You are on page 1of 20

Effects of atom losses on a one-dimensional lattice gas of hardcore bosons

François Riggio, Lorenzo Rosso† , Dragi Karevski and Jérôme Dubail



Université Paris-Saclay, CNRS, LPTMS, 91405 Orsay, France
Université de Lorraine, CNRS, LPCT, F-54000 Nancy, France ∗

Atom losses occur naturally during cold atoms experiments. Since this phenomenon is unavoid-
able, it is important to understand its effect on the remaining atoms. Here we study a gas of
hard-core bosons on a lattice subject to K-body losses (where K = 1, 2, 3, . . . is the number of
atoms lost in each loss event), and in particular we investigate the effect of losses on the rapidity
arXiv:2307.02298v1 [cond-mat.quant-gas] 5 Jul 2023

distribution ρ(k) of the atoms. Under the assumption that losses are weak enough so that the sys-
tem relaxes between two loss events, we are able to determine the loss functional F [ρ](k) encoding
the loss process for K-body losses. We derive closed expressions for the cases of one- and two-body
losses, and show their effects on the evolution of the total number of particles. Then we add a
harmonic trapping potential and study the evolution of the position-dependent rapidity distribution
of this system by solving numerically the evolution equation for one-, two- and three-body losses.

I. INTRODUCTION a singular property: the stationary states of the system


are modeled by Generalized Gibbs Ensembles (GGE). A
During the past two decades, cold atom experiments GGE is constructed by maximizing the entropy under
have become a key simulation platform to investigate the constraints imposed by all conservations laws [52].
the physics of low-dimensional quantum many-body sys- However, under dissipative processes like atom losses, the
tems [1]. While these experiments are typically well iso- conserved quantities of an integrable system are not con-
lated and their dynamics at short times is well approx- served anymore, as the coupling between the system and
imated by unitary dynamics, cold atom setups are al- its environnement typically causes integrability breaking
ways subject to atom losses. The effects of the latter on [53]. In principle, one then expects that integrability
the dynamics of the gas can become important at longer breaking leads to thermalization at very long times, pos-
times, and they are typically hard to describe theore- sibly with a pre-thermalization phenomenon at interme-
tially. Cold atom experiments involve various loss pro- diate time scales [54]. Some studies seem to confirm the
cesses, for instance one-body losses resulting from scat- thermalization but other works suggest otherwise [17].
tering with background thermal atoms [2], which can be In this paper we propose a model of hard-core bosons
significant. Inelastic two-body collisions can occur natu- on a lattice subject to weak K-body atom losses: when-
rally or be intentionally engineered, leading to two-body ever K atoms occupy K neighboring sites, they can es-
losses [3–10]. Three-body losses, where a highly bound cape the system with some rate. Our aim is to un-
diatomic molecule is formed, are invariably present and derstand how integrability breaking caused by such loss
typically dominate the overall loss process [11–13]. In events affects the system. Of course, under atom losses,
principle, loss processes involving more than three atoms the stationary state at very long times is always the vac-
also exist. In particular, losses involving four atoms have uum. However, what we want to investigate is how the
been reported in Refs. [14, 15] vacuum is approached, and in particular whether or not
Loss processes can sometimes be controlled and engi- to the gas is in a quasi-stationary thermal state. We in-
neered [16] to bring out some physical phenomena such vestigate for instance the mean density n(t) = hN (t)i /L
as cooling [17–24] and the quantum Zeno effect [25–39]. (where N is the number of atoms and L is the number of
Despite the relevance of loss processes, a consistent gen- lattice sites) for different initial states and observe differ-
eral theory describing the loss events does not exist. One ent non-trivial behaviors depending on the number K of
standard way to model atom losses is to use the Lindblad atoms lost in each loss event. In particular, for two-body
equation [25, 26, 40–42]. Several studies have inspected losses (K = 2), we analytically show that n(t) ∝ 1/t
the interplay between the unitary dynamics and the lossy or ∝ 1/t1/2 , depending on a specific property of the ini-
one in both bosonic [25, 26, 40, 43, 44] and fermionic tial state (see below). More generally, for other values
gases [44–51]. of K ≥ 2 our numerical results are in agreement with a
The question of losses in integrable systems is partic- power-law decay n(t) ∝ tα where the exponent α typi-
ularly interesting. An integrable model admits a macro- cally depends on the initial state, and generically differs
scopic number of conserved charges in comparison with from the mean-field result 1/(K − 1). This observation
non-integrable models where only the energy and/or the holds also if we add a harmonic potential; however in
number of particle are conserved. Due to this large set that case the exponent α changes and is different from
of conserved quantities, an isolated integrable system has the one found in the homogeneous case.
The paper is organized as follows. In Section II we de-
fine the model and discuss our assumptions. Our main
hypothesis of slow losses, and fundamental concepts such
∗ francois.riggio@univ-lorraine.fr as the rapidity distribution and the loss functional are
2

K=3 with nearest-neighbor hopping,


1X + − +
HHCB = − (σj σj+1 + σj+1 σj− ). (1)
2
j∈Z

This Hamiltonian generates the unitary part of the evo-


K=2 lution of the gas. In addition, we assume that the gas is
subject to incoherent K-body loss processes, with K a
positive integer. To describe the losses, we assume that
the dynamics is Markovian, and we consider the following
Lindblad equation for the density matrix ρ̂,
K=1 ˙ = −i[HHCB , ρ̂(t)]
ρ̂(t)
X † 1 †

