )|hsurf , (4)
hsurf →0 IJ
ij
would inspire discovery of more exotic equilibrium phases
and transitions rather than reaching completely beyond. with the variance Var(zi j ) = 1 − zi j 2 and the average zi j 
obtained by randomly sampling hsurf . This BTS response
quantifies the degree of bulk fluctuations induced by perturba-
II. EMERGENT CHAOTIC DYNAMICS IN A tions at the boundary and reflects the chaotic structures of the
HAMILTONIAN GROUND STATE dual dynamics. We also introduce a space-resolved local BTS
We first consider a discrete classical system which contains response χ j = I −1 i Var(zi j )|hsurf to diagnose the criticality.
binary degrees of freedom zi j = ± on a two-dimensional (2D) Their behavior across the phase transition is shown in Fig. 1.
grid, with the indices i ∈ [0, I − 1] and j ∈ [0, J − 1]. Our The BTS response vanishes in the rigid phase and becomes
theory starts from constructing a Hamiltonian ground state finite in the volatile phase, defining our peratic phase transi-
that supports the phase transition scenario described by the tion. The BTS response is consistent with the bulk entropy
BTS response in Eq. (1). The Hamiltonian contains a bulk and of the system (see Appendix A). By increasing the system
a surface term, Hbulk and Hsurf , size, we find that the BTS response has a divergent second
  derivative at the critical point. The local BTS response shows
 M
an exponential decay in the rigid phase, with a decay length
Hbulk = − zi j sgn u j + w [imj] zi+m, j−1 ,
that tends to diverge approaching the critical point. At the
i j>0 m=−M
 critical point, the local BTS response exhibits a power-law
Hsurf = − [i]
hsurf zi,0 , (2) scaling,
i χ j ∝ j −η , (5)
[i j]
where w represents Ising couplings, u j is a local field in the a signature of nontrivial criticality for the peratic phase tran-
[i]
bulk, and hsurf is a local field on the surface introduced to study sition. We argue that the critical behavior is described by a
the BTS response. The Ising couplings are randomly drawn universal finite-size scaling function
from a Gaussian distribution with zero mean and variance
σ 2 . The bulk local field u j takes random binary values ±1 χBTS (t, L) = L −η G(Lt ν ), (6)
with equal probability. We adopt an open boundary condition with I, J ∝ L, t = (σ − σc )/σc , and determine the anomalous
along the j axis and a periodic boundary along the other axis dimension η = 0.172(2) and ν exponent ν = 2.73(8) (see Ap-
(Fig. 1). The system has a layered structure along the j axis: pendix B). This phase transition scenario does not rely on the
We have interlayer couplings only. symmetry or the topology of the system, in sharp contrast to
Minimizing each term of the Hamiltonian in Eq. (2) leads the standard phase transitions. Whether it can be described by
to a set of equations, the replica type of spontaneous symmetry breaking [30–32]
  as established in disordered spin systems [33–35] is worth
g  [i]  g
M
[i j] g
zi0 = sgn hsurf , zi, j>0 = sgn u j + w m zi+m, j−1 , future investigation. We emphasize that the chaotic structures
m=−M or the BTS response reported here for a highly anisotropic
(3) disordered spin system may shed light on our understanding

043009-2
PERATIC PHASE TRANSITION BY BULK-TO-SURFACE … PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

FIG. 2. The peratic phase transition in the frustrated Ising model FIG. 3. Peratic phase transition with Rydberg atom arrays.
[Eq. (7)]. (a) The classical peratic phase transition. The dashed (a) Schematic illustration of the Rydberg system. The atoms are
lines show the results of the frustration-free model for comparison. dressed with a Rydberg p-wave state and located on a 2D square
(b) The quantum peratic phase transition of the frustrated Ising model lattice, whose lattice constant is a. (b) The BTS response across the
adding the transverse field hT . We choose a field strength hT = 10. peratic phase transition. The strengths of longitudinal fields are fixed
The BTS response is averaged over different surface terms with along the i axis and randomly drawn from the binary values ±V with
p = 0.1, 0.3, 0.5. In this plot, we choose I = 5, J = 5, and M = 2. equal probability along the j axis. The BTS response is averaged over
The results are calculated by averaging over 104 samples, and the all configurations of the different longitudinal fields and different
statistical data error is smaller than the symbol size. surface terms, with p = 0.3, 0.4, 0.5. Here, we set the transverse
field strength hT = 0.5V and the system size as 5 × 5. The results are
calculated by averaging over all possible surface spin polarizations.
of anisotropic spin glasses [33,36,37], for which exact models
are lacking, to our knowledge.
One fascinating property of the volatile phase is its ro- The two-body Hamiltonian also allows a natural way to
bustness: The BTS response is stable irrespective of different generalize the phase transition to quantum systems, by simply
choices of surface manipulations. In our numerical calcula- promoting the Ising variables zi j to Pauli operators ẑi j and
tions, each surface term hsurf,i takes a positive value with a adding a transverse field coupling H = i j hT x̂i j (x̂ being
probability p and a negative value with 1 − p. The BTS re- the Pauli-x operator) to generate quantum fluctuations. As
sponse in the volatile phase as constructed above does not vary shown in Fig. 2(b), the peratic phase transition still persists
with the p value [Fig. 1(b)], i.e., having robustness against in the presence of quantum fluctuations. We find that quantum
different surface manipulations. This nontrivial property can effects further stabilize the rigid phase and shift the transition
be attributed to the presence of a dynamical fixed point of point towards the volatile phase, which is somewhat counter-
the Hamming distance evolution in the dual chaotic dynam- intuitive. This can be attributed to the fact that the transverse
ics [38]. This makes the volatile phase sharply distinctive from field couples the degenerate ground states and develops a
a trivial case with the bulk trivially determined by the surface, tendency for gap opening, rendering the bulk more rigid.
for example, with zi j = zi0 , where the BTS response would As a concrete experimental candidate, we consider a sys-
strongly depend on p. tem of Rydberg p-wave dressed atoms, which has been
used to construct quantum spin ice Hamiltonians [41]
and programmable quantum annealing [42]. Two atomic
III. EXPERIMENTAL CANDIDATES
hyperfine states, |↑ = |52 S1/2 , F = 2, mF = 0 and |↓ =
Although the frustration-free Hamiltonian in Eq. (2) |52 S1/2 , F = 1, mF = 0, of 87 Rb atoms are selected to form
has the nice property of being polynomially solvable, its a spin-1/2 lattice with I rows and J columns. Their couplings
experimental realization is challenging. For experimental re- are controllable by performing a microwave-induced transi-
alization, we further consider a frustrated Hamiltonian with tion, which is described by a transverse field HT = r hT x̂r ,
two-body Ising couplings only, with the field strength hT determined by the Rabi frequency
  of the microwave. Taking a Rydberg p-wave dressing scheme
 
