You are on page 1of 11

Topology by Dissipation: Transport properties

Gal Shavit1 and Moshe Goldstein1


1
Raymond and Beverly Sackler School of Physics and Astronomy, Tel Aviv University, Tel Aviv 6997801, Israel
Topological phases of matter are the center of much current interest, with promising potential
applications in, e.g., topologically-protected transport and quantum computing. Traditionally such
states are prepared by tuning the system Hamiltonian while coupling it to a generic bath at very low
temperatures; This approach is often ineffective, especially in cold-atom systems. It was recently
shown that topological phases can emerge much more efficiently even in the absence of a Hamiltonian,
by properly engineering the interaction of the system with its environment, to directly drive the
arXiv:1903.05336v1 [cond-mat.quant-gas] 13 Mar 2019

system into the desired state. Here we concentrate on dissipatively-induced 2D Chern insulator
(lattice quantum Hall) states. We employ open quantum systems tools to explore their transport
properties, such as persistent currents and the conductance in the steady state, in the presence of
various Hamiltonians. We find that, in contrast with equilibrium systems, the usual relation between
the Chern topological number and the Hall conductance is broken. We explore the intriguing edge
behaviors and elucidate under which conditions the Hall conductance is quantized.

I. INTRODUCTION details of the engineered bath coupling.


Topologically non-trivial phases of matter [13] have
The adverse affects of a dissipative environment or been at the forefront of condensed matter physics during
bath on a quantum system coupled to it are well-known: the last several decades, since the discovery of the integer
Processes of decay and decoherence lead to a degrada- and fractional quantum Hall effects [14, 15]. The more
tion of the desired quantum state and shortening of its recently discovered topological insulators [16–18], super-
coherence time [1, 2]. For these reasons, a prime goal of conductors [19, 20], and semimetals [21] have opened the
quantum device engineering is to mitigate those effects, door to novel exciting possibilities, e.g., topologically pro-
mainly by trying to reduce the system-bath interaction tected quantum computing [22]. The idea of employing
strength. However, it has been realized that an engi- the engineered bath interactions scheme to stabilize a
neered dissipative interaction to an external bath can topologically ordered ground state has already received
possibly be exploited, in order to reliably prepare inter- some attention in recent years [9, 23–31].
esting non-trivial quantum states in open systems; This We focus on the 2D Chern insulator phase discussed
approach has recently been brought to many-body sys- in Ref. [31], stabilized by employing purely dissipative
tems [3–9]. This is typically done by tailoring the bath dynamics of Lindblad type, i.e., the master equation de-
interaction such that the required quantum state would scribing the evolution of the system density matrix ρ,
emerge as a unique steady state solution to the open
d
quantum (Lindblad [10]) master equation, independent ρ = −i [H, ρ] + Dρ, (2)
of the initial conditions. If the sought after state is rep- dt
resented by the density matrix ρD , one would aim to has the Hamiltonian H = 0, and where D is the engi-
manipulate the system-bath coupling such that ρD is a neered dissipator super-operator acting on ρ. This work
“dark state” of the dynamical evolution of the open sys- addresses the effects of including H 6= 0 in the Lindblad
tem, i.e., dynamics (2). This step is crucial, as it allows one to
define current operators in the system, such that trans-
d
ρD = 0 (1) port properties of the dissipative state may be explored.
dt
These properties, e.g, the Hall conductance, are the hall-
This idea of dissipative preparation contrasts the con- mark of the well-known equilibrium counterpart of this
ventional Hamiltonian approach, that heavily relies on phase, and as such are important in characterizing the
reaching sufficiently low temperatures to attain the in- engineered state. We will show that although the dissipa-
teresting properties of the desired quantum ground state. tively engineered states could be arbitrarily close to equi-
This approach might prove ineffective in some cases. Of librium quantum Hall states, they do not always present
particular interest is the case of quantum simulators im- the same transport features. These depend on the de-
plemented using ultra-cold atomic gases [11, 12], where tails of the coherent dynamics, the relative amplitude of
equilibrium at low temperatures compared to the trap- those dynamics compared to the dissipation energy scale,
ping potential energy scale is experimentally challeng- and on the implementation of artificial fields within the
ing to achieve. A dissipative preparation scheme, albeit dissipative scheme.
with its own challenges and complexities, may circum- The rest of the paper is organized as follows. In Sec. II
vent such issues, as the accuracy of the final state will be we briefly introduce the dissipative scheme developed in
determined solely by the degree of control one has on the [31]. We introduce the Hamiltonian dynamics, and the
2

flux per plaquette in units of flux quantum, introduced


using the Peierels substitution. α is assumed to be a ra-
tional fraction α = pq (with p, q, integers with no common
factor). The Hamiltonian (3) is diagonalized by moving
to two-dimensional momentum space,

k,λ a†k,λ ak,λ ,


X
H ref = (4)
k,λ

with k the two-dimensional momentum, and λ =


1, 2, ..., q the band index. The spectrum obtained for
α = 71 with periodic boundary conditions along one di-
rection is shown in Fig. 2(a), where the distinct bands
are apparent, as well as the edge state spectrum.
The goal of the construction of [31] is to achieve a state
as close as possible to the ground state of the reference
Hamiltonian with chemical potential in the first gap (i.e.,
the lowest band completely filled and all the other bands
FIG. 1. Schematic depiction of the dissipative scheme de- completely empty) by purely dissipative dynamics. For
scribed in the text. (a) Orange particles, representing the that we introduce a different Hamiltonian, built using the
a particles, may hop to a nearest neighbor site and become matrix elements of the Hofstadter reference Hamiltonian
a b particle (blue). This hopping has a phase which is de- (3),
termined by the artificial magnetic field implemented in the
lattice, according to Eq. (5). (b) The b particles are not
trapped, and escape the lattice, which imposes a band and
X
H diss = −tdiss b†m,n e±ilx 2πα alx ,ly ±1 + alx ±1,ly