+Γ Lj ρ̂(t) Lj − {Lj Lj , ρ̂(t)} , (2)
2
j∈Z

Figure 1. We consider a lattice gas of hardcore bosons (pink Here Γ is a constant that sets the loss rate, while the
dots) that can jump from site to site. Two bosons cannot be Lindblad operators
on the same site. Losses are represented by the large black
arrow which connects the bosons chain to its environment K−1
Y
(blue background). Here we show the situation for one-body −
Lj = σj+l (3)
losses (K = 1), two-body losses (K = 2) and three-body losses l=0
(K = 3). For each cases, K consecutive atoms are removed
and are lost in the ‘environment’. In this paper losses are remove K bosons from the K consecutive sites j, j +
modeled by the Lindblad equation (2). 1, . . . , j + K − 1.
We stress that, with these loss terms, the model is
not exactly solvable, so it is necessary to develop some
effective approaches to tackle it.
introduced. In Section III we present our calculation
of the loss functional for K-body losses, which crucially
depends on the parity of K. In Section IV we investi- B. Adiabatic losses, effective description by slow
gate the effect of atom losses on the rapidity distribution motion of the charges
and on the mean density in the homogeneous gas, by
combining analytical and numerical calculations. In Sec-
To simplify the description of the system, we follow the
tion V, we add a harmonic trapping potential. We use a
approach of Ref. [40]. There the approach was developed
hydrodynamic-like approach similar to Generalized Hy-
for losses in a continuous gas, and it is based on the as-
drodynamics [55, 56] where we incorporate the loss func-
sumption that the losses are slow, so that the gas remains
tional (following similar proposals in Refs. [40, 43]), and
in a Generalized Gibbs Ensemble (GGE) with parame-
design a numerical method for solving the resulting evo-
ters that slowly drift in time (see e.g. [57] for a review).
lution equation for the rapidity distribution in the gas.
A similar description of slowly-evolving nearly integrable
We conclude in Section VI.
systems with weak integrability breaking had previously
appeared in Refs. [58–60]. Following these ideas, here we
also assume that the loss processes occur on very long
times compared to the relaxation time scale due to the
II. THE MODEL unitary evolution of the gas. In that limit, the gas has
time to reach a local stationary state after each loss event.
A. Definition: lattice Tonks-Girardeau gas with To efficiently exploit that assumption, we look at the slow
K-body losses dynamics of the conserved charges.
For convenience, from now on we focus on a finite sys-
tem of L ≫ 1 sites, with periodic boundary conditions.
We consider a lattice Tonks-Girardeau gas subject to The Hamiltonian HHCB commutes with an infinite set
atom losses. Each site j ∈ Z is occupied by either zero of hermitian operators Qa , a = 0, 1, 2 . . . that can be
or one boson. We write σj+ /σj− for the operator that cre- constructed using the Jordan-Wigner mapping to non-
ates/annihilates a boson on site j. Because of the hard- interacting fermions (see Subsec. II C for details). These
core constraint, these operators do not satisfy the usual operators also commute among themselves, [Qa , Qb ] =
bosonic canonical commutation relations. Instead, they [HHCB , Qa ] = 0. Moreover, they are local in the sense
satisfy the algebra of Pauli matrices, (σj+ )2 = (σj− )2 = 0, PL
that Qa = j=1 qa,j where qa,j is a charge density op-
and [σi+ , σj− ] = δi,j σjz . erator that has compact support (i.e. it acts on a finite
We consider the hard-core boson (HCB) Hamiltonian number of sites around j).
3

The time evolution of the expectation value hQa i (t) = the Hamiltonian (1) becomes
tr[ρ̂(t)Qa ], is obtained from Eq. (2),
L
1X †
L
Γ XD HHCB = − (c cj+1 + c†j+1 cj ). (9)
2 j=1 j
E
hQ˙a i(t) = L†j [Qa , Lj ] + [L†j , Qa ]Lj (t). (4)
2 j=1
Moreover, the fermions satisfy antiperiodic (resp. peri-
Moreover, the hermiticity of Qa implies h[L†j , Qa ]Lj i = odic) boundary conditions if the number of particles N
in the system is even (resp. odd):
hL†j [Qa , Lj ]i∗ . Thus,

L
c†L+1 = (−1)N −1 c†1 . (10)
D E
hQ˙a i(t) = Γ L†j [Qa , Lj ]
X
Re (t). (5) The Fourier modes are
j=1
L
1 X ikj †
This equation is exact and it is a direct consequence of c† (k) = √ e cj (11)
Eq. (2). Notice that, in this form, it is not particularly L j=1
useful, because to evaluate the r.h.s one needs to know
the exact density matrix ρ̂(t). with k ∈ 2π 1
L (Z + 2 ) if N is even, and k ∈

L Z if N is
Now comes the crucial step. Importantly, the operator odd. Either way the Hamiltonian reads

