Anda di halaman 1dari 7

Persistence of locality in systems with power-law interactions

Zhe-Xuan Gong,
1
Michael Foss-Feig,
1
Spyridon Michalakis,
2
and Alexey V. Gorshkov
1
1
Joint Quantum Institute, NIST/University of Maryland, College Park, Maryland 20742, USA
2
Institute for Quantum Information & Matter, California Institute of Technology, Pasadena, CA 91125, USA
(Dated: January 27, 2014)
Motivated by recent experiments with ultra-cold matter, we derive a new bound on the prop-
agation of information in D-dimensional lattice models exhibiting 1/r

interactions with > D.


The bound contains two terms: One accounts for the short-ranged part of the interactions, giving
rise to a bounded velocity and reecting the persistence of locality out to intermediate distances,
while the other contributes a power-law decay at longer distances. We demonstrate that these two
contributions not only bound but qualitatively reproduce the short- and long-distance dynamical
behavior following a local quench in an XY chain and a transverse-eld Ising chain. In addition to
describing dynamics in numerous intractable long-range interacting lattice models, our results can
be experimentally veried in a variety of ultracold-atomic and solid-state systems.
PACS numbers: 75.10.Pq, 05.70.Ln
Lieb-Robinson bounds [1, 2] on the propagation of
information in many-body quantum systems underly
our understanding of numerous equilibrium and non-
equilibrium phenomena, including entanglement growth
after quenches [35], stability to perturbations of the
area law for entanglement entropy [6, 7], exponential
decay of correlations in gapped ground states [8], and
the Lieb-Schultz-Mattis theorem [9]. While such bounds
are well established for systems with short-range interac-
tions, their generalization to systems with long-range in-
teractions is far from complete [8, 10, 11]. Meanwhile, nu-
merous currently available atomic, molecular, and optical
systems exhibiting long-range interactions are emerging
as versatile platforms for studying quantum many-body
physics both in and out of equilibrium. These long-range
interactions include dipolar (1/r
3
) interactions between
electric [12, 13] or magnetic [1419] dipoles, strong van-
der-Waals (1/r
6
) interactions between Rydberg atoms
[12, 20] or polaritons [21], along with 1/r

and even more


general forms of interactions between trapped ions [22
25] or atoms in multimode cavities [26].
An important consequence of Lieb-Robinson bounds
for short-range interacting systems is an emergent veloc-
ity for the propagation of information, which gives rise to
a linear light cone and the associated notion of locality.
While these bounds have been generalized to long-range
interacting systems by Hastings and Koma [8], it is not
yet clear to what extent such locality persists. For exam-
ple, while the Hastings-Koma bound allows for a causal
region that grows exponentially in time, and thus a di-
vergent velocity (see Fig. 1), to the best of our knowledge
there are no models that explicitly demonstrate such be-
havior. Conversely, linear light cones have been observed
in certain systems with long-ranged interactions [10, 11],
yet are manifestly absent in the existing Hastings-Koma
bound. In this Letter, we derive a new Lieb-Robinson-
type bound for systems with 1/r

interactions, which is
consistent with the Hastings-Koma bound at long dis-
t
r
r
c
e
r
r
c
r

causal
(a) (b)
vt r
vt log r
region
FIG. 1: (Color online). (a) Illustration of the causal region
(shaded) resulting from Eq. (1) for the case v1 = v2 = v. The
boundary switches from linear to logarithmic at a critical rc
satisfying rc log rc. (b) The decay of the signal outside
the causal region changes from exponential to algebraic at rc.
tances, but explicitly captures the possibility for a linear
light cone at intermediate distances. Our results are rel-
evant to recent experiments in trapped ions [24, 25], and
build on recent theoretical work studying post-quench
dynamics in several long-range interacting systems, in-
cluding Ising models both with [10, 27, 28] and without
[11, 29] a transverse eld, the XXZ chain [11, 30], and
spin models with boson-mediated interactions [31].
We consider general spin Hamiltonians of the form
H =
1
2