momentum dependent effective depletion rate for the a parti-
out lx ,ly
cles, γk,λ . An external refilling reservoir for a is added, with
b†lx ,ly alx ,ly + h.c. ,
X
a tuned wavevector- and band-independent refilling rate γ in . ∗
−µ (5)
lx ,ly

tools required to analyze the steady state in its presence describing a coupling of the lattice fermions a to an
in Sec. III. This step enables us to discuss the persistent additional fermionic species b (e.g., a different hyper-
steady state currents that develop for different classes of fine state). A discussion regarding the cold-atom im-
Hamiltonians in Sec. IV. Then, the electric conductance plementation for this Hamiltonian is presented in [31],
response is calculated in Sec. V for the different cases, and is based on previous constructions suggested in Refs.
and compared with the known results for the equilibrium [34, 35]. In the eigenbasis of the reference Hamiltonian
scenario. Finally, we summarize our findings and conclu- we obtain
sions in Sec. VI.
(k,λ − µ∗ ) b†k,λ ak,λ + h.c. ,
X
H diss = (6)
k,λ
II. DISSIPATIVELY INDUCED TOPOLOGICAL
STATE with k,λ now proportional to tdiss instead of tref . We as-
sume that, in the cold-atom implementation, the b parti-
In this section, we recapitulate the dissipative recipe cles are not trapped by the confining potential in the
presented in Ref. [31] to realize a dissipative lattice in- direction perpendicular to the 2D optical lattice, and
teger quantum Hall state. The main components in the escape to infinity. Treating them as a “bath” for the
scheme are illustrated in Fig. 1. Consider the Harper- trapped lattice particles, we can integrate them out, and
Hofstadter model [32, 33], describing nearest-neighbor arrive at a contribution to the Lindblad master equation
hopping on a two-dimensional square lattice pierced by for the density matrix of the a particles,
a magnetic flux, which we will use as a reference in the  o
1n †
construction below, ak,λ ρa†k,λ −
X
Dout ρ = out
γk,λ ak,λ ak,λ , ρ . (7)
2
X † k,λ
H ref = −tref alx ,ly eilx 2πα alx ,ly +1 + alx +1,ly + h.c. ,


lx ,ly out
Calculation of the rate γk,λ can be done using Fermi’s
(3) golden rule [36], to find
with alx ,ly a fermionic annihilation operator on the site 2π 2
(lx , ly ), tref the hopping strength, and α is the magnetic
out
γk,λ = ν0 |k,λ − µ∗ | , (8)
~
3

with ν0 , the density of available b-states, taken constant


[31]. We thus define a typical dissipative scale for the
system,

2π 2
γ0 ≡ ν0 |tdiss | . (9)
~

The implication of (8) is that given a flat band, i.e.,


k,λ = 0 for one value of λ, it is possible to fine-tune
µ∗ such that the depletion rate goes to zero for that par-
ticular band alone. The system will exponentially decay
into a state where the finely-tuned flat band is the only FIG. 2. Example of the purely-dissipative scheme. (a)
occupied band, with its occupancy depending on the ini- Calculated spectrum k,λ of the reference Hamiltonian. The
tial state of the lattice filling. This is analogous to a low different bands are visibly shown, with the edge state spec-
temperature equilibrium scenario where the chemical po- trum appearing in the interband gaps. (b) Using the reference
tential lies in a gap between bands. Hamiltonian to construct H diss , the steady state occupation
Since a topologically non-trivial (non-zero Chern num- numbers are obtained. The bottom band is almost entirely
populated, whereasPall the rest are nearly depleted. We set
ber) exactly flat-band in a finite-range hopping Hamilto-
µ∗ = k,1 ≡ Lx1Ly k k,1 , and γopt
in
was calculated according
nian, separated by a finite gap from the other bands, is
to Eq. (12). In both plots we used α = 17 , Lx = Ly = 301,
not possible [37, 38], we consider a band which is nearly
periodic boundary conditions along the y direction, and open
flat, i.e., with a width much smaller compared to its mini- boundary conditions in the x direction (cylindrical geometry).
mal distance to the other bands. Then, a very small value
out
of γk,λ for that particular band is attainable. We now in-
troduce another ingredient to the dissipative scheme, a III. INTRODUCING HAMILTONIAN
global filling reservoir, replenishing all bands at a rate DYNAMICS
γ in . The full master equation now reads
 o We now consider adding some coherent dynamics to
d 1n †
ak,λ ρa†k,λ −
X
out the a particles which live on the lattice. This will modify
ρ= γk,λ ak,λ ak,λ , ρ
dt 2 the master equation (2), since if H 6= 0, one must include
k,λ
X †
1n o the commutator term between the Hamiltonian and the
+ γ in ak,λ ρak,λ − ak,λ a†k,λ , ρ . (10) density matrix. In this work, we explore the consequences
2
k,λ of two kinds of hopping Hamiltonians,