M
where the |↑ state is selectively dressed with a Rydberg
Hbulk = − zi j u j + wm zi+m, j−1 , (7)
[i j]
p-wave state |n2 P3/2 , m = 3/2 [the quantization axis along
i j>0 m=−M
the i direction in Fig. 3(a)], we introduce interactions between
which could describe a broad range of spin systems from Ry- neighboring layers as labeled by j [41]. One key feature of this
dberg atomic systems [39] and superconducting qubits [40] to system is that it has interlayer interactions but no intralayer
anisotropic spin glasses [33]. We study its ground-state phase interactions, and thus the description of this system closely
transition by numerically minimizing the energy. The results resembles Eq. (7). Arranging the atoms periodically in a 2D
are shown in Fig. 2(a). We observe that the BTS responses array as shown in Fig. 3, the resultant interaction between the
for the frustrated and frustration-free models are quite similar two qubits at r and r is given by
to each other with a tiny difference that is barely noticeable.  V sin4 θrr
The peratic phase transition is still preserved in the frustrated HR = (ẑr + 1)(ẑr + 1)/2, (8)
model. r,r
1 + (|r − r |/rc )6

043009-3
QIU, WANG, XIA, AND LI PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

FIG. 4. Quantum peratic phase transition. (a) Schematic illustration of the frustration-free qudit model. The model contains a 2D array of
qudits, with its four states marked by “·” and “ .” The two “ ” states are further specified as “ ” and “⊗.” (b) The forbidden configurations.
The corresponding projector Hamiltonian is provided in Appendix C. The consequent low-energy subspace corresponds to the two types of
line configurations as illustrated in (a) (see main text). (c) The BTS response across the quantum peratic phase transition. Here, we sample the
surface terms hsurf with different p values from 0.1 to 0.5. The quantum ground-state phase transition as constructed is dual to the dynamical
phase transition from quantum ergodic to many-body localized. At small σ , the bulk is robust against surface manipulations giving a vanishing
BTS, which is dual to the dynamical quantum ergodic phase. Above a certain threshold at σ > σc , the bulk is stringently tied with the surface,
corresponding to the dynamical MBL phase. The results are calculated by averaging over 100 samples, with the standard deviations illustrated
by the shaded error bands. The inset in (c) shows the system size dependence of the BTS response (with p fixed at 0.5), which indicates that
the phase transition becomes sharper at larger system size.