i=j
h
ij
, dened on an innite D-dimensional
regular cubic lattice (D = 1, 2, 3), where h
ij
is an opera-
tor supported on sites i and j and h
ij
h
ji
[39]. We note
that all of the results presented below apply to Hamilto-
nians with time dependence and/or arbitrary onsite in-
teractions [32]. We further assume that all interactions
are bounded by a power-law decay: |h
ij
| J
ij
1/r

ij
,
where r
ij
is the Euclidean distance (in lattice units) be-
tween sites i and j, and for convenience we dene J
ii
= 1.
The notation |O| represents the norm of the operator O
(the largest magnitude of an eigenvalue of O). In what
follows, we prove that arbitrary operators A and B, sup-
ported on single sites a distance r apart, obey the bound
|[A(t), B]|
2 |A| |B|

c
1
e
v1t
1
e
r
+c
2
e
v2t
1
[(1 )r]

, (1)
a
r
X
i
v
:
1
4
0
1
.
6
1
7
4
v
1


[
q
u
a
n
t
-
p
h
]


2
3

J
a
n

2
0
1
4
2
where A(t) = e
iHt
Ae
iHt
. The constants c
1
, c
2
, v
1
, v
2
are nite for all > D and independent of t and r, while
(0, 1) is an adjustable parameter. As shown in Fig.
1(a), we can dene a causal region as the part of the
r-t plane where the right-hand-side of Eq. (1) is larger
than a given value. The rst term on the right-hand-side
of Eq. (1) is the familiar short-range bound of Ref. [1];
alone it would lead to a causal region bounded by a linear
light cone (v
1
t r), and thus to a nite velocity for the
propagation of information. The second term is similar
to the long-range bound of Ref. [8]; alone it would lead
to a causal region with a logarithmic boundary v
2
t
log r, and an actual velocity that grows exponentially
in time. The two terms together give a hybrid boundary,
which switches from linear to logarithmic behavior at a
critical r
c
satisfying r
c
log r
c
. As shown in Fig. 1(b),
the decay of the signal outside the causal region changes
from exponential to polynomial at r
c
. Note that, as
, and the Hamiltonian approaches a nearest-neighbor
model, r
c
and the bound reduces to the short-range
case.
We emphasize that this hybrid bound cannot be ob-
tained by simply adding a short-range contribution to the
long-range bound of Ref. [8]. The long-range bound [8]
alone, in order to remain valid in the large- limit, has
the property that v
2
diverges as . Therefore, as
shown in Fig. 2(a), it leads to a causal region that grows
larger for shorter-range interactions. To the contrary, in
the long-range piece of Eq. (1), v
2
remains nite in the
large- limit. Thus we obtain a much more physical sce-
nario, in which the causal region shrinks for progressively
shorter-range interactions, eventually coinciding with the
linear light cone familiar from short-range Lieb-Robinson
bounds [Fig. 2(b)].
Equation (1) has many important applications; for ex-
ample it can be used to predict dynamics following a
local quench of a lattice system [25]. Specically, we
envision applying a unitary operator U (acting only on
site j) to an arbitrary state [. The eect of this
quench on the expectation value of an arbitrary oper-
ator A, acting only on site i a distance r from site j,
is captured by the experimentally measurable quantity
Q(t) =

[U

A(t)U[ [A(t)[

. Because Q(t) =

[U

[A(t), U][

|[A(t), U]|, we can bound Q(t) us-


ing Eq. (1).
Long-range Lieb-Robinson bounds. We rst briey
summarize the origin of the Hastings-Koma bound de-
rived in Ref. [8]. For operators A and B supported on
sites i and j, respectively, |[A(t), B]| can be bounded by
an innite series in time [8],
|[A(t), B]|
2 |A| |B|

n=1
(2t)
n
n!

n
(i, j), (2)

n
(i, j)

k1,...,kn1
J
ik1
J
k1k2
. . . J
kn1j
. (3)
t
r
region
(a) (b)
I
n
c
r
e
a
s
i
n
g