The
n steady-state band occupation numbers, nk,λ =
X †
alx ,ly eilx 2πα alx ,ly +1 + alx +1,ly + h.c. ,

o Hcomp = −tcomp

Tr ρak,λ ak,λ can now be calculated, lx ,ly
(13)
in
γ
nk,λ = out . (11)
γ in + γk,λ
a†lx ,ly alx ,ly +1 + alx +1,ly + h.c. , (14)
X 
Hinc = −tinc
lx ,ly
Given the values
n oof the maximal rate for the bottom
out
band, max γk,1 , and the minimal one for the next
n o where Hcomp is compatible with the dissipative interac-
out tion Hamiltonian, i.e., it has the dynamics of particles
(closest) band min γk,2 , an optimal choice of the tun-
able refilling rate, such that nk,1 ∼ 1 and nk,λ>1 ∼ 0 is hopping on a square lattice with the same flux α as
thus before, and Hinc is a simple nearest-neighbor hopping
r Hamiltonian which is incompatible.
in
n o n
out · min γ out .
o First, we want to understand how the presence of either
γopt = max γk,1 k,2 (12) (13) or (14) affects the steady state we arrived at using
the engineered dissipative scheme. We notice right away
Fig. 2(b) shows an example of results for this scheme, that the steady state solution for ρ with H = 0 commutes
where we have properly tuned µ∗ and γ in , and achieved with Hcomp , since it is diagonal in the same k, λ basis as
almost completely full or empty occupation of the lower γ out . As a result, Eq. (11) holds exactly even for a non-
band or upper bands, respectively. This mixed state is zero compatible Hamiltonian H = Hcomp .
not only very close to the pure ground state of the refer- Conversely, if H = Hinc the steady state can be much
ence Hamiltonian, but also shares with it the value -1 of different than (11), depending on the ratio tγinc
0 . We use

the topological index, the Chern number [31]. the full master equation to calculate the steady state sin-
4
n o
Tr ρ (t) a†lx ,ly alx ,ly (see Appendix B for the full deriva-
tion). Using the master equation (2) for the evolution
of the density matrix, one can separate the local change
in particle number to a coherent contribution ṅH and a
dissipative part ṅD ,
d D
nl ,l (t) = ṅH
lx ,ly (t) + ṅlx ,ly (t) . (17)
dt x y
The coherent part, controlled by the [H, ρ] term in the
master equation, must obey a continuity equation involv-
ing the particle current, ṅH + ∇ · j = 0, where (∇·) is
a lattice version of the divergence operator, or more ex-
plicitly,
   
y y
ṅH x x
lx ,ly (t) = − jlx +1,ly − jlx ,ly − jlx ,ly +1 − jlx ,ly .
(18)
FIG. 3. Calculation of occupation numbers with H = Hinc .
Different values of tinc and γ in were used in each plot. (a)
By examining the expression for ṅH , which depends on
tinc = 0.01γ 0 , γ in = 1.5γopt
in
, (b) tinc = 0.03γ 0 , γ in = 10γopt
in
, the details of the Hamiltonian, one can extract and define
0 in in
(c) tinc = 0.1γ , γ = 80γopt , and (d) tinc = 0.5γ , γ = 0 in a proper current operator. As shown in Appendix B, the
in
400γopt . For all plots we used α = 71 , Lx = Ly = 140, and steady state expectation values of the current are fully
the same cylindrical geometry as in Fig. 2. given in terms of elements of P , so no further complicated
calculations are required in order to obtain them. Since
d
in the steady state dt n = 0, the dissipative steady state
gle particle density matrix, whose eigenstates are the oc- contribution can be calculated from the divergence of the
cupation numbers nk,λ , current, ṅD = −ṅH = ∇ · j, and can be decomposed into
n o incoming and outgoing terms,
Plx ,ly ;lx0 ,ly0 = Tr ρ (t → ∞) a†lx ,ly alx0 ,ly0 . (15)
D,in D,out
ṅD
lx ,ly ≡ Jlx ,ly − Jlx ,ly , (19)
This requires us to solve a Sylvester-type matrix equation
(see Appendix A for details) which are proportional to the single particle matrix el-
    ements of γ in and γ out , respectively (see Appendix B).
1 in out 1 in out We present our results below for the different classes of
− ih = γ in ,
 
γ +γ + ih P +P γ +γ
2 2 Hamiltonian we have considered.
(16)
with h, γ in , and γ out the single-particle matrices corre-
sponding to H, γ in , and γ out , respectively. We find that A. Compatible hopping
the engineered nearly-pure steady state deteriorates with
increasing tinc . As shown in Fig. 3, the incompatibility The expectation value of the electric current operator
of the Hamiltonian requires one to use a faster refilling in the different directions is given in terms of elements of
rate γ in to maintain the high occupation of the bottom the P matrix by (Appendix B)
band. This in turn leads to non-negligible occupation
jlxx ,ly = itcomp Plx +1,ly ;lx ,ly − Plx ,ly ;lx +1,ly ,

of the higher energy levels, deviating from the desired (20a)
state. This phenomenon becomes significant at about
tinc ∼ 0.1γ 0 , i.e., when the Hamiltonian is no longer neg- jlyx ,ly = itcomp eilx 2πα Plx ,ly +1;lx ,ly − e−ilx 2πα Plx ,ly ;lx ,ly +1 .

ligible compared to the dissipative energy scale. We note
(20b)
that the gap in the spectrum of P (purity gap) is still
Since P itself is not affected by the presence of the Hamil-
finite, and the associated Chern number retains its value jlµ ,l
x y
of -1 [31]. tonian, the normalized current tcomp is completely in-
dependent of tcomp . As shown in Fig. 4, we find chi-
ral currents strongly localized near the edges, much like
IV. PERSISTENT CURRENTS in the ground state of the equilibrium quantum Hall ef-
fect. We note that the current in this compatible case is
Including an Hamiltonian for the a species finally divergence-free, as expected for a chiral edge mode. As
allows us to define a sensible current operator in a consequence, the dissipative current ṅD ∝ ∇ · j is zero,
the system, by looking at the time evolution of the signaling a zero local net flux of particles into or out of
local particle density expectation value nlx ,ly (t) ≡ the baths.
5

FIG. 5. (a) The current distribution calculated in a square


geometry and the Landau gauge choice used in (5). Each
arrow is proportional to the current density vector at the
corresponding lattice site. (b) The normalized dissipative
current ṅD , indicating the local exchange of particles with
the reservoir. Parameters used are tinc = 10−3 γ 0 , α = 17 ,
FIG. 4. The current distribution calculated in a square
Lx = Ly = 70, and the optimal values for µ∗ and γ in .
geometry. Each arrow is proportional to the current density
vector at the corresponding lattice site. We used α = 71 , Lx =
Ly = 70, and the optimal values for µ∗ and γ in . Localization
of a chiral edge mode is visible. Inset: calculation with a
cylindricalPgeometry (periodic boundaries along the y axis)
of jlyx = y
ly jlx ,ly , using the same parameters, except for
Lx = Ly = 140.