with the coupling strength V and the Rydberg interaction We consider a 2D qudit lattice model with a four-
range rc determined by the Rabi frequency and the detun- dimensional local Hilbert space. The four levels are labeled
ing of the one-photon transition of the Rydberg dressing as |νz  with ν (= 0, 1) and z (= ±1) [Fig. 4(a)]. The lattice
scheme [41]. The on-site longitudinal field is tunable by ma- contains I rows and J rungs, with two types of rungs labeled
nipulating the detuning of the microwave with respect to the by “A” and “B.” Different qudits are labeled according to their
hyperfine splitting of 6.8 GHz. The longitudinal fields are position on the lattice by (i, j), with i ( j) being the row (rung)
described by HL = r hr ẑr . index. The large 4IJ -dimensional Hilbert space of the qudit
As shown in Fig. 3(b), the BTS response shows distinctive system is constrained to a low-energy subspace by introducing
behaviors at small and large Rydberg interaction range rc . local projectors. The forbidden configurations are illustrated
At small rc , the system is in a rigid phase having a vanish- in Fig. 4(b), with the corresponding Hamiltonian realization
ing χBTS with bulk spin polarization robust against different given in Appendix C. With the configurations forbidden by
choices of surface polarizations. When rc is larger than a the horizontal rules, there is at most only one “·” in one row.
certain threshold (roughly 0.8 in units of the lattice con- All sites on the left (right) of the “·” have to be “⊗” (“ ”).
stant), the bulk becomes volatile, fluctuating with different By the vertex rules, the “·” sites on each rung have to form
surface polarizations, as characterized by a finite χBTS . The a continuous line, which can either go straight on the same
Rydberg system thus supports a peratic phase transition char- rung or bend over to its nearby rungs. By the vertical rules
acterized by the BTS response. Since the two spin states in on the A (B) rungs, the continuous line has to reach to the
our proposed setup have a hyperfine splitting on the order bottom (top) on the A (B) rungs. With all these constraints,
of gigahertz, the peratic phase transition can be detected by the allowed states in the low-energy subspace correspond to
microwave spectroscopy, which is a standard technique in the two types of continuous lines—type A and type B, as il-
cold-atom experiments. Local addressability can be reached lustrated in Fig. 4(a). The type-A (type-B) line starts from the
by creating spatially resolved Stark shifts using focused laser bottom (top) at A (B) sites, goes straight upward (downward),
fields [43]. bends at most once to the right, and then continues upward
(downward) following the B (A) rungs to the top (bottom).
There are a total number of N = I (J − 1) + 1 such lines,
IV. QUANTUM PERATIC PHASE TRANSITION
with each line uniquely labeled by (i, j), according to the
site where the line bends over, or the last site if the line does
We now provide a rigorous quantum model supporting not bend. The quantum state on the line is |z10 z11 ·· ·· ·· zI−1
1
,
the quantum peratic phase transition. This is achieved by with the subscripts labeled by the order from top to bottom
constructing a quantum frustration-free Hamiltonian, whose on the lattice. The qudits to its left (right) all reside on the
ground-state phase transition has an exact duality with the 0
state |−1  (|01). All the low-energy states can then be labeled
dynamical MBL transition of the Floquet quantum dynamics
of a one-dimensional disordered spin chain [12,13]. as |i, j; z [z ≡ (z0 , z1 , . . . , zI−1 )].

043009-4
PERATIC PHASE TRANSITION BY BULK-TO-SURFACE … PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

We construct a 2D local qudit Hamiltonian (see Ap- rendering a vanishing BTS response for the 2D ground state.
pendix C), whose projection on the low-energy subspace takes The numerical results are provided in Fig. 4(c) and agree well
the form of a Feynman-Kitaev clock Hamiltonian [44,45], with the theoretical analysis. The BTS response thus defines
 a peratic quantum phase transition in the frustration-free
1
Heff = − (|γ0 (z )γ0 (z )| − |γN−1 (z )γN−1 (z )|) quantum ground states. Since the existence of the MBL phase
2 has been proven for one-dimensional systems by Imbrie [47],
z
our constructed exact duality establishes a rigorous scenario

N−2
1 for the quantum peratic phase transition.
+ (|γl (z )γl (z )|− |γl (z )γl+1 (z )|+ H.c.) .
2 With the exact construction presented above and the nu-
l=0
merical results for the transverse field Ising model, we expect
(9) the quantum peratic phase transition to be generic, not relying
Here, the γ states are |γl (z ) = z ψl (z)|l; z, and the se- on the duality with the MBL-to-ergodic phase transition. The
quential index l is introduced for a compact representation of unconventional phase transition defined by the bulk-to-surface
(i, j)—li j = jI + i for j ∈ A and of li j = ( j + 1)I − i − 1 for response could arise in a broad range of quantum simulation
j ∈ B (see one explicit example in Appendix C, Fig. 7). platforms as well as anisotropic spin glass materials.
The wave functions ψl (z) are defined through a sequential
unitary transformation starting from ψ0 (z) = δzz . The update V. CONCLUSION AND OUTLOOK
of ψl is designed to follow the Floquet quantum dynamics of
a one-dimensional spin-1/2 system [46]. From l = 0 to l = We propose a peratic phase transition defined by a bulk-to-
I − 1, the update of ψl corresponds to a unitary gate e−ιx̂i x̂i+1 /10 surface response in both classical and quantum ground states.
(i from 0 to I − 1), where ι is the imaginary unit, and these By constructing frustration-free models, we establish rigorous
unitary gates are then applied in backward order from l = I duality from the peratic phase transition to the order-to-chaos
to 2I − 1. Then in the same order, we apply the unitary gates transition in the classical system, and to the MBL-to-ergodic
e−ιŷi ŷi+1 /10 from l = 2I to l = 4I − 1, and e−ι(ẑi ẑi+1 /10+δi ẑi ) from transition in the quantum setting. With numerical results, we
l = 4I to 6I − 1. This unitary update process is then repeated show that the peratic phase transition is also preserved in
forward for J/6 periods. The amplitude δi is a random num- anisotropic spin glass models with two-body Ising couplings
ber drawn from a Gaussian distribution with zero mean and only. We predict that the system of Rydberg p-wave dressed
variance σ 2 . atoms supports the peratic phase transition. Our theory im-
The constructed 2D local qudit Hamiltonian is semipos- plies that dynamical phase transitions would inspire exotic
itive definite and frustration-free [44], and its ground state scenarios in equilibrium systems rather than reaching beyond.
is |G(z ) = √1N l |γl (z ), having a 2I -fold degeneracy as Our approach also provides an alternative way to characterize
labeled by z . The ground state is an equal-amplitude quan- dynamical phase transitions from the perspective of equi-
tum superposition of those geometrical line configurations in librium phase transitions, which could unify the description
Fig. 4(a). of nonequilibrium phases within the equilibrium framework,
We introduce a bulk observable to diagnose the peratic as the constructed duality from dynamical phases to static
phase transition, Hamiltonian ground states is quite generic (see Appendix D).
 