I
n
c
r
e
a
s
i
n
g

r
t
causal
region
causal
FIG. 2: (Color online). (a) The long-range bound in Ref. [8]
leads to a logarithmically bounded causal region that expands
with decreasing interaction range (increasing ). (b) The new
bound in Eq. (1) gives rise to a hybrid boundary and a causal
region that contracts with decreasing interaction range, even-
tually converging to a linear light-cone for .
Here
n
(i, j) can be thought of as the total contribution
from all n
th
order hopping processes connecting sites i
and j, and the factor

k
J
ik
is nite for all > D.
Following Ref. [8], one can show that, for any i, j, and
> D, the following reproducibility condition holds

k
J
ik
J
kj
2

J
ij
. (4)
Repeated application of this bound to Eq. (3) yields

n
(i, j) (2

)
n1
/r

, where it has now been assumed


that i ,= j and r r
ij
. Substituting this result into
Eq. (2) yields the Hastings-Koma bound
|[A(t), B]|
2 |A| |B|
c
e
vt
1
r

, (5)
where v = 2
2
2

and c = (2

)
1
. Equation (5) holds
for all > D, so naively one would expect to be able
to recover a short-ranged Lieb-Robinson bound [e.g. the
rst term in Eq. (1)] by taking the limit . How-
ever, because the velocity v in Eq. (5) diverges exponen-
tially with , the causal region encompasses all r and t
for short-range ( ) interactions [Fig. 2(a)]. Below,
we derive the new bound in Eq. (1), which recovers the
correct short-range physics in the large limit, and man-
ifestly preserves the eects of short-range interactions at
intermediate distance scales.
Recovering locality. Our strategy for obtaining the
bound on |[A(t), B]| given in Eq. (1) begins by separat-
ing the innite series in Eq. (2) into two parts:
|[A(t), B]|
2 |A| |B|

[r]1

n=1
(2t)
n
n!

n
(i, j)+

n=[r]
(2t)
n
n!

n
(i, j),
(6)
where is an adjustable parameter satisfying 0 < < 1,
and [r] represents the smallest integer satisfying [r]
r. The intuition for this separation is that the rst
part contains a relatively small number of hops, where
the long-range part of the interactions gives important
contributions, while the second part contains a relatively
3
. . . . . .
. . . . . .
. . . . . .
j i
A B
r
J
ij
n = 1
n = 2
n = 3
FIG. 3: (Color online). The largest magnitude terms con-
tributing to Jn(i, j) for n = 1, 2, 3, shown in 1D for simplicity.
large number of hops, where the short-range part of the
interactions is dominant.
By using the inequality J
kn1j
1, we immediately
obtain
n
(i, j)
n1
. Therefore the second term in
Eq. (6) is bounded by

n=[r]
(2t)
n
n!

n
(i, j)

n=[r]
(2
2
t)
n
n!e
rn
< c
1
e
v1t
1
e
r
, (7)
where v
1
= 2
2
e and c
1
=
1
. For nearest-neighbor
interactions ( ), we can simply take the limit 1
so that [r] = [r], and the rst sum in Eq. (6) vanishes
as one needs at least [r] hops to get from site i to site j.
Hence the bound is directly given by Eq. (7), which is the
usual Lieb-Robinson bound for short-range interactions.
The rst term in Eq. (6) could be bounded using
Eq. (4); however, this would once again lead to a velocity
that diverges with . The origin of this divergence is the
attempt to bound repeated nearest-neighbor hops (which
have unity amplitude for all ) by a single long-range hop
(whose amplitude decreases with ). To resolve this is-
sue, we include the contribution from nearest-neighbor
hops explicitly, thereby arriving at the following modi-
ed reproducibility condition (valid for all i, j) [33]

k
J
ik
J
kj
3
D

r
ik
<2
J
ik
J
kj
. (8)
Here the notation

r
ik
<2
implies a sum over all sites k for
which r
ik
< 2. This bound is a major improvement over
Eq. (4) used for deriving the Hastings-Koma bound, as
it does not contain a coecient that grows exponentially
with . Applying this result iteratively in Eq. (3), we
nd

n
(i, j) (3
D
)
n1

r
ik
1
<2,...,r
k
n2
k
n1
<2
J
i,k1
. . . J
kn1,j
.
(9)
The maximum possible value for each summand is given
by (r n + 1)