Lj [Qa , Lj ] is local, because both the operator Lj and the X
charge density qa,j have compact support. This, together HHCB = ε(k)c† (k)c(k), (12)
with the assumption of slow losses, allows us to use the k
idea of local relaxation in the system. Namely, we expect
with ε(k) = − cos k.
that, under unitary evolution, the density matrix of a
It is clear from the form (12) that any operator of the
small subsystem quickly relaxes to a Generalized Gibbs
form
Ensemble (GGE). Expectation values of local observables
can then be evaluated with respect to the GGE density
X
Q[f ] = f (k)c† (k)c(k), (13)
matrix, k
P
− βa Qa
ρ̂GGE,{hQa i} ∝ e a , (6) for any function f (k), commutes with the Hamiltonian
HHCB . Moreover these conserved charges also com-
where the Lagrange multipliers βa are fixed by the expec- mute among themselves. Convenient choices for f (k) are
tation values of the charges hQa i, which must be equal to cos(nk) or sin(nk) for n ∈ N, which leads to a hermi-
tr[ρ̂GGE,{hQb i} Qa ]. Evaluating the r.h.s of Eq. (4) in the tian basis set of charges, where each charge has a charge
GGE density matrix leads to a closed evolution equation density that is compactly supported.
for the slow motion of the charges induced by the losses, However, for the purposes of this paper, rather than
L to work with a specific choice of basis for the space of
d D E
conserved charges Qa (or Q[f ]), it is more convenient to
Re L†j [Qa , Lj ]
X
hQa i = Γ . (7)
dt j=1
GGE,{hQb i} work directly with the occupation number, or ‘rapidity
distribution’,
It is this evolution equation that we study in great detail

in this paper. For lattice HCB, the description is fur- ρ(k) = c (k)c(k) , ρ(k) ∈ [0, 1]. (14)
L→∞
ther simplified by specifying the form of the conserved
charges Qa . This is what we do next, by introducing the It is clear that if we know the rapidity distribution ρ(k),
distribution of rapididites. then we also know the expectation
P values of any charge
Q[f ], because hQ[f ]i = k f (k)ρ(k).
Following Ref. [40], we can turn the evolution equation
C. Slow evolution of the rapidity distribution for the slow motion of the charges (4) into an equation
for the slow evolution of the rapidity distribution itself,
Hard-core bosons can be mapped to free fermions by
a Jordan-Wigner transformation, ρ̇(k) = −ΓF [ρ](k), (15)

j−1 j−1 where the loss functional



c†i ci
(−1)ci ci c†j ,
Y Y
σj+ = σj− = (−1) cj . (8) L D E
Re L†j [Lj , c† (k)c(k)]
X
i=1 i=1 F [ρ](k) = , (16)
GGE,ρ
j=1
Here the operators c†j /cj create/annihilate a fermion on
site j. They satisfy the canonical anticommutation rela- and the GGE density matrix itself is parameterized
tions {ci , c†j } = δij . Under the Jordan-Wigner mapping, by the rapidity distribution ρ(k). More precisely,
4

the GGE density matrix is Gaussian for the fermions the Fourier modes of the fermions with periodic bound-
c† , c , and it is characterized by its two-point function ary conditions to the ones with anti-periodic boundary

j † j conditions. For conciseness, let us introduce the two cor-


c (k)c(k ′ ) GGE,ρ = ρ(k)δk,k′ . All higher-order cor-

responding sets of momenta,
relations can be computed using Wick’s theorem for
fermionic operators. 2π
The functional (16) is the central object of this paper. Qp = × {1, 2, . . . , L}, (19)
L
In the next section we compute it explicitly for K-body 2π 1 3 1
loss processes. For one- and two-boson loss processes we Qap = × { , , . . . , L − }. (20)
L 2 2 2
get simple closed expressions. For loss events involving
larger numbers K of bosons, we will see that we can Then we have the following identities,
express the loss functional as a small determinant, which i X ei(q−k)/2
follows from applying Wick’s theorem to Eq. (16). (k ∈ Qp ) c(k) = c(q),
L ap
sin((q − k)/2)
q∈Q
(21)
III. DERIVING THE LOSS FUNCTIONAL i X e i(q−k)/2
(k ∈ Qap ) c(k) = c(q).
L p
sin((q − k)/2)
q∈Q
In this section we compute the functional F [ρ](k) ex-
plicitly. Importantly, in our calculation we uncover a dif- (22)
ferent structure depending on the parity of the number We can insert them into the second term of Eq. (17),
K of bosons lost in each loss event, which can be traced which leads to
back to the Jordan-Wigner string appearing in the map- D E
′ †
ping (8) to non-interacting fermions.
+ † 1 X ei(q−q )/2 c1 c† (q ′ ) c(q)c1
σ1 c (k)c(k)σ1− = 2

.
L ′
sin((q − k)/2) sin((q ′ − k)/2)
q,q
A. One-body losses (23)
This correctly implements the change of boundaries of
For K = 1 the Lindblad dissipators are Lj = σj− . Us- the fermions. Next we can apply Wick’s theoremEto eval-
ing translational invariance, the loss functional (16) that
D
† † ′
uate the four-fermion correlator c1 c (q ) c(q)c1 . This
we need to compute is
leads to
F [ρ](k) = L σ1+ [σ1− , c† (k)c(k)] GGE,ρ


L σ1+ c† (k)c(k)σ1− =

= L σ1+ σ1− c† (k)c(k) GGE,ρ − L σ1+ c† (k)c(k)σ1− GGE,ρ ,





!2
  2
hN i X ρ(q) 1X q−k hN i
(17) − cot ρ(q) − .
L2 q sin2 ( q−k2 )
L q 2 L2
where L is the length of the system and the loss operator (24)
acts on the site j = 1. Both terms in the second line The first term in the above equation has a pole of order 2
of (17) can be computed using the fact that the GGE at q = k + 2πZ, and
is a gaussian state for the fermions, which allows us to Pit is convenient to reduce its degree
using the identity q 1/ sin2 ( q−k
2 ) = L 2
. This leads to
use Wick’s
P theorem. For the first term we have (using the equivalent expression
c1 = √1L q eiq c(q)):
hN i X ρ(q) − ρ(k)
L σ1+ c† (k)c(k)σ1− = 2