1 1 1 1
Oi j = − , (10) ACKNOWLEDGMENTS
1 1 −1 −1
We would like to thank Wei Wang for suggesting the
that acts on the site (i, j). The ground-state degeneracy would
ancient Greek “péras” for naming the phase transition and
be lifted up by adding a perturbation on the edge H =
Dong-Ling Deng and Meng Cheng for helpful discussions.
i hsurf [1 − Oi,0 ]: Different choices of hsurf select ground
[i]
This work is supported by National Program on Key Basic
states |G(z ) with different z . The corresponding BTS re-
Research Project of China (Grants No. 2021YFA1400900 and
sponse is
  No. 2017YFA0304204), National Natural Science Foundation
1  2 of China (Grants No. 11774067 and No. 11934002), Shanghai
χBTS = lim Oi j  − Oi j  |hsurf ,
2
Municipal Science and Technology Major Project (Grant No.
hsurf →0 I
ij 2019SHZDZX01), and Shanghai Science Foundation (Grants
No. 21QA1400500 and No. 19ZR1471500). X.Q. acknowl-
given by χBTS = 1
JI 2 i,l {( z zi |ψl (z)|
2 )2 −
2
edges support from National Natural Science Foundation of
( z zi |ψl }, where · · · denotes averaging over different
(z)|2 ) China (Grant No. 12104098).
z . Treating the unitary update from l to l + 1 as quantum time
evolution, the dynamics of ψl (z) has two distinctive phases:
APPENDIX A: RELATION BETWEEN BULK-TO-SURFACE
quantum ergodic and MBL. For the MBL dynamics, the
RESPONSE AND BULK FLUCTUATIONS
wave function holds the memory of the initial configuration
of z , and the resultant BTS response χBTS of the dual 2D In this Appendix, we show that the bulk-to-surface (BTS)
quantum ground state is finite and strongly depends on z , response reflects the entropy of bulk fluctuations in the sys-
[i]
namely, the p value in sampling hsurf . In contrast, for the tem. To probe the bulk phase transition directly in our model,
quantum ergodic dynamics, the local observables would we define a half-system entropy S = − σ P(σ ) log2 P(σ ),
thermalize as l proceeds and are then independent of z , with σ indexing the configuration of one-half of the system

043009-5
QIU, WANG, XIA, AND LI PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

rms
FIG. 6. Finite-size scaling analysis for the classical peratic phase
transition. (a) Power-law decay of the local bulk-to-surface (BTS)
response in χ j and the η exponent. This exponent is obtained by
fitting our numerical results for χ j to a power law at the critical point.
The inset shows the rms error for fitting χ j near the critical point to
a power-law function, and the location of the minimum indicates the
critical point. In this plot, we choose I = 500, J = 5I, p = 0.5, and
M = 2. The η exponent is obtained to be 0.172(2). (b) Data collapse
FIG. 5. Phase transition in the frustration-free Ising model [see in the scaling analysis for the BTS response. Here, we choose M = 2,
Eq. (2) in the main text]. Both the half-system entropy and the BTS p = 0.5, L = I, J = 5I. In (b), we take ν = 2.732, which gives the
response can determine two distinctive phases. The inset shows the best-quality data collapse.
linear scaling of the entropy with the system size I, where we set σ =
104 . The results are averaged over 5 × 104 random samples. Here, we group fixed point [30]. Taking a renormalization group trans-
choose I = 16, J = 5I, p = 0.5, and M = 2. formation, t → tζ 1/ν , L → L/ζ (the system size I, J ∝ L),
with ζ being a scaling factor and ν being a critical exponent,
and P(σ ) being the corresponding probability. Its behavior we have
is shown in Fig. 5. We find that the half-system entropy is
χ j/ζ (tζ 1/ν , L/ζ ) = ζ η χ j (t, L) (B2)
vanishing and subextensive on the two sides of the peratic
phase transition. Here, “subextensive” means that the entropy in the neighborhood of the fixed point. Here, η is the anoma-
increases linearly with the system size I instead of I × J. Both lous dimension. The ν and η exponents are to be fixed with
the BTS response and the half-system entropy characterize our numerical results. The scaling form implies that
the fluctuations of the bulk system and serve properly as an
order parameter for the peratic phase transition. However, χ j (t, L) = j −η A( jt ν , Lt ν ), (B3)
for our frustration-free model, it is so much more convenient
to compute the BTS response than the entropy. Computing with A(· · · ) being a universal function. It then follows directly
the BTS response takes polynomial time, whereas the time it that a thermodynamic limit system right at the peratic phase
takes to compute the entropy scales exponentially. The BTS transition point has a power-law local BTS response,
response is thus more convenient to diagnose peratic phase
χ j ∝ j −η . (B4)
transition.
The scaling of the total BTS response obeys
APPENDIX B: PHENOMENOLOGICAL THEORY
AND SCALING ANALYSIS χBTS (tζ 1/ν , L/ζ ) = ζ η χBTS (t, L), (B5)