, which is achieved by combining n 1


nearest-neighbor hops with one remaining hop of distance
r n + 1 (see Fig. 3) [40]. The number of sites with
r
ik
< 2 for a D-dimensional cubic lattice is bounded by
3
D
, therefore
n
(i, j) (9
D
)
n1
(r n + 1)

and
[r]1

n=1
(2t)
n
n!

n
(i, j) c
2
e
v2t
1
[(1 )r]

, (10)
with v
2
= 2
2
9
D
and c
2
= (9
D
)
1
. Combining
Eqs. (6,7,10) we arrive at our bound in Eq. (1).
A key feature of our combined bound is that both ve-
locities v
1
and v
2
actually decrease (through the implicit
-dependence of ) with shorter interaction range (larger
), consistent with the expected physical picture. Note
that the parameter can be optimized to give the best
possible bound for a particular range of interactions. For
small , 1 recovers the Hastings-Koma bound at
large r (since the long-range part of the bound dominates
at suciently large r for any ,= 0), whereas for large
, one can choose closer to unity in order to improve
the short-range part of the bound.
Applications to experimentally realizable models We
now show that the coexistence of behavior consistent
with both terms in Eq. (1) can be seen in experimen-
tally realizable lattice spin models. We consider a 1D
1/r

interacting spin-1/2 chain, with (a) XY interac-


tions: H
XY
=
1
4

i=j
(
x
i

x
j
+
y
i

y
j
)/r

ij
, and (b) Ising
interactions with a transverse eld (TFIM): H
TFIM
=
1
2

i=j

x
i

x
j
/r

ij
+ B
z

z
i
. Ions in a linear rf-Paul
trap have already been used to simulate H
TFIM
with
variable B
z
and (0, 3) [22, 24, 25], and in the limit
B
z
1 (where the total spin excitation fraction is con-
served) can also simulate H
XY
[24, 25, 34]. Alternately,
for = 3, both H
XY
and H
TFIM
can be simulated with
polar molecules [13, 30, 35, 36]. In both models, we take
a spin-polarized initial state [ =

i
[
z
i
= 1 and
apply a local quench operator U
0
=
1

2
(1 +
x
0
) on the
0
th
site. We then numerically calculate the measurable
quantity Q
r
(t) [[U

x
r
(t)U
0
[ [
x
r
(t)[[ at a
specic time t. For a chain with N spins, we choose the
time 1 t N small enough to avoid boundary eects
and large enough to prohibit a perturbative treatment of
the time evolution.
For the long-range XY model subjected to the stated
local quench, we can restrict our attention to the sin-
gle spin-excitation subspace during the entire time evo-
lution. As a result, we can map the spin model to
a solitary free particle, making numerical calculation
trivial for hundreds of spins. It is also easy to show
that Q
r
(t) = |[
x
r
(t), U
0
]|, and thus the commutator
norm in Eq. (1) is measurable. For N = 501 spins and
= 2, 3, 6, , Fig. 4(a) demonstrates that at a specic
time, the distance dependence of Q
r
(t) can be divided
into several regions: (I) 1 r r
LC
v
max
t, where v
max
denotes the maximum group velocity of the free particle.
Q
r
(t) increases to its maximum value at r r
LC
. (II)
r
LC
< r < r
c
, where Q
r
(t) decays faster than a power
law. Note that for = 3 and = 6, Q
r
(t) is almost
unchanged by the addition of long-ranged interactions
for r < r
c
. Thus the behavior of Q
r
(t) in this region is
4
10
0
10
1
r
(b)


!=2
~1/r
2
!=3
~1/r
3
!=6
~1/r
6
!="
10
0
10
2
10
!5
10
0
r
Q
r
(
t
)
(a)


!=2
t/r
2
!=3
t/r
3
!=6
t/r
6
!="
10
1
10
0
10
5
10
2
10
0
10
0
(a) (b)
=6
=3
=2
=
t/r
6
t/r
3
t/r
2
r r
Q
r
(
t
)
FIG. 4: (Color online). (a) Qr(t) from a local quench of 1/r

XY chain (periodic boundary conditions, N = 501, t = 5).