+ hN i ρ(k)
sin2 ( q−k
D E
L σ1+ σ1− c† (k)c(k) GGE,ρ = L c†1 c1 c† (k)c(k)

L q 2 )
GGE,ρ !2
  2
i(q−q′ ) 1X p−k hN i
X
† ′ †

= e c (q )c(q)c (k)c(k) GGE,ρ − cot ρ(p) − .
qq′ L p 2 L2
X ′
= ei(q−q ) hc† (q ′ )c(q)ihc† (k)c(k)i+ (25)
qq′
Putting the two terms (18)-(25) together and taking the
hc† (q ′ )c(k)ihc(q)c† (k)i

thermodynamic limit L → ∞, we arrive at the following
= hN iρ(k) + ρ(k)(1 − ρ(k)). (18) form of the one-body loss functional
 π   2
The second term requires more care, because the oper- 2 dp k−p
F [ρ](k) = ρ(k) − ρ (k) + cot ρ(p)
ator c† (k)c(k) is inserted between σ1+ and σ1− , and the −π 2π 2
latter change the parity of the number of particles in the
!
π
dq ρ(k) − ρ(q)
system. The boundary conditions for the fermions are +n n+ 2 k−q
, (26)
modified according to Eq. (10). Thus we need to relate −π 2π sin ( 2 )
5
ffl
where n = hN i /L is the density of particle and means C. K-body losses with K even
the Cauchy principal value of the integral. This is our
main result for one-body losses. It is similar to (but In this subsection we show that is possible to
different from) the formula given in Ref. [40] for the find a closed formula for the loss functional defined
Tonks-Girardeau gas in the continuum. Notice that the in (16) where the Lindblad operator is given by Lj =
functional is non-linear in ρ(k), and also non-local in ra- σj σj+1 . . . σj+K−1 . Taking the Fourier transform of Lj
pidity space. and L†j in (16), the loss functional reads

K
( )
B. Two-body losses even 1 X X
F [ρ](k) = exp i (ql − ql′ )l ×
LK−1 q1 ,...,qK l=1
For K = 2, the dissipators are Lj = σj− σj+1
− q1′ ,...,qK

. Under
the Jordan-Wigner mapping they become

† ′
c (qK ) . . . c† (q1′ )[c(q1 ) . . . c(qK ), c† (k)c(k)] .

† (32)
Lj = σj− σj+1

= cj (−1)cj cj cj+1 = −cj cj+1 . (27)
As mentionned in III B, in the case of even K-body losses
Then, to compute the functional F , we simply need to the commutator in (32) reduces to K terms
insert the dissipator (27) in the definition (16),
† ′
c (qK ) . . . c† (q1′ )[c(q1 ) . . . c(qK ), c† (k)c(k)]

X

+
σj+ [σj− σj+1

, c† (k)c(k)]

F [ρ](k) = σj+1 = δkqK c† (qK


) . . . c† (q1′ )c(q1 ) . . . c(qK−1 )c(k)

j
− δkq(K−1) c† (qK ′
) . . . c† (q1′ )c(q1 ) . . . c(qK−2 )c(qK )c(k)


X ei(2p′ +p−2q−q′ )

c† (q)c† (q ′ )[c(p)c(p′ ), c† (k)c(k)] .



= + ... (33)
′ ′
M
q,q ,p,p
(28) and applying Wick’s theorem on
each terms leads to a
product of
K terms of the form c† (q ′ )c(q) . Using the
Expanding the commutator in the braket leads to two property c† (q ′ )c(q) = δqq′ ρ(q) and taking the ther-
terms modynamic limit, the loss functional in (32) can be ex-

† pressed as a sum of K terms each consisting of a K by K
c (q)c† (q ′ )[c(p)c(p′ ), c† (k)c(k)]

matrix determinant. Let us introduce the K × K matrix
(j)
= c† (q)c† (q ′ )c(p)c(p′ )c† (k)c(k)


A[ρ] with matrix elements
− c† (q)c† (q ′ )c† (k)c(k)c(p)c(p′ ) .


(29) π
1 i(b−a)q
 ´
(j)
[A[ρ] ]ab = 2π −π dq e ρ(q) if b 6= j
(34)
The first term can be expressed as i(b−a)k
e ρ(k) if b = j,


c (q)c† (q ′ )c(p)c(p′ )c† (k)c(k)

for indices a, b = 1, . . . , K. The superscript j indicates
= c† (q)c† (q ′ )c† (k)c(k)c(p)c(p′ )

which column depends on the rapidity k. Apart from
the j th column, the matrix essentially contains Fourier
+ δp′ k c† (q)c† (q ′ )c(p)c(k) − δpk c† (q)c† (q ′ )c(p′ )c(k) .



transforms of the rapidity distribution ρ(k). The loss
(30) functional then reduces to
K
The second term is the expectation value of c† (k)c(k) in even
X 
(j)

a state where two atoms have been removed, the parity of F [ρ](k) = det A[ρ] . (35)
the initial number of particle is then unchanged. Hence, j=1

in contrast to the K = 1 case treated in the previous


subsection, here the parity of the number of atoms does The relation (35) is another fundamental result of this
not change. paper.
One can then apply Wick’s theorem, and take the ther-
modynamic limit to obtain the loss functional,
D. K-body losses with K odd
2 π
 
k−q
ˆ
F [ρ](k) = dq sin2 ρ(q) ρ(k). (31)
π −π 2 In this last subsection we investigate the case of losses
process with an odd number K of lost atoms. The rea-
This is our main result for two-body losses. That func- soning is similar to the one we developped in the previous
tional presents some similarities with the loss functional subsection, however like in the K = 1 case, we need to
found in the relation (5) of Ref. [25]. be careful about the change of boundary conditions for
We now generalise this calculation to the case of K- the fermions. We start by taking the Fourier transform
body losses for K an arbitrary even integer. of the Lindblad operators in the definition (16), which
6