To analyze the peratic phase transition, we propose a phe- which implies


nomenological theory and perform scaling analysis. Since
the BTS response that characterizes the phase transition also χBTS (t, L) = L −η G(Lt ν ), (B6)
probes the degree of fluctuations in the bulk system, resem- with G(· · · ) being a universal function.
bling the entropy, the phenomenological free energy near the The η exponent is determined by fitting our numerical
phase transition takes the form results to χ j ∝ j −η at the critical point [Fig. 6(a)], from which
 3  we get η = 0.172(2). We first fit all numerical results near
f (t, χBTS ) = −r(t )χBTS + κ (t )χBTS
2
+ O χBTS , (B1)
the critical point for χ j to a power law and locate the critical
where we have t = (σ − σc )/σc , r > 0 (r < 0) for t > 0 (t < point by minimizing the root-mean-square (rms) error of the
0), and κ > 0. This free energy does not have Ising symmetry. fitting. The value of the η exponent and its error are obtained
Nonetheless, minimizing the free energy under the physical by fitting the data right at the critical point.
constraint χBTS  0 produces a phase transition, i.e., the per- The ν exponent is extracted by performing a data collapse
atic phase transition established in the main text. taking the scaling form in Eq. (B6) [Fig. 6(b)], and we get
The scaling analysis is performed by postulating that the ν = 2.73(8). The ν exponent is calculated using the analysis
phase transition is described by a certain renormalization in Ref. [48], with the error estimated by bootstrap.

043009-6
PERATIC PHASE TRANSITION BY BULK-TO-SURFACE … PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

FIG. 7. Line configurations for the quantum frustration-free qudit model. These line configurations represent the allowed states in the
low-energy subspace after enforcing the local projectors. Here, we choose I = 3 and J = 4 for illustration.

APPENDIX C: CONSTRUCTION OF A Projector P4 contains four-body interactions acting on the sites


FRUSTRATION-FREE QUANTUM MODEL {(i, j − 1), (i, j), (i, j + 1), (i ± 1, j)},
We consider a 2D array of qudits. Its Hilbert space is   0 0 0 1 0 0 0 1
P4 = .
projected to a low-energy subspace by introducing an energy z1 z2 z3 z4 z1 z2 z3 z4
i j z1 z2 z3 z4
penalty
The low-energy subspace (|low ) is defined by
HP = λ(P1 + P2 + P3 + P4 ), P1,2,3,4 |low  = 0. The forbidden configurations by the
with P1,2,3,4 being four local projectors as described below. above Hamiltonian projectors are illustrated in Fig. 4(b) in
Projector P1 acts on columnwise nearby sites (i, j) and (i + the main text. The low-energy quantum states in the subspace
1, j) (see Fig. 4 in the main text for an illustration), correspond to line configurations as shown in Fig. 4(a) in the
  main text. We assume that the strength of the energy penalty
  1 0 1 0  0 1 0 1 λ is large enough to suppress all possible fluctuations leaving
P1 = + . the subspace.
z z z z z z z z
zz i, j∈A i, j∈B
We now construct a microscopic local Hamiltonian that
Projectors P2 and P3 both act on row-wise nearby sites (i, j) produces the effective Hamiltonian for low-energy quantum
and (i, j + 1), states in Eq. (9) in the main text. The microscopic Hamilto-
  nian takes the form
  1 1 1 1 0 0 0

0  
P2 = + , H= HiAj + HiBj , (C1)
z z z z 1 −1 1 −1
ij zz i, j∈A i, j∈B
   
  1 0 1 0 0 1 0 1 with HiAj and HiBj being three-local interactions acting on the
P3 = + . sites {(i j), (i + 1, j), (i, j + 1)} and {(i − 1, j), (i, j), (i, j +
z −1 z −1 1 z 1 z
ij z 1)}, respectively, defined by

 0    
1 1 0 1 1 1  0 1 1 1 1 0
HiAj = − Uz[i1 zj]2 ;z z + H.c. ,
−1 z1 z2 −1 z1 z2 2 1 2 −1 z2 z1 z1 z2 1
z1 z2 z1,2 ,z1,2
 1    
0 1 1 0 1 1  1 0 1 1 1 0
HiBj = − Uz[i1 zj]2 ;z z + H.c. .
z1 −1 z2 z1 −1 z2 2 1 2 z1 −1 z2 z1 z2 1
z1 z2 z1,2 ,z1,2