B. Incompatible hopping

In the case of the hopping Hamiltonian (14), the cur-


rent is similarly given by

jlxx ,ly = itinc Plx +1,ly ;lx ,ly − Plx ,ly ;lx +1,ly ,

(21a)
FIG. 6. (a) The current distribution calculated in a square
geometry and the symmetric gauge A ~ = (−By/2, Bx/2, 0).
jlyx ,ly

= itinc Plx ,ly +1;lx ,ly − Plx ,ly ;lx ,ly +1 . (21b)
Each arrow is proportional to the current density vector at
We find that the steady state currents which develop in the corresponding lattice site. (b) The normalized dissipative
the system in such a scenario are completely different current ṅD , indicating the local exchange of particles with the
reservoir. All parameters are the same as in Fig. 5.
than those of the compatible Hamiltonian, as we show
in Fig. 5. The current is not localized near the edges,
but rather flows along the y direction with a structure
whose periodicity is determined by the value of α. The invariance: A gauge transformation would modify both
current trajectories terminate at the system edges where H diss and H, whereas here we incorporate the “magnetic
phase” using different choices of A ~ only in H diss . As an
they appear to backscatter through a net flux of a parti-
cles leaving and re-enetering the system at nearby lattice example, in Fig. 6 we make use of the symmetric gauge
choice A~ = (−By/2, Bx/2, 0). The current pattern which
sites. This manifests as a finite spatially oscillating dis-
sipative current ṅD at the edges. emerges now has circulating currents around plaquettes
This result seems to be somewhat peculiar at first, with the size q × q.
since the current in the system appears to prefer one Qualitatively, the current patterns do not change ap-
direction (y) over the other (x). This can only be ac- preciably as tinc increases beyond the tinc  γ 0 limit.
counted for by the phase factors appearing in the dis- What we observe is a steep decline in the amplitude of
j
sipative Hamiltonian, Eq. (5), which imply the choice the normalized current tinc as tinc becomes comparable
of gauge for the vector potential A ~ = (0, Bx, 0), with to the dissipative scale γ 0 . This is understood by the
B the magnetic field, which we refer to as the Landau depletion of current carriers in the system as the steady
gauge. Changing this choice of gauge affects the resul- state occupation deteriorates, and is compensated by suf-
tant currents. This is of course not a violation of gauge ficiently increasing the value of the refilling rate γ in .
6

V. CONDUCTANCE

One of the most remarkable properties of the quantum


Hall ground state is its exactly quantized transverse elec-
trical conductance. In order to reveal the response and
transport properties of the dissipatively prepared state,
we introduce a small perturbation to the Hamiltonian,
playing the role of voltage bias along the x direction,
applied on parts of the system which will serve as leads,
"L Lx
#
lead
V X X
a†lx ,ly alx ,ly − a†lx ,ly alx ,ly
X
δH = ,
2
ly lx =1 lx =Lx −Llead +1
(22)
with V the applied voltage difference, and Llead is the size
of the leads. The exact value of Llead was numerically
verified to have only a negligible effect on the conduc-
tance, as long as it is small enough compared to Lx , and
larger than the few-site width of the edge regions, which
can be inferred from Figs. 4–6. An alternative way to FIG. 7. (a) conductance calculations for the case of a com-
define the conductance, using a perturbation which de- patible Hopping Hamiltonian. Gxx (blue) and Gyx (red) are
scribes a constant electric field along the system is briefly t
displayed as a function of the ratio comp . The behavior in the
γ0
described in Appendix C, and produced very similar re- dissipative regime is indicated, whereas in the Hamiltonian
sults. regime the transverse conductance is quantized toGyx ≈ − eh ,
2

As we are only interested in the linear response regime, the value of the topological invariant, the Chern number.
we calculate the first-order change in the occupation (b-c) Representative transverse current patterns of jlyx =
P y 0
numbers matrix δP due to a finite value of V , which ly jlx ,ly along the systems, for (b) tcomp = 0.01γ and (c)
amounts to solving another matrix Sylvester equation, tcomp = 100γ 0 [the corresponding values are marked by arrows
 in in (a)]. Throughout this figure we used a cylindrical geometry
γ + γ out
 in
γ + γ out
 