0.0 % lost 25.9 % lost 55.1 % lost 75.3 % lost 91.8 % lost
9.5 % lost 33.0 % lost 63.2 % lost 79.8 % lost 98.1 % lost
18.1 % lost 45.1 % lost 69.9 % lost 86.5 % lost Eq.45

T = 0.1 T =1 T = 10 T = 100
1.0 0.5
0.5
0.8 0.6 0.4
0.4
0.6
ρ(k, t)

0.4 0.3 0.3


0.4 0.2 0.2
0.2
0.2 0.1 0.1
0.0 0.0 0.0 0.0
−π 0 π −π 0 π −π 0 π −π 0 π
Rapidity k Rapidity k Rapidity k Rapidity k
Figure 2. Effect of one-body losses on different rapidity distributions. Taking the initial distribution as ρ(k) = (1 + exp(− cos(k)/T ))−1 ,
we change the initial distribution by modifying the temperature T . The time evolution is obtained solving (15) with Runge-Kutta method.
From left to right the temperature is decreasing T = 0.1, 1.0, 10.0, 100.0. The black dashed lines correspond to formula (45) and show a
perfect agreement with numerical simulations.

leads to the relation (32). The commutator in the loss term in the right-hand side of (38) can be written as two
functional gives two terms determinants of two matrices B and C. The matrices

† ′ B and C are (K + 1) × (K + 1) hermitian matrices and
c (qK ) . . . c† (q1′ )[c(q1 ) . . . c(qK ), c† (k)c(k)]

their matrix elements depend on the Fourier and Hilbert
= c† (qK′
) . . . c† (q1′ )c(q1 ) . . . c(qK )c† (k)c(k) transforms [61] of ρ(k)

− c† (qK′
) . . . c† (q1′ )c† (k)c(k)c(q1 ) . . . c(qK ) .

 1 ´π
(36)  2π −π dq ei(b−a)q ρ(q) if a, b < K + 1
[B[ρ] ]ab = e−iak ρ(k) if b = K + 1
As we have already discussed in the subsection III A,
0 if a = b = K + 1,

the first term is the expectation value of c† (k)c(k) in a
state where the initial number of particles is preserved. (39)
However the second term corresponds to the expectation
value of c† (k)c(k) in a state where K atoms have been 1
´π
dq ei(b−a)q ρ(q) if a, 
b<K+ 1

2π −π
removed. Since K is an odd number, the parity of the

k − q


 1
ffl π
−i(a−1)q
number of particle is changed and one needs to use the 2π −π dq e ρ(q)(cot + i)


2

relations (21) to express c† (k)c(k) in the appropriate par-

[C[ρ] ]ab = if b = K + 1
ity sector. Inserting the relations (21) in the second term, 
 ffl π ρ(q) − ρ(k)
1
dq if a = b = K + 1.

one has 
2π −π k−q


sin2 (


† ′  )
c (qK ) . . . c† (q1′ )c† (k)c(k)c(q1 ) . . . c(qK ) =

2
(40)
X ei(q−q′ )/2 c† (qK ′
) . . . c† (q1′ )c† (q ′ )c(q)c(q1 ) . . . c(qK )


. The loss functional for odd K takes the final form

L2 sin((q − k)/2) sin((q ′ − k)/2)
q,q  
(37) K  
(j)
X
odd
FK [ρ](k) =  det A[ρ] 
Before using Wick’s theorem on the above formula, one j=1
can notice that the first term in the right-hand side   
of (36) can be written as + det B[ρ] − det C[ρ] . (41)

† ′ It is possible to write a general expression valid both
c (qK ) . . . c† (q1′ )c(q1 ) . . . c(qK )c† (k)c(k)

1 − (−1)K
= c† (qK


) . . . c† (q1′ )c† (k)c(k)c(q1 ) . . . c(qK )
for even and odd K by introducing the factor
2
which vanishes for K even, so that our final result, valid
+ δkqK c† (qK ′
) . . . c† (q1′ )c(q1 ) . . . c(qK−1 )c(k)


in all cases, reads
− δkq(K−1) c† (qK ′
) . . . c† (q1′ )c(q1 ) . . . c(qK−2 )c(qK )c(k)


 
K
+ ..., (38)
 
(j)
X
FK [ρ](k) =  det A[ρ] 
where we used the anti-commutation relation for the j=1

fermionic operators. As we proceed in the previous sub- 1 − (−1)K   


section, the Wick’s contractions of (37) and of the first + det B[ρ] − det C[ρ] . (42)
2
7

IV. EVOLUTION OF THE RAPIDITY 3.0


DISTRIBUTION IN A HOMOGENEOUS GAS
2.5 a)

ρ(k, t)/n(t)
2.0
Having established the general form of the loss func-
tional F [ρ] for K-body losses, Eqs. (26)-(31)-(35)-(41)- 1.5
(42), we now turn to the time evolution of the rapidity 1.0 t=0
distribution. We solve the evolution equation t→∞
0.5 Fit with a1 ecos(k)/a2
ρ̇(k) = −ΓF [ρ](k) (43) 0.0 π
−π − π2 0 2 π
numerically (and analytically for the special case K = 1),
and we focus in particular on the time evolution of the 2.5 t=0
´π dk t→∞
atom density n = −π ρ(k) 2π . 2.0