(b) Qr(t) from a local quench of 1/r

TFIM chain (open


boundary conditions, N = 23, Bz = 0.5, t = 1).
a direct consequence of nearest-neighbor interactions in
the system, and is captured by the rst term in Eq. (1).
(III) r > r
c
, where Q
r
(t) decays algebraically as 1/r

due to the second term in Eq. (1). Note, however, that


Q
r
(t) t/r

(which is asymptotically exact in the limit


of t/r 0 [37]) does not saturate the time dependence
exp(v
2
t)1 in Eq. (1). This exponential time dependence
in our bound (as well as in Ref. [8]) results from the
n
in Eq. (2) adding in phase. For the XY model, a more
careful analysis shows that the contributions for dierent
n do not add constructively, causing Q
r
(t) to depend lin-
early on t [37]. These issues not-withstanding, it is abun-
dantly clear (especially in the = 3 and = 6 cases)
that the distance dependence of Q
r
(t) is a combination
of a nearest-neighbor-interaction contribution (leading to
rapid decay outside of a well-dened light cone) and a
long-range-interaction contribution scaling as 1/r

.
In the TFIM, the long-ranged interactions prevent a
mapping onto a free model, and therefore our numeri-
cal calculation is limited to a relatively small chain size
(N = 23). Setting B
z
= 0.5, which accentuates the
role of quantum uctuations, we calculate Q
r
(t) numer-
ically for = 2, 3, 6, using a Krylov-subspace projec-
tion method. Figure 4(b) shows that a local quench of
the TFIM yields qualitatively similar behaviors to the
XY model. For large r, we see a clear power-law decay
1/r

. For intermediate r, we see hints of faster than


power-law decay similar to the nearest-neighbor case.
Bound on the propagation of correlations.We now
use Eq. (1) to bound the spread of correlations follow-
ing a global quench [24, 27]. Specically, we consider
connected correlation functions C(t) = A(t)B(t)
A(t)B(t), evaluated for initial product states, with
A and B supported on sites i and j, respectively.
We rst dene an operator

A(t) = e
iHit
Ae
iHit
evolv-
ing under H
i

1
2

r
ik
,r
il
<r
h
kl
, with r r
ij
/2 (simi-
larly for

B). Because the supports of

A(t) and

B(t) are
nonoverlapping, the operators A(t)

A(t) and B(t)

B(t)
determine the correlation function C(t), leading to
[C(t)[ 2(|A(t)

A(t)||B|+|B(t)

B(t)||A|). (11)
Because H
i
agrees with H in the neighborhood of i, we
expect

A(t) A(t) to be small; indeed, it is bounded as
|A(t)

A(t)|

t
0
d

r
ik
<r,r
il
r
|[

A(), h
kl
]|, (12)
where h
kl
are the terms in H that expand the support of

A(t). Applying Eq. (1) and carrying out the summations,


we arrive at the following bound for correlation functions
[C(t)[ 2|A||B|

c
3
e
v1t
1
e
r/2
+c
4
e
v2t
1
[(1 )r/2]
D

,
(13)
where the constants c
3
and c
4
are nite for all > D
[33]. We see that, interestingly, while local-quench and
global-quench dynamics obey the same form of the short-
range part of the bound, the long-range part has dier-
ent r dependence for the two types of quenches (1/r

vs. 1/r
D
). The intuition behind the appearance of
1/r
D
is in the summation over l in Eq. (12), which can
be thought of as an eective integration of 1/r