with edges along the x̂ direction with Lx = Ly = 70, α = 17 ,
+ ih δP +δP − ih = i [P0 , δh] ,
2 2 Llead = 7, and the optimal values for µ∗ and γ in .
(23)
with P0 the solution of Eq. (16) without the voltage
leads, and δh is the single particle matrix correspond- currents flows near system edges in a similar fashion (Fig.
ing to δH. We calculate the modification of the val- 7b,c).
ues of the current due to this perturbation, δjlµx ,ly = t
In the large compγ 0 case, which we refer to as the Hamil-
jlµx ,ly (P → P + δP )−jlµx ,ly (P ), and find the conductance tonian regime, we observe quantum-Hall-like behavior,
by summing over all the lattice sites, namely the quantization of the transverse response in ac-
cordance with the topological Chern number associated
1 X µ with the steady-state density matrix, which is equal to
Gµν = δjlx ,ly . (24)
Vν the one associated with the lowest band of H diss [31],
lx ,ly
alongside a vanishing longitudinal response with increas-
Eq. (22) corresponds to a choice of Vx = V and Vy = 0. ing tcomp .
Similarly to the previous section, we discuss our findings In the opposite limit, the dissipative regime, both
separately for the two possible Hamiltonian classes. transverse and longitudinal responses are small, and in-
crease as tcomp becomes larger. While Gxx grows lin-
t
early with compγ 0 , the transverse conductance Gyx is pro-
A. Compatible hopping  2
tcomp
portional to γ 0 . To understand why the linear
Unlike our discussion regarding the persistent currents, contribution to Gyx vanishes, although the current it-
it is clear from Eq. (23) that even in the compatible self is proportional to tcomp , one must examine the h = 0
case, the relative amplitude of the hopping Hamiltonian, limit of Eq. (23). Performing a Fourier transform of
as compared to dissipation rate, may play an important all the fermionic operators Pin the system with respect to
role. We indeed find two regimes for the conductance in the y direction, alx ,ly ∼ ky eiky ly alx ,ky , one finds that
t t 
our system, corresponding to comp γ0  1 and compγ0  1, in this basis the matrices γ in + γ out , P0 , and δh are
see Fig. 7a. Interestingly, in both regimes the transverse purely real, which would make δP in this limit purely
7

FIG. 8. Parameter dependence of the transverse conductance


Gyx in the Hamiltonian regime of the compatible hopping sce-
nario. (a) µ∗ is changed and γ in = γoptin
is kept constant. (b)
in ∗ ∗
γ is changed and µ = µopt is kept constant. The dashed
light blue line marks the quantized value Gyx = −1. Through- FIG. 9. The conductance calculated in the case of an in-
out this figure we used a cylindrical geometry with edges along compatible hopping Hamiltonian. Gxx (blue) and Gyx (red)
the x direction with Lx = Ly = 70, α = 17 , Llead = 7, and are displayed as a function of the ratio tγinc
0 . The behavior

tcomp = 100γ 0 . at small tγinc


0 is indicated. We used a cylindrical geometry
with edges along the x direction with Lx = Ly = 70, α = 17 ,
Llead = 7, and the optimal values for µ∗ and γ in .
imaginary. Since
D E δP is hermitian, its diagonal elements

alx ,ky alx ,ky vanish. But j y is composed of exactly
these vanishing averages. Hence Gyx is zero to first order tized value for Gyx , see Fig. 9. Again, this should not
t be surprising, as we have already seen that a large tinc
in compγ0 .
We have also investigated how the quantized conduc- negates our ability to manipulate the particles into the
tance (in the Hamiltonian regime) changes when the pa- desired dissipatively engineered steady state, which pos-
rameters controlling the dissipative scheme, i.e., µ∗ and sesses some QHE-like features. We note that in the dis-
γ in , are modified, and do not assume their optimal val- sipative regime, the current distribution is also indistin-
ues. We find that a precise tuning of µ∗ is required for guishable from the behavior for the compatible Hamilto-
the nearly precise quantization (Fig. 8a), and that a nian, e.g., Fig 7b.
small deviation from the optimal value deteriorates Gyx Lastly, we find that the conductance matrix is
significantly. This is because µ∗ should be tuned to min- anisotropic, namely that Gxx 6= Gyy and Gyx 6= −Gxy
out in this incompatible dissipative regime, which is not the
imize γk,1 , such that the λ = 1 band would be nearly
filled in the steady state. Once µ∗ is detuned away from case for the compatible Hamiltonian. Once more, this is
the middle of the first band, this band will always be par- nothing but a consequence of the choice of Landau vector
tially filled (unless we increase γ in to the point that higher potential A ~ implicit in the Hamiltonian (5), which explic-
bands are no longer nearly empty). In contrast, the trans- itly has a preferable axis. Choosing a symmetric artificial
verse response is somewhat less sensitive to changes in vector potential (which is not related to the Landau case
γ in , and we find that only changes of order of magnitude by a gauge transformation, as noted in Sec. IV B) in-
have an appreciable effect (Fig. 8b): A small change of stead removes this anisotropy, as one would expect. In
the refilling rate does little to change the occupation of Fig. 10 we show calculation of different elements in the
the different bands, leading to small deviations from the conductance matrix for both choices of A ~ we have dis-
quantized value. cussed. For the longitudinal conductance, i.e., Gxx and
Gyy , we find not only different values, but also differ-
ent functional form, which depends on both A ~ and the
B. Incompatible hopping direction. A new trend of the conductance that goes as
 3
∝ tγinc
0 is found for Gyy with the Landau choice and for
Similarly to what we have seen for the steady state per-
Gxx in the symmetric choice, establishing that the lon-
sistent currents, things change when one has an Hamilto-
gitudinal conductance in these cases is only finite when
nian incompatible with the magnetic flux in (5). Whereas
changes in h and P0 induced by finite tinc are taken into
in the dissipative regime with tinc  γ 0 the conduc-
account in (23). The transverse response however, shows
tance features similar dependence on tγinc 0 as in the co-  2
tinc
herent case, at higher γ 0 the conductance begins to de- a more universal trend of ∝ tγinc0 , albeit with different
cline in amplitude, never reaching the topological quan- amplitudes for Gxy , Gxy in the asymmetric case and for
8

tem. On the other hand, a hopping Hamiltonian lacking


the compatible magnetic field, induces bulk currents in
a pattern and a direction determined by the chosen “ar-
tificial gauge”. Near the edges the net flow to/from the
reservoirs becomes finite locally (though it sums up to
zero globally), facilitating current backscattering at the
edges via the particle reservoir.