ρ(k, t)/n(t)
1.5
A. Results for K = 1 1.0

For one-body losses the atom density always decays 0.5


b)
exponentially, 0.0 π
−π − π2 0 2 π
−Γt
n(t) = n(0)e . (44) Rapidity k
This simply follows from Eq. (4) applied to the total
Figure 3. Long time behavior of different rapidity distributions
particle number Qa = N and to Lj = σj− : it gives rescaled by the corresponding density. a) The initial distribu-
hṄ i = −ΓhN i, which implies Eq. (44) for the atom den- tion (blue curve) is a Fermi-Dirac distributions ρ0 (k) = (1 +
sity n = N/L. exp(− cos(k)/T ))−1 with T =0.1. The red curve is the distribu-
tion at long time according to Eq. (45). The black dashed line is a
It turns out the evolution equation (43) for the rapidity numerical fit with a Boltzmann distribution a1 exp{cos(k)/a2 }. b)
distribution can be solved exactly for the loss functional The initial distribution (blue curve) is a cosinus function ρ0 (k) =
for K = 1 (see Eq. (26)). The solution is derived in (1 − cos(sk))/2. The red curve is the distribution for t → ∞. Here
Appendix A; it reads it is clear that the distribution does not become thermal.
!
tanh n0 (e−Γt − 1) + ni0 I(t, k)

−Γt
ρ(t, k) = n0 e Re ,
1 + ni0 tanh(n0 (e−Γt − 1)) I(t, k)
ρ(k, t)/n(t) in the limit t → ∞, and fit the result with
(45) a Boltzmann distribution. The agreement is very good,
even though it is clear from the exact formulas (45)-(46)
where n0 = n(0) is the initial atom density, ρ0 (k) = ρ(t = that the distribution ρ(k, t) never becomes exactly ther-
0, k) is the initial rapidity distribution, and I(t, k) is the mal, even at infinite time.
integral
ˆ π To find a more striking signature of the fact that the
dq ρ0 (q) system never goes to a low-density thermal distribution,
I(t, k) =  . (46)
−π 2π k−q
tan 2 + in0 (1 − e−Γt ) we consider the case of an oscillating initial rapidity dis-
tribution ρ0 (k) = (1 − cos(sk))/2, where s is an integer.
A related expression for the continuous Tonks-Girardeau In that case the long-time limit of the integral (46) can be
gas with one-body losses was obtained in Ref. [40]. evaluated analytically, lim I(t, k) = (1 − e−2sn0 eiks )/2,
t→∞
In Fig. (2) we show the evolution of the rapidity distri- and when injected in Eq. (45) it leads to a late-time ra-
bution from thermal initial states at different tempera- pidity distribution of the form
tures. We display the analytical result (45), as well as the
numerical solution of Eq. (43) obtained from the Runge-
Kutta method; they are in perfect agreement, as they ρ(k, t) α + β cos(sk)
= , (47)
should. n(t) t→∞ γ + δ cos(sk)
We see that loss processes spread the distribution in
rapidity space. In the limit of large temperature for the
initial state, the rapidity distribution is flat, and remains where the coefficients α, β, γ, δ depend on the initial den-
flat at all times. For smaller initial temperatures, it sity n0 and on the integer s. Thus, even at long time, the
evolves into a bell-shape distribution at late times, which rescaled rapidity distribution is sensitive to the structure
is close to a Boltzmann distribution ρ(k) ∝ exp[cos(k)/T ] of the inital distribution (see also Fig.3.(b)). We con-
with a density going to zero according to Eq. (44). This clude that in general the rapidity distribution does not
is further illustrated in Fig. 3.(a), where we plot the ratio go to a low-density thermal distribution at long times.
8

0.0 % lost 22.7 % lost 45.5 % lost 66.7 % lost 86.6 % lost
8.3 % lost 28.9 % lost 52.1 % lost 73.7 % lost 94.7 % lost
15.2 % lost 36.4 % lost 60.9 % lost 80.8 % lost Eq.49

T = 0.1 T =1 T = 10 T = 100
1.0 0.5
0.5
0.8 0.6 0.4
0.4
ρ(k, t)

0.6 0.4 0.3 0.3


0.4 0.2 0.2
0.2
0.2 0.1 0.1
0.0 0.0 0.0 0.0
−π 0 π −π 0 π −π 0 π −π 0 π
Rapidity k Rapidity k Rapidity k Rapidity k

Figure 4. Effect of two-body losses on the rapidity distribution. The initial rapidity distribution ρ(k, 0) = 1/(1 +
exp(− cos(k)/T )) is chosen with T = 0.1, 1, 10, 100 from left to right. The black dashed curve is the analytic solution (see
Eq. (49)). After several loss events the rapidity distribution takes the form of a gaussian centered at k = 0. As in the one-body
case, if initially the rapidity distribution is flat then it remains flat under lossy evolution.