il
over the
D-dimensional cubic lattice outside the support of

A(t).
Outlook. In addition to being relevant to a variety
of equilibrium [2, 69] and short-time non-equilibrium
[35] phenomena, we also expect the derived bound to
shed light on long-time relaxation processes in quantum
many-body systems [4]. It would be very interesting to
try to either saturate or tighten the time dependence
in the long-range part of the bound, thereby proving or
ruling out the possibility of quantum state transfer [38]
in time t log r. Similarly, whether it is possible to
saturate or tighten the 1/r
D
dependence for the spread
of correlations after a global quench is an interesting open
question deserving of further study.
We thank J. Preskill for asking whether the bound de-
rived in Ref. [8] reduces to the nearest-neighbor case as
. We thank A. M. Rey, K. Hazzard, C. Monroe,
L.-M. Duan, C. Senko, P. Richerme, M. Maghrebi, A. Da-
ley, J. Schachenmayer, A. Lee, J. Smith, M. Kastner, and
S. Manmana for discussions. This work was supported
by the JQI and the PFC at JQI. MF thanks the NRC
for support. SM acknowledges funding provided by the
Institute for Quantum Information and Matter, an NSF
Physics Frontiers Center with support of the Gordon and
Betty Moore Foundation through grant GBMF1250.
[1] E. H. Lieb and D. W. Robinson, Comm. Math. Phys. 28,
251 (1972).
[2] B. Nachtergaele and R. Sims, arXiv:1004.2086 (2010).
[3] T. Fukuhara, P. Schausz, M. Endres, S. Hild, M. Cheneau,
I. Bloch, and C. Gross, Nature (London) 502, 76 (2013).
5
[4] T. Langen, R. Geiger, M. Kuhnert, B. Rauer, and
J. Schmiedmayer, Nature Phys. 9, 640 (2013).
[5] A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalat-
tore, Rev. Mod. Phys. 83, 863 (2011).
[6] S. Michalakis and J. Pytel, Comm. Math. Phys. 322, 277
(2011).
[7] S. Michalakis, arXiv:1206.6900 (2012).
[8] M. Hastings and T. Koma, Comm. Math. Phys. 265, 781
(2006).
[9] M. B. Hastings, Phys. Rev. B 69, 104431 (2004).
[10] P. Hauke and L. Tagliacozzo, Phys. Rev. Lett. 111
(2013).
[11] J. Eisert, M. van den Worm, S. R. Manmana, and
M. Kastner, Phys. Rev. Lett. 111, 260401 (2013).
[12] M. Saman, T. G. Walker, and K. Mlmer, Rev. Mod.
Phys. 82, 2313 (2010).
[13] B. Yan, S. A. Moses, B. Gadway, J. P. Covey, K. R. A.
Hazzard, A. M. Rey, D. S. Jin, and J. Ye, Nature (London)
501, 521 (2013).
[14] K. Aikawa, A. Frisch, M. Mark, S. Baier, A. Rietzler,
R. Grimm, and F. Ferlaino, Phys. Rev. Lett. 108, 210401
(2012).
[15] M. Lu, N. Q. Burdick, and B. L. Lev, Phys. Rev. Lett.
108, 215301 (2012).
[16] L. Childress, M. V. Gurudev Dutt, J. M. Taylor, A. S. Zi-
brov, F. Jelezko, J. Wrachtrup, P. R. Hemmer, and M. D.
Lukin, Science 314, 281 (2006).
[17] G. Balasubramanian, P. Neumann, D. Twitchen,
M. Markham, R. Kolesov, N. Mizuochi, J. Isoya,
J. Achard, J. Beck, J. Tissler, et al., Nature Mater. 8,
383 (2009).
[18] J. R. Weber, W. F. Koehl, J. B. Varley, A. Janotti, B. B.
Buckley, C. G. Van de Walle, and D. D. Awschalom, Proc.
Natl. Acad. Sci. 107, 8513 (2010).
[19] F. Dolde, I. Jakobi, B. Naydenov, N. Zhao, S. Pezzagna,
C. Trautmann, J. Meijer, P. Neumann, F. Jelezko, and
J. Wrachtrup, Nature Phys. 9, 139 (2013).
[20] P. Schau, M. Cheneau, M. Endres, T. Fukuhara, S. Hild,
A. Omran, T. Pohl, C. Gross, S. Kuhr, and I. Bloch, Na-
ture (London) 491, 87 (2012).
[21] O. Firstenberg, T. Peyronel, Q.-Y. Liang, A. V. Gor-
shkov, M. D. Lukin, and V. Vuletic, Nature (London) 502,
71 (2013).
[22] R. Islam, C. Senko, W. C. Campbell, S. Korenblit,
J. Smith, A. Lee, E. E. Edwards, C.-C. J. Wang, J. K.
Freericks, and C. Monroe, Science 340, 583 (2013).
[23] J. W. Britton, B. C. Sawyer, A. C. Keith, C. C. J. Wang,
J. K. Freericks, H. Uys, M. J. Biercuk, and J. J. Bollinger,
Nature (London) 484, 489 (2012).
[24] P. Richerme, Z.-X. Gong, A. Lee, C. Senko, J. Smith,
M. Foss-Feig, S. Michalakis, A. V. Gorshkov, and C. Mon-
roe, arXiv:1401.5088 (2014).
[25] P. Jurcevic, B. P. Lanyon, P. Hauke, C. Hempel, P. Zoller,
R. Blatt, and C. F. Roos, arXiv:1401.5387 (2014).
[26] S. Gopalakrishnan, B. L. Lev, and P. M. Goldbart, Phys.
Rev. Lett. 107, 277201 (2011).
[27] J. Schachenmayer, B. P. Lanyon, C. F. Roos, and A. J.
Daley, Phys. Rev. X 3, 031015 (2013).
[28] M. Knap, A. Kantian, T. Giamarchi, I. Bloch, M. D.
Lukin, and E. Demler, Phys. Rev. Lett. 111, 147205
(2013).
[29] M. Foss-Feig, K. R. A. Hazzard, J. J. Bollinger, and A. M.
Rey, Phys. Rev. A 87, 042101 (2013).
[30] K. R. A. Hazzard, S. R. Manmana, M. Foss-Feig, and
A. M. Rey, Phys. Rev. Lett. 110, 075301 (2013).
[31] J. Juenemann, A. Cadarso, D. Perez-Garcia,
A. Bermudez, and J. J. Garcia-Ripoll, Phys. Rev.
Lett. 111, 230404 (2013).
[32] S. Bachmann, S. Michalakis, B. Nachtergaele, and
R. Sims, Comm. Math. Phys. 309, 835 (2012).
[33] See Supplemental Material at link X for details.
[34] Z.-X. Gong and L.-M. Duan, New Journal of Physics 15,
113051 (2013).
[35] A. V. Gorshkov, S. R. Manmana, G. Chen, E. Demler,
M. D. Lukin, and A. M. Rey, Phys. Rev. A 84, 033619
(2011).
[36] A. V. Gorshkov, S. R. Manmana, G. Chen, J. Ye,
E. Demler, M. D. Lukin, and A. M. Rey, Phys. Rev. Lett.
107, 115301 (2011).
[37] M. Foss-Feig, Z.-X. Gong, and A. V. Gorshkov, (in prepa-
ration) (2014).
[38] S. Bose, Contemp. Phys. 48, 13 (2007).
[39] We note that generalizations to fermionic/bosonic models
and arbitrary lattice geometris are possible [8]. Moreover,
for nite lattices, the bounds we derive can be tightened
and extended to < D.
[40] This can be easily proven by projecting the entire path
of hoppings onto the line connecting sites i and j.
Supplementary Material for: Persistence of locality in systems with power-law
interactions
S1. PROOF OF EQ. (9) IN THE MAIN TEXT
Consider rst the case when i ,= j, then for any > D 1,