The interaction elements Uz[i1 zj]2 ;z form a unitary matrix, with The Hamiltonian is frustration-free within the low-energy
1 z2
[i j] [i j]∗ subspace. Taking l as a time step, the unitary update of
= δz1 z1 δz2 z2 . Projecting the microscopic
z1 z2 Uz1 z2 ;z1 z2 Uz1 z2 ;z1 z2 ψl (z) corresponds to unitary time evolution under a series of
Hamiltonian to the low-energy subspace, we reach the effec- two-qubit gates defined by their matrix representation, U [i j] .
tive Hamiltonian provided in the main text, with the γ states One explicit example can be found in Fig. 7; the low-energy
defined according to its wave function ψl (z) by subspace is spanned by the states illustrated by the line con-
 [i j] figurations. Our Hamiltonian construction has been inspired
ψli j (z) = Uzi zi+1 ;z z ψli j −1 (z0 · · · zi−1 zi zi+1 · · · zI−1 ).
i i+1 by the geometrical Feynman-Kitaev clock states, which have
zi zi+1

043009-7
QIU, WANG, XIA, AND LI PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

been used to prove the computational quantum Merlin-Arthur APPENDIX E: THE EXACT GROUND-STATE DUAL
(QMA) completeness of 2D quantum lattice models [44,45]. OF THE TIME CRYSTAL
In this Appendix, we show that the nonequilibrium time
APPENDIX D: MAPPING BINARY VARIABLE DYNAMICS crystal phase transition [7,8,17–20] also has an exact dual to
TO HAMILTONIAN GROUND STATES a phase transition in the ground states of a static Hamiltonian.
We consider a generic dynamical evolution of a binary This Hamiltonian has the same form as that in Eq. (6) in
sequence zt ≡ [z0,t , z1,t , . . . , zI−1,t ]T (a column vector), with t the main text, but with the γ states |γl (z ) = z ψl (z)|l; z,
being the evolution time. A generic binary variable dynamics updated in a different way. Here, for the time crystal, we
is described by a dynamical equation, update the wave functions ψl (z) through a Floquet time crystal
dynamics of a one-dimensional (1D) spin-1/2 system [20].
zt = sgn[Ft (zt−1 )], (D1) From l = 0 to l = I − 1, the update of ψl corresponds to the
unitary gates eιx̂i hx (i from 0 to I − 1), and then the unitary
with Ft being an arbitrary function. A Turing complete binary
gates e−ι(J ẑi ẑi+1 +hz ẑi +hx x̂i ) (i from I − 1 to 0) are applied from
[i] [i] [i]

circuit can also be formulated in this form. The time-evolved


configuration is the ground state of a static Hamiltonian, l = I to 2I − 1. The amplitudes J [i] , hz[i] , and hx[i] are chosen
 from certain random distributions. This unitary update process
H =− ztT sgn[Ft (zt−1 )], (D2) is then repeated forward for J/2 periods.
t In Appendix C, we have constructed a 2D local qudit
Hamiltonian, whose projection on the low-energy subspace
by treating t as a real-space column index. The solution exis-
gives the effective Hamiltonian and whose ground state
tence of Eq. (D1) guarantees that the Hamiltonian ground state
|G(z ) = √1IJ l |γl (z ) encodes the whole Floquet quan-
is frustration-free. If we further impose a local constraint on
the binary variable dynamics, namely, assuming that zi,t is de- tum dynamics. From the construction, we observe that this
termined by its neighboring binary variables zi ∈[i−M,i+M],t−1 local qudit Hamiltonian has lattice translation symmetry
only (M is a finite number), the local dynamical equation takes (LTS) along the j axis, with A and B sites forming a sublattice
the form structure. Note that this LTS is dual to the time-translation
symmetry (TTS) of the Floquet quantum dynamics. The time
zi,t = sgn[Fi,t ({zi ∈[i−M,i+M],t−1 })]. (D3) crystal phase transition arising from TTS breaking has an ex-
act dual to the phase transition in the ground state |G(z ) due
The Hamiltonian in Eq. (D2) is then reduced to
to LTS breaking. The operator given in Eq. (8) in the main text

H =− zi,t sgn[Fi,t ({zi ∈[i−M,i+M],t−1 })], (D4) also characterizes this symmetry breaking. The corresponding
i,t
order parameter reads
 
which only contains local couplings on a 2D grid. We thus 1 
C(O) = (−1) j/2 G(z )|Oi j |G(z ) . (E1)
conclude that generic local Ising variable dynamics can be I i, j∈A
mapped to ground states of a local Hamiltonian. We expect
that this holds for arbitrary dynamical evolution in discrete In the thermodynamic limit, this order parameter vanishes in
settings, as the discrete degrees of freedom can be rigorously the symmetric phase and becomes finite in the LTS-broken
encoded by Ising variables. phase.