The electrical conductance of the dissipative steady


state was also examined. Quantization of the transverse
conductance, consistent with the theoretical Chern num-
ber, was shown to arise for the compatible Hamiltonian,
provided it is sufficiently larger as compared to the dissi-
pative rates. The quality of this quantization is affected
by the accuracy of the assigned dissipative parameters,
FIG. 10. Different elements of the conductance matrix, cal-
culated for the incompatible Hamiltonian in the dissipative namely the refilling rate γ in and µ∗ , which allows to tune
~ The conductance for
regime for the different choices of A. the evaporation rate of the filled band. We find the quan-
both directions in the Landau choice (red and blue), and for tization is much less sensitive to the former compared to
the symmetric choice (yellow) is presented for (a) the longi- the latter. A weaker compatible Hamiltonian featured
tudinal and (b) the transverse elements. For the symmetric mainly currents in the vicinity of the edges mediated by
choice we present only one element of each, since the remain- a dissipative current, together with a weak second or-
ing ones are identical. The behavior as a function of tγinc
0 is
der transverse response, also located near the edges of
indicated next to each plot. The calculations are performed
the system. The conductance matrix in the incompati-
with edges in both directions, with Lx = Ly = 56, α = 17 ,
ble regime is slightly more complicated in structure, as
Llead = 7, and the optimal values for µ∗ and γ in .
the choice of effective vector potential in the dissipative
interaction may render it anisotropic: we observe longi-
Gyx in the symmetric one. tudinal response in the x direction identical to the one
obtained with a weak compatible Hamiltonian, and dif-
ferent power law trends depending on the choice of A. ~
VI. CONCLUSIONS The transverse response seems to have a universal ∝ t2inc
behavior, but with anisotropic vector potential depen-
dent absolute values.
We investigate the transport properties of the purely
dissipative theoretical scheme, presented in Ref. [31],
which reproduces a state synonymous with a very low
temperature equilibrium quantum Hall state in a 2D lat- Under the circumstances of a compatible hopping
tice exploiting engineered dissipation. These properties Hamiltonian, sufficiently larger in amplitude than the
can not be probed without introducing some coherent dissipative energy scale, the potential for obtaining some
Hamiltonian dynamics for the lattice particles. We have of the equilibrium quantum Hall properties, such as chi-
demonstrated that a departure from the completely dis- ral edge-states and quantized conductance, may be put
sipative scheme, crucial for the transport study, as well to the test by checking its robustness to some finite
as being more experimentally realistic, does not necessar- amount of disorder, both in the bath interaction and lat-
ily harm the engineered steady state, provided that the tice Hamiltonians, which will be the subject of a future
Hamiltonian is compatible with the evaporative part, or study. The way the transverse current is carried through
alternatively, that the incompatible part of that Hamil- the system in its dissipative steady state, under such per-
tonian is sufficiently small in magnitude compared to the turbations, can be probed using the tools we have devel-
dissipative energy scale. oped, and may shed further insights into the evaporative
Having introduced the Hamiltonian dynamics, we processes occurring in the system. This also opens up the
could explore the persistent currents that flow once the possibility for engineering increasingly complex coupling
dissipative steady state is reached in the system. In the to the bath, allowing exploration of more interesting pos-
case where the lattice Hamiltonian has the same mag- sibilities and topological traits through the engineering of
netic field present in the evaporative dynamics the main dissipative two-particle interactions. This in turn would
observation is the well-known chiral edge-modes, charac- allow the exploration of the dissipative analogues of ex-
teristic of equilibrium quantum Hall and Chern insulator otic fractional states [39, 40], and perhaps the equivalents
systems, and the lack of any other currents in the sys- of anyons [22, 41, 42].
9

VII. ACKNOWLEDGMENTS namics of this matrix is given by

d Dh iE
hAB a†C aD , a†A aB
X
We would like to thank B. A. Bernevig, B. Bradlyn, PCD (t) = −i
J. I. Cirac, Y. Gefen, A. Kemenev, N. Schuch, H. H. dt t
A,B
Tu, and P. Zoller for useful discussions. This work has * ( )+
X (1) † † † a†C aD
been supported by the Israel Science Foundation (Grant + ΓAB aA aC aD aB − aA aB ,
No. 227/15), the German Israeli Foundation (Grant No. 2
A,B t
I-1259-303.10), the US-Israel Binational Science Founda-
* ( )+

X (2) a aD
tion (Grants No. 2014262 and 2016224), and the Israel + ΓAB aB a†C aD a†A − aB a†A , C .
2
Ministry of Science and Technology (Contract No. 3- A,B t
12419). (A5)

Using the definition for PAB (t), Wick’s theorem, and the
fermionic anti-commutation relations, we find the matrix
Appendix A: Steady state occupation numbers
equation

In this appendix, we re-derive some of the results of d 1 n (1) o


P (t) = −i [h, P (t)] − Γ + Γ(2) , P (t) + Γ(2) .
Ref. [43], and bring the parts relevant to our discussion dt 2
for the reader’s convenience. We begin by introducing a (A6)
quadratic Hamiltonian for the lattice particles a (which The steady state version of Eq. (A6) can be manipulated
in this work will be either the compatible Hamiltonian into a Sylvester equation for the matrix P , Eq. (16) of
(13), or the incompatible Hamiltonian (14)), the main text. If Γ(1) and Γ(2) can be simultaneously
diagonalized, such as in our case, where Γ(2) is a constant
times unity matrix, the steady solution for (16) in the
hAB a†A aB , purely dissipative regime (h → 0) is given (in the basis
X
H= (A1)
A,B
where the Γ matrices are diagonal) by

with generalized indexes A, B, which may each represent Γ(2)