k
J
ik
J
kj

r
ik
<2
J
ik
J
kj
+

r
jk
<2
J
ik
J
kj
+

r
ik
2
r
jk
2
r

ik
r

kj
(S1)
2

r
ik
<2
J
ik
J
kj
+ 2

ij

r
ik
2
r

ik
, (S2)
6
where the second inequality uses r

ij
(r
ik
+r
kj
)

[(2r
ik
)

+(2r
kj
)

]/2. The notation

r
ab
<c
indicates a summation
over all b for which r
ab
< c. Now let x

= 2y

+w

be the Cartesian coordinates of the vector r


ik
, where w

= 0, 1 for
y

1, w

= 0, 1 for y

= 0, and w

= 0, 1 for y

1. The constraint r
ik
2 is then equivalent to

D
=1
y
2

,= 0
for D = 1, 2, 3. We can now write

r
ik
2
r

ik
=

y
2

=0

[
D

=1
(2y

+w

)
2
]
/2
(S3)

y
2

=0
3
D
[
D

=1
(2y

)
2
]
/2
(S4)
= 3
D
2

( 1). (S5)
As a result,

k
J
ik
J
kj
2

r
ik
<2
J
ik
J
kj
+J
ij
3
D
( 1) (S6)
3
D

r
ik
<2
J
ik
J
kj
+ 3
D
( 1)

r
ik
<2
J
ik
J
kj
(S7)
= 3
D

r
ik
<2
J
ik
J
kj
. (S8)
Consider now the situation when i = j, in which case we have

k
J
2
ik

k
J
ik
< 3
D

r
ik
<2
J
2
ik
. Therefore, Eq. (9)
of the main text holds for any i and j.
S2. PROOF OF EQS. (12-14) IN THE MAIN TEXT
For an initial product state, we have