[1] J. Demsar, Non-equilibrium phenomena in superconductors [6] J. Smith, A. Lee, P. Richerme, B. Neyenhuis, P. W. Hess, P.
probed by femtosecond time-domain spectroscopy, J. Low Hauke, M. Heyl, D. A. Huse, and C. Monroe, Many-body
Temp. Phys. 201, 676 (2020). localization in a quantum simulator with programmable random
[2] N. Rivera and I. Kaminer, Light–matter interactions with pho- disorder, Nat. Phys. 12, 907 (2016).
tonic quasiparticles, Nat. Rev. Phys. 2, 538 (2020). [7] S. Choi, J. Choi, R. Landig, G. Kucsko, H. Zhou, J. Isoya, F.
[3] E. Altman, K. R. Brown, G. Carleo, L. D. Carr, E. Demler, C. Jelezko, S. Onoda, H. Sumiya, V. Khemani, C. von Keyserlingk,
Chin, B. DeMarco, S. E. Economou, M. A. Eriksson, K.-M. C. N. Y. Yao, E. Demler, and M. D. Lukin, Observation of discrete
Fu, M. Greiner, K. R. Hazzard, R. G. Hulet, A. J. Kollár, time-crystalline order in a disordered dipolar many-body sys-
B. L. Lev, M. D. Lukin, R. Ma, X. Mi, S. Misra, C. Monroe tem, Nature (London) 543, 221 (2017).
et al., Quantum simulators: Architectures and opportunities, [8] J. Zhang, P. Hess, A. Kyprianidis, P. Becker, A. Lee, J. Smith,
PRX Quantum 2, 017003 (2021). G. Pagano, I.-D. Potirniche, A. C. Potter, A. Vishwanath, N. Y.
[4] M. Schreiber, S. S. Hodgman, P. Bordia, H. P. Lüschen, Yao, and C. Monroe, Observation of a discrete time crystal,
M. H. Fischer, R. Vosk, E. Altman, U. Schneider, and I. Nature (London) 543, 217 (2017).
Bloch, Observation of many-body localization of interacting [9] P. Roushan, C. Neill, J. Tangpanitanon, V. M. Bastidas,
fermions in a quasirandom optical lattice, Science 349, 842 A. Megrant, R. Barends, Y. Chen, Z. Chen, B. Chiaro, A.
(2015). Dunsworth, A. Fowler, B. Foxen, M. Giustina, E. Jeffrey, J.
[5] S. Deng, Z.-Y. Shi, P. Diao, Q. Yu, H. Zhai, R. Qi, and H. Kelly, E. Lucero, J. Mutus, M. Neeley, C. Quintana, D. Sank
Wu, Observation of the Efimovian expansion in scale-invariant et al., Spectroscopic signatures of localization with interacting
Fermi gases, Science 353, 371 (2016). photons in superconducting qubits, Science 358, 1175 (2017).

043009-8
PERATIC PHASE TRANSITION BY BULK-TO-SURFACE … PHYSICAL REVIEW RESEARCH 4, 043009 (2022)