B. Results for K = 2 and in Fig. 5 we show the corresponding evolution of the


mean atom density n(t). It appears that, except for an
We now consider the time evolution equation for the initial infinite temperature√state, the mean density al-
rapidity distribution for the K = 2 case, which is charac- ways decays as n(t) ∝ 1/ t at very long times, while
terized by the functional given by Eq. (31). For simplic- the density decreases as 1/t for an infinite temperature.
ity, here we focus on initial rapidity distributions ρ0 (k) These two behaviors follow from Eqs. (48)-(49), as we
that are symmetric under reflection k → −k. Since the now explain.
master equation is also invariant under k → −k, this Eq. (48) reveals that initial rapidity
´ π distributions that
property is conserved throughout the entire evolution. have a vanishing first Fourier mode −π dq cos(q)ρ(q) al-
Then Eq. (31) simplifies to the following expression, ways follow the exact same dynamics as the mean density,
ˆ π namely
1 1
F [ρ] = 2 ρ(k) n(t) − cos (k)ρ(k) dq cos (q) ρ(q). n(t) = , (50)
π −π 1 + 2n(0)Γt
(48)
Eq. (48) highlights the two distinct contributions to the characterized by a long-term decay as ∼ 1/t. This power
time evolution of ρ(k, t): the first term in the right- law is then always found for initial rapidity distributions
hand side represents a mean-field contribution, as it with vanishing first Fourier mode, including infinite tem-
does not introduce any structure in rapidity space, while perature states.
the second term is responsible for generating quantum In contrast, initial rapidity distributions that have √
correlations and, consequently, introducing structure in a non-vanishing first Fourier mode decay as ∼ 1/ t
k−space. at long times. This can be understood by looking at
After some algebra presented in App. B, one can derive the long time limit of Eq. (49). Let us introduce the
´t
an exact (implicit) expression for the rapidity distribu- two time-dependent functions g(t) = 0 n(τ ) dτ and
tion at all times: ´t q
∂τ n(τ )
f (t) = 0 1 + 2Γ n(τ )2 n(τ )dτ . Numerically we ob-
ρ(k, t) = ρ0 (k) serve that |∂τ n(τ )| ≪ 2Γn(τ )2 at long times, as soon
(
t
s ! ) as the initial rapidity distributions has a non-vaishing
∂τ n(τ )
ˆ
× exp −2Γ 1 − σ0 cos (k) 1+ n(τ )dτ , first Fourier mode. Then, expanding at first order in
0 2Γ n(τ )2 |∂τ n(τ )|/(2Γn(τ )2 ), f (t) becomes
(49) 1

n(t)

´π f (t) ≃ g(t) + ln , (51)
where σ0 = sgn( −π cos (k)ρ0 (k)dk), with sgn(x) = ±1 4Γ n(0)
the sign function. Eq. (49) shows that ρ(k, t) is entirely implying that, at large t, the difference between f (t) and
determined by ρ0 (k) and n(t). Notably, it reveals that g(t) grows as ln(t). Integrating Eq. (49) over k leads to
rapidities are distributed according to a cosine law, with the mean density
the k = 0 mode that has the longest lifetime.
e−2Γg(t) π
ˆ
In Fig. 4 we show the evolution of the rapidity distribu- n(t) = dk ρ0 (k) e2Γσ0 cos (k)f (t) . (52)
tion for initial thermal states at different temperatures, 2π −π
9

8
100 a)

ρ(k, t)/n(t)
t=0
6 Γt = 96
4 Γt = 3.24 × 105
10−1 2
Fit with a1 ea2 cos(k)
n(t)/n(0)

0
10−2 −π − π2 0 π
2 π
T = 0.1
T = 0.5 b)

ρ(k, t)/n(t)
t=0
T =1
4 Γt = 12
10−3 T = 100 Γt = 3.24 × 105
Eq.(52) 2 Fit with a1 ea2 cos(k)
∼ t−1/2
10−4
0
−π − π2 0 π
π
10−2 10−1 100 101 102 103 104 2

Γt 2 c)

ρ(k, t)/n(t)
t=0
Γt = 3.24 × 105

Figure 5. The mean density under two-body losses for differ- 1


ent thermal-like initial rapidity distributions. The chemical
potential is µ=0.0 so that the initial density is n(0) =0.5.
0 π
Colored curves are obtained by solving numerically the time −π − π2 0 2 π
evolution equation of ρ(k) for the loss functional (31) with a
Rapidity k
time step dt=0.05 and a loss rate Γ =0.1. The red dashed line
is the mean density associated to an initial rapidity distribu-
tion which is flat as shown in (57). The long time behavior at Figure 6. Long time behavior of different rapidity distributions
rescaled by the corresponding density under two-body losses. For
small temperatures is presented by the black dashed line (see
the subplots a, b and c, the blue curve represents the initial ra-
Eq.(55)). pidity distribution while the red curve is the rapidity distribution
at long time. The violet curve is the rescaled rapidity distribution
at an intermediate time. a) The initial distribution is a Fermi-
Since the function f (t) diverges at large t, the latter inte- Dirac distributions ρ0 (k) = (1 + exp(− cos(k)/T ))−1 with T =0.1.
gral can be evaluated by the saddle-point approximation; The green dashed curve is a fit with a Boltzmann distribution
a1 exp{a2 cos(k)} where a2 is positive. b) The initial rapidity dis-
we denote kσ∗0 the saddle point: kσ∗0 = 0 if σ0 = +1 and tribution is (1 − cos(k))/2 which has a non-vanishing first Fourier
kσ∗0 = π if σ0 = −1. We are thus left with: mode. The green dashed curve is a fit with a Boltzmann distribu-
ˆ π tion a1 exp{a2 cos(k)} where a2 is negative. c) The initial rapididty
distribution is (1 − cos(2k))/2 which has no first Fourier mode.
dk ρ0 (k) e2Γσ0 cos (k)f (t)
−π
ˆ ∞
2
2Γ f (t)
≃ e dk ρ0 (k) e−Γ k f (t)
t→∞ −∞
We conclude this subsection with an investigation
r
π of the long time behavior of the rapidity distribution
= ρ0 (kσ∗0 ) e2Γ f (t) . (53) ρ(k, t), which is determined by the mean density n(t) ac-
Γf (t) cording to Eq. (49). We have just we etablished that
the first Fourier mode of the initial rapidity distribu-
Using Eq. (51), we find
tion strongly influences the long time behavior. In the
1 ρ0 (kσ∗0 )2 1 ρ0 (kσ∗0 )2 case of a vanishing first Fourier mode, the rapidity den-
n(t) ≃ ≃ , (54) sity at time t is simply given (see Eqs. (50)-(49)) by
t→∞ Γf (t) 4πn(0) Γg(t) 4πn(0)
ρ(k, t) = ρ0 (k) n(t)/n(0). The ratio ρ(k, t)/n(t) is then
where, in the second identity, we have used the fact time-independent, as illustrated in Fig. 6(c). In contrast,
that the logarithmic term in Eq. (51) is subleading. when the first Fourier mode of the initial distribution
Since n(t) = ∂t g(t), we arrive at an ordinary differen- ρ0 (k) is non-zero, the ratio ρ(k, t)/n(t) loses its depen-
tial equation of the form ∂t g(t) ∝ 1/g(t). Consequently, dence on the initial rapidity distribution at very long