A(t)

B(t) =

A(t)

B(t). Eq. (12) in the main text then follows from


[C(t)[ = [[A(t)

A(t)]B(t) +

A(t)[B(t)

B(t)] A(t)

A(t)

B(t) A(t)B(t)

B(t)[, (S9)
and Eq. (13) follows immediately from the inequality
|A(t)

A(t)| = |

t
0
d
d
d
(e
iH
e
iHi
Ae
iHi
e
iH
)|

t
0
d|[

A(), H H
i
]|. (S10)
To prove Eq. (14), we rst note the bound given by Eq. (1) in the main text can be generalized to operators
supported on more than one site by using the shortest distance between the supports of the two operators, and by
multiplying the bound by sizes (the number of sites) of the two supports [S1]. Thus, using our bound Eq. (1) in the
main text, for i ,= k we obtain
|[

A(), h
kl
]| 4|A||h
kl
|

c
1
e
v1
e
r
ik
+c
2
e
v2
[(1 )r
ik
]

. (S11)
The case of i = k can be bounded by the rst term of Eq. (S11) if we replace c
1
= 1/ with 1. Therefore,

r
ik
<r
r
il
r
|[A(), h
kl
]|
4 |A|

r
ik
<r
r
il
r
r

kl
e
v1
e
r
ik
+

0<r
ik
<r
r
il
r
r

kl
c
2
e
v2
[(1 )r
ik
]

. (S12)
The second sum above can be easily bounded using Eq. (4) in the main text, giving

0<r
ik
<r
r
il
r
r

kl
c
2
e
v2
[(1 )r
ik
]


c
2
e
v2
(1 )

r
il
r
r

il
(S13)
[
2
9(1 )
]
D

1
e
v2
[(1 )r/2]
D
, (S14)
7
where we have used the inequality

r
il
r
r

il

1
/r
D
for some D-dependent constant
1
. To bound the rst
sum in Eq. (S12), we note that for r
ik

1+
2
r,

r
ik

1+
2
r
r
il
r
r

kl
e
r
ik

r
il
r
(r
il

1 +
2
r)

k
e
r
ik
(S15)


2
[(1 )r/2]
D

1
, (S16)
where we have used the fact that

r
il
r
(r
il

1+
2
r)


2
/(
1
2
r)
D
for some constant
2
, and have dened

k
e
r
ik
. For r
ik
>
1+
2
r, the inequality

r
ik
>
1+
2
r
e
r
ik

2
e
r/2
holds for some constant
2
, and thus

1+
2
r<r
ik
<r
r
il
r
r

kl
e
r
ik

r
ik
>
1+
2
r
e
r
ik

r
kl
>0
r
kl

(S17)

2
e
r/2
. (S18)
Combining Eqs. (S12,S14,S16,S18) and noting that v
2
> v
1
, we obtain

r
ik
<r
r
il
r
|[A(), h
kl
]|
4 |A|

2
e
v1r/2
+ (
2

1
+ [
2
9(1 )
]
D

1
)
e
v2
[(1 )r/2]
D
. (S19)
As the above inequality has exactly the same form when replacing the operator A by the operator B, one can readily
obtain Eq. (13) in the main text by dening
c
3
8
2
/v
1
and c
4
8(
2

1
+
1
[
2
9(1 )
]
D
)/v
2
. (S20)
Note that
1
,
2
are D dependent constants and
1
,
2
both depend on and D. For D = 1, they have the simple
explicit forms

1
=
2
= and
1
=
2
= coth(/2). (S21)
[S1] M. Hastings and T. Koma, Comm. Math. Phys. 265, 781 (2006).

Anda mungkin juga menyukai