[10] Z. Yan, Y.-R. Zhang, M. Gong, Y. Wu, Y. Zheng, S. Li, C. [31] S. F. Edwards and P. W. Anderson, Theory of spin glasses,
Wang, F. Liang, J. Lin, Y. Xu, C. Guo, L. Sun, C.-Z. Peng, J. Phys. F: Met. Phys. 5, 965 (1975).
K. Xia, H. Deng, H. Rong, J. Q. You, F. Nori, H. Fan, X. [32] G. Parisi, Infinite Number of Order Parameters for Spin-
Zhu et al., Strongly correlated quantum walks with a 12-qubit Glasses, Phys. Rev. Lett. 43, 1754 (1979).
superconducting processor, Science 364, 753 (2019). [33] K. Binder and A. P. Young, Spin glasses: Experimental facts,
[11] K. Xu, Z.-H. Sun, W. Liu, Y.-R. Zhang, H. Li, H. Dong, W. Ren, theoretical concepts, and open questions, Rev. Mod. Phys. 58,
P. Zhang, F. Nori, D. Zheng, H. Fan, and H. Wang, Probing 801 (1986).
dynamical phase transitions with a superconducting quantum [34] D. Pierangeli, A. Tavani, F. Di Mei, A. J. Agranat, C. Conti,
simulator, Sci. Adv. 6, eaba4935 (2020). and E. DelRe, Observation of replica symmetry breaking in
[12] R. Nandkishore and D. A. Huse, Many-body localization and disordered nonlinear wave propagation, Nat. Commun. 8, 1501
thermalization in quantum statistical mechanics, Annu. Rev. (2017).
Condens. Matter Phys. 6, 15 (2015). [35] R. Harris, Y. Sato, A. J. Berkley, M. Reis, F. Altomare, M. H.
[13] D. A. Abanin, E. Altman, I. Bloch, and M. Serbyn, Colloquium: Amin, K. Boothby, P. Bunyk, C. Deng, C. Enderud, S. Huang,
Many-body localization, thermalization, and entanglement, E. Hoskinson, M. W. Johnson, E. Ladizinsky, N. Ladizinsky, T.
Rev. Mod. Phys. 91, 021001 (2019). Lanting, R. Li, T. Medina, R. Molavi, R. Neufeld et al., Phase
[14] T. Kitagawa, E. Berg, M. Rudner, and E. Demler, Topological transitions in a programmable quantum spin glass simulator,
characterization of periodically driven quantum systems, Phys. Science 361, 162 (2018).
Rev. B 82, 235114 (2010). [36] U. Atzmony, E. Gurewitz, M. Melamud, H. Pinto, H. Shaked,
[15] L. Jiang, T. Kitagawa, J. Alicea, A. R. Akhmerov, D. Pekker, G. Gorodetsky, E. Hermon, R. M. Hornreich, S. Shtrikman,
G. Refael, J. I. Cirac, E. Demler, M. D. Lukin, and P. Zoller, and B. Wanklyn, Anisotropic Spin-Glass Behavior in Fe2 TiO5 ,
Majorana Fermions in Equilibrium and in Driven Cold-Atom Phys. Rev. Lett. 43, 782 (1979).
Quantum Wires, Phys. Rev. Lett. 106, 220402 (2011). [37] M. Dragomir, I. Arčon, P. A. Dube, J. C. Beam, A. P.
[16] M. S. Rudner, N. H. Lindner, E. Berg, and M. Levin, Anoma- Grosvenor, G. King, and J. E. Greedan, Family of anisotropic
lous Edge States and the Bulk-Edge Correspondence for spin glasses Ba1–x La1+x MnO4+δ , Phys. Rev. Mater. 5, 074403
Periodically Driven Two-Dimensional Systems, Phys. Rev. X (2021).
3, 031005 (2013). [38] N. Bertschinger and T. Natschläger, Real-time computation at
[17] A. Shapere and F. Wilczek, Classical Time Crystals, Phys. Rev. the edge of chaos in recurrent neural networks, Neural Comput.
Lett. 109, 160402 (2012). 16, 1413 (2004).
[18] F. Wilczek, Quantum Time Crystals, Phys. Rev. Lett. 109, [39] M. Saffman, T. G. Walker, and K. Mølmer, Quantum informa-
160401 (2012). tion with Rydberg atoms, Rev. Mod. Phys. 82, 2313 (2010).
[19] V. Khemani, A. Lazarides, R. Moessner, and S. L. Sondhi, [40] M. Kjaergaard, M. E. Schwartz, J. Braumüller, P. Krantz,
Phase Structure of Driven Quantum Systems, Phys. Rev. Lett. J. I.-J. Wang, S. Gustavsson, and W. D. Oliver, Superconducting
116, 250401 (2016). qubits: Current state of play, Annu. Rev. Condens. Matter Phys.
[20] D. V. Else, B. Bauer, and C. Nayak, Floquet Time Crystals, 11, 369 (2020).
Phys. Rev. Lett. 117, 090402 (2016). [41] A. W. Glaetzle, M. Dalmonte, R. Nath, I. Rousochatzakis, R.
[21] J. Kelso, A. Mandell, and M. Shlesinger, Adaptation toward the Moessner, and P. Zoller, Quantum Spin-Ice and Dimer Models
edge of chaos, in Dynamic Patterns in Complex Systems (World with Rydberg Atoms, Phys. Rev. X 4, 041037 (2014).
Scientific, Singapore, 1988), pp. 293–301. [42] X. Qiu, P. Zoller, and X. Li, Programmable quantum anneal-
[22] J. M. Deutsch, Quantum statistical mechanics in a closed sys- ing architectures with Ising quantum wires, PRX Quantum 1,
tem, Phys. Rev. A 43, 2046 (1991). 020311 (2020).
[23] M. Srednicki, Chaos and quantum thermalization, Phys. Rev. E [43] Y. Wang, A. Kumar, T.-Y. Wu, and D. S. Weiss, Single-qubit
50, 888 (1994). gates based on targeted phase shifts in a 3D neutral atom array,
[24] M. Rigol, V. Dunjko, and M. Olshanii, Thermalization and Science 352, 1562 (2016).
its mechanism for generic isolated quantum systems, Nature [44] A. Y. Kitaev, A. Shen, M. N. Vyalyi, and M. N. Vyalyi,
(London) 452, 854 (2008). Classical and Quantum Computation (American Mathematical
[25] L. D. Landau, On the theory of phase transitions. I, Phys. Z. Society, Providence, RI, 2002).
Sowjetunion 11, 26 (1937). [45] D. Aharonov, W. van Dam, J. Kempe, Z. Landau, S. Lloyd,
[26] S. Weinberg, The Quantum Theory of Fields (Cambridge Uni- and O. Regev, Adiabatic quantum computation is equivalent
versity Press, Cambridge, 1995), Vol. 2. to standard quantum computation, SIAM J. Comput. 37, 166
[27] M. Z. Hasan and C. L. Kane, Colloquium: Topological insula- (2007).
tors, Rev. Mod. Phys. 82, 3045 (2010). [46] P. Ponte, Z. Papić, F. Huveneers, and D. A. Abanin, Many-Body
[28] X.-L. Qi and S.-C. Zhang, Topological insulators and supercon- Localization in Periodically Driven Systems, Phys. Rev. Lett.
ductors, Rev. Mod. Phys. 83, 1057 (2011). 114, 140401 (2015).
[29] X.-G. Wen, Colloquium: Zoo of quantum-topological phases of [47] J. Z. Imbrie, On many-body localization for quantum spin
matter, Rev. Mod. Phys. 89, 041004 (2017). chains, J. Stat. Phys. 163, 998 (2016).
[30] A. Altland and B. D. Simons, Condensed Matter Field Theory [48] S. M. Bhattacharjee and F. Seno, A measure of data collapse for
(Cambridge University Press, Cambridge, 2010). scaling, J. Phys. A: Math. Gen. 34, 6375 (2001).

043009-9

You might also like