P = . (A7)
for example two spatial indexes (e.g., A = lx , ly and B = Γ(1) + Γ(2)
lx + 1, ly − 1).
This reduces to Eq. (11) in the main text. Notice how-
The dissipator is also assumed to be quadratic, so that
ever, that this result remains intact even in the presence
the Lindbladian master equation for the density matrix
of an Hamiltonian h which is diagonal in the same basis
is
as the Γ matrices, due to the cancellation of the commu-
d h i tator term in (A6). This is the reason for the distinction
hAB a†A aB , ρ +
X
ρ = −i between the compatible and the incompatible Hamiltoni-
dt
A,B ans: In the former case (A7) holds, but not in the latter.
X (1)  † 1n † o
ΓAB aB ρaA − a aB , ρ +
2 A
A,B
Appendix B: The current operators
X (2)  † 1n †
o
ΓAB aA ρaB − aB aA , ρ , (A2)
2
A,B Consider the expectation value of the particle den-
sity
D in the site E (lx , ly ) of the lattice, nlx ,ly (t) ≡
with the Γ matrices encapsulating the dissipative pro- †
ρ (t) alx ,ly alx ,ly = Plx ,ly ;lx ,ly (t). Its time-evolution is
cesses. In this work, we have the matrices
given by
2π d
Γ(1) = γ out = 2
ν0 Hdiss , (A3) nl ,l (t) = ṅH D
~ lx ,ly (t) + ṅlx ,ly (t) , (B1)
dt x y
with
Γ(2) = γ in . (A4) X D E
ṅH
lx ,ly (t) = −i hlx ,ly ;nx ,ny a†nx ,ny alx ,ly
t
We define the Dfollowing matrix of expectation val- nx ,ny
E X D † E

ues PAB (t) = aA aB , with the shorthand notation +i alx ,ly anx ,ny hnx ,ny ;lx ,ly , (B2)
t t
hM it = Tr {ρ (t) M }. According to Eq. (A2), the dy- nx ,ny
10

1 X h (1) i D E
ṅD
lx ,ly (t) = − Γ + Γ(2) a†nx ,ny alx ,ly
2 n ,n lx ,ly ;nx ,ny t
x y

1 D E h i
a†lx ,ly anx ,ny
X
− Γ(1) + Γ(2)
2 n ,n t lx ,ly ;nx ,ny
x y

(2)
+ Γlx ,ly ;lx ,ly . (B3)

Plugging in Hcomp , we get


D E

ṅH
lx ,ly (t) = −it comp a a
lx +1,ly lx ,ly
D E t

+ itcomp alx ,ly alx −1,ly FIG. 11. Distribution of the transverse current along the
t
D E system, with the voltage applied in the x direction. (a) The
ilx 2πα †
− itcomp e alx ,ly +1 alx ,ly perturbation used is δH [Eq. (22)] with Llead = 7. (b) Here
D Et we use δHE with a constant electric field throughout the sam-
ilx 2πα † ple. The black dashed line emphasizes the position of zero
+ itcomp e alx ,ly alx ,ly −1 + h.c. . (B4)
t current, and that in panel (b) there is a significant bulk re-
sponse. Parameters used for both cases are Lx = Ly = 105,
The r.h.s. can be written as a discrete divergence, α = 17 , tcomp = 100γ 0 , and the optimal values for µ∗ and γ in .
   
y y
ṅH x x
lx ,ly (t) = − jlx +1,ly − jlx ,ly − jlx ,ly +1 − jlx ,ly , R
(B5) we incorporate a term synonymous with dxρ (x) φ (x),
which allows us to define the currents as in Eq. (20). by using the perturbation
A transformation tcomp → tinc , α → 0, then gives the
(Ex lx + Ey ly ) a†lx ,ly alx ,ly ,
X
δHE = − (C2)
current in the presence of Hinc , Eq. (21).
lx ,ly
A closer look at ṅD reveals that it can be separated
into two terms, corresponding, respectively, to flow of with the electric field E = (Ex , Ey , 0). The rest of the
particles from the refilling reservoir into the system, and calculations take place by solving Eq. (23), as usual.
particles leaving the system into the evaporative reser- Whereas the calculated conductance values and their
voir, trends remain the same when using δHE , the current
distributions are somewhat different, as illustrated in
Fig. 11 for the case of a compatible Hamiltonian, in
1 n (2) o 1 n (1) o
ṅD
lx ,ly (t) = Γ ,1 − P − Γ ,P the tcomp  γ 0 regime. Most notably, the bulk is more
2 lx ,ly ;lx ,ly 2 lx ,ly ;lx ,ly
active in carrying the current when the electric field is
≡ JlD,in
x ,ly
− JlD,out
x ,ly
. (B6) applied throughout the sample, in contrast to the case of
voltage leads, where the response is localized around the
Also note that for a compatible Hamiltonian, plugging leads (which is the only place the voltage drops).
in the steady state P , given by Eq. (A7), leads to the
steady state ṅD vanishing exactly. More generally, in the
d
steady state dt n = 0, hence the “dissipative current” is
immediately found (for any choice of Hamiltonian) to be
[1] C. W. Gardiner and H. Haken, Quantum noise, Vol. 2
ṅD
lx ,ly = −ṅH
lx ,ly , (B7) (Springer Berlin, 1991).
[2] O. Hirota, A. S. Holevo, and C. M. Caves, Quan-
with the r.h.s. given by Eq. (B5). tum Communication, Computing, and Measurement
(Springer Science & Business Media, 2012).
[3] J. F. Poyatos, J. I. Cirac, and P. Zoller, Phys. Rev. Lett.
77, 4728 (1996).
Appendix C: Conductance with a constant electric [4] S. Diehl, A. Micheli, A. Kantian, B. Kraus, H. P. Büchler,
field and P. Zoller, Nature Physics 4, 878 EP (2008).
[5] B. Kraus, H. P. Büchler, S. Diehl, A. Kantian, A. Micheli,
A possible alternative to the definition we present in and P. Zoller, Phys. Rev. A 78, 042307 (2008).
Sec. V is to use a perturbative Hamiltonian which mod- [6] F. Verstraete, M. M. Wolf, and J. Ignacio Cirac, Nature
Physics 5, 633 EP (2009).
els the application of an electric field along the sample. [7] H. Krauter, C. A. Muschik, K. Jensen, W. Wasilewski,
Using the relation between the electric field E and the J. M. Petersen, J. I. Cirac, and E. S. Polzik, Phys. Rev.
electrostatic potential φ, Lett. 107, 080503 (2011).
[8] J. Otterbach and M. Lemeshko, Phys. Rev. Lett. 113,
E = −∇φ, (C1) 070401 (2014).
11

[9] E. Kapit, M. Hafezi, and S. H. Simon, Phys. Rev. X 4, [26] R. König and F. Pastawski, Phys. Rev. B 90, 045101
031039 (2014). (2014).
[10] G. Lindblad, Comm. Math. Phys. 48, 119 (1976). [27] E. Kapit, M. Hafezi, and S. H. Simon, Phys. Rev. X 4,
[11] I. M. Georgescu, S. Ashhab, and F. Nori, Rev. Mod. 031039 (2014).
Phys. 86, 153 (2014). [28] J. C. Budich, P. Zoller, and S. Diehl, Phys. Rev. A 91,
[12] I. Bloch, J. Dalibard, and W. Zwerger, Rev. Mod. Phys. 042117 (2015).
80, 885 (2008). [29] F. Iemini, D. Rossini, R. Fazio, S. Diehl, and L. Mazza,
[13] X. G. Wen, Advances in Physics 44, 405 (1995). Phys. Rev. B 93, 115113 (2016).
[14] K. v. Klitzing, G. Dorda, and M. Pepper, Phys. Rev. [30] Z. Gong, S. Higashikawa, and M. Ueda, Phys. Rev. Lett.
Lett. 45, 494 (1980). 118, 200401 (2017).
[15] D. C. Tsui, H. L. Stormer, and A. C. Gossard, Phys. [31] M. Goldstein, “Dissipation-induced topological insu-
Rev. Lett. 48, 1559 (1982). lators: A no-go theorem and a recipe,” (2018),
[16] C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 146802 arXiv:1810.12050.
(2005). [32] P. G. Harper, Proceedings of the Physical Society. Section
[17] B. A. Bernevig, T. L. Hughes, and S. Zhang, Science A 68, 874 (1955).
314, 1757 (2006). [33] D. R. Hofstadter, Phys. Rev. B 14, 2239 (1976).
[18] M. König, S. Wiedmann, C. Brüne, A. Roth, H. Buh- [34] D. Jaksch and P. Zoller, New Journal of Physics 5, 56
mann, L. W. Molenkamp, X. Qi, and S. Zhang, Science (2003).
318, 766 (2007). [35] F. Gerbier and J. Dalibard, New Journal of Physics 12,
[19] L. Fu and C. L. Kane, Phys. Rev. Lett. 100, 096407 033007 (2010).
(2008). [36] P. A. M. Dirac, Proceedings of the Royal Society of Lon-
[20] R. M. Lutchyn, E. P. A. M. Bakkers, L. P. Kouwenhoven, don A 114, 243 (1927).
P. Krogstrup, C. M. Marcus, and Y. Oreg, Nature Re- [37] L. Chen, T. Mazaheri, A. Seidel, and X. Tang, Journal
views Materials 3, 52 (2018). of Physics A: Mathematical and Theoretical 47, 152001
[21] N. P. Armitage, E. J. Mele, and A. Vishwanath, Rev. (2014).
Mod. Phys. 90, 015001 (2018). [38] N. Read, Phys. Rev. B 95, 115309 (2017).
[22] C. Nayak, S. H. Simon, A. Stern, M. Freedman, and [39] J. K. Jain, Phys. Rev. Lett. 63, 199 (1989).
S. Das Sarma, Rev. Mod. Phys. 80, 1083 (2008). [40] M. Levin and A. Stern, Phys. Rev. Lett. 103, 196803
[23] S. Diehl, E. Rico, M. A. Baranov, and P. Zoller, Nature (2009).
Physics 7, 971 EP (2011). [41] K. P. Schmidt, S. Dusuel, and J. Vidal, Phys. Rev. Lett.
[24] C.-E. Bardyn, M. A. Baranov, E. Rico, A. İmamoğlu, 100, 057208 (2008).
P. Zoller, and S. Diehl, Phys. Rev. Lett. 109, 130402 [42] B. Bauer, L. Cincio, B. P. Keller, M. Dolfi, G. Vidal,
(2012). S. Trebst, and A. W. W. Ludwig, Nature Communica-
[25] C.-E. Bardyn, M. A. Baranov, C. V. Kraus, E. Rico, tions 5, 5137 EP (2014).
A. İmamoğlu, P. Zoller, and S. Diehl, New Journal of [43] F. Schwarz, M. Goldstein, A. Dorda, E. Arrigoni, A. We-
Physics 15, 085001 (2013). ichselbaum, and J. von Delft, Phys. Rev. B 94, 155142
(2016).

You might also like