g(t) ∝ t, and then times. Indeed, in that case the rapidity distribution goes
to a low-density, low-temperature, Boltzmann distribu-
n(t) ∝ t−1/2 , (55) tion of the form ρ(k, t)/n(t) ≃ eβ(t) cos k with the effective
inverse temperature β(t) = 2σ0 Γf (t). This is illustrated
as expected from our numerical results, see Fig. 5. We for the case of an initial thermal rapidity distribution
note that a similar result was found recently in a lattice in Fig. 6(a), where we see that the ration ρ(k, t)/n(t)
gas with similar but different two-body loss term [25] as gets concentrated around k = 0 and is very close to a
well as in its continuous analog [43], although we stress Boltzmann distribution. Notice also that the effective
that the loss functionals and rate equations for these temperature β(t) is negative when the sign of the first
models are different from the ones of this paper. Fourier mode of the initial rapidity distribution is neg-
10

0.0 % lost 16.2 % lost 34.4 % lost 65.1 % lost 86.5 % lost
4.6 % lost 23.6 % lost 43.4 % lost 73.5 % lost 92.0 % lost
9.5 % lost 30.2 % lost 52.5 % lost 81.2 % lost 95.9 % lost

T = 0.1 T =1 T = 10 T = 100
1.0 0.5
0.5
0.8 0.6 0.4
0.4
ρ(k, t)

0.6 0.4 0.3 0.3


0.4 0.2 0.2
0.2
0.2 0.1 0.1
0.0 0.0 0.0 0.0
−π 0 π −π 0 π −π 0 π −π 0 π
Rapidity k Rapidity k Rapidity k Rapidity k

Figure 7. Effect of three-body losses on the rapidity distribution. The initial rapidity distribution ρ(k, 0) = 1/(1 +
exp(− cos(k)/T )) is chosen with T = 0.10, 1.0, 100.0 from top left to bottom right.

ative. This is illustrated in Fig. 6(b), where we display


the ratio ρ(k, t)/n(t) at late time for the far-from-thermal 100
initial rapidity distribution ρ0 (k) = (1 − cos(k))/2. We
observe that, at late times, the distribution gets concen-
trated around k = π and corresponds to a Boltzmann
distribution at negative temperature.
Remarkably, these observations are in stark contrast 10−1
with our findings for the K = 1 case. While we found
n(t)/n(0)

that, for K = 1, the rapidity distribution never goes


0.5(1 − cos(k))
to a thermal distribution at late time, here for K = 2
0.5(1 − cos(2k))
the distribution goes to a low-density, low-temperature
(possibly negative), thermal distribution. This is always 10−2 T = 0.1
T = 0.5
true, except in the special case where the first Fourier T = 1.0
mode of the rapidity distribution vanishes; in that case T = 2.0
the rapididity distribution is simply rescaled by a factor T = 100
n(t)/n(0) under lossy evolution. ∼ (Γt)−1/2
10−3
∼ (Γt)−0.38
∼ (Γt)−0.21
C. Generic observations for arbitrary K 10−1 100 101 102 103 104 105 106
Γt
We now turn to the case of higher K, and draw some
general conclusions.
Figure 8. The mean density under three-body losses for different
Numerically, we solve the time evolution equation of initial rapidity distributions. The chemical potential is µ=0.0 so
the rapidity distribution for three-body losses (K = 3), that the initial density is n(0) =0.5. Colored curves are obtained
see Fig. 7 for the evolution of the rapidity distribution by solving numerically the time evolution equation of ρ(k) with
from an initial thermal state, and Fig. 8 for the atom den- an non-regular time step and a loss rate Γ =0.1. From blue to
red, the simulation is performed with an initial distribution which
sity n(t). In Fig. 7 we see that the effect of three-body is a Fermi-Dirac distribution ρ0 (k) = (1 + exp(− cos(k)/T ))−1 .
losses is to spread the rapidity distribution in rapidity The dark and light green curves are respectively obtained from
space, as already observed for one-body and two-body initial rapidity distributions (1 − cos(k))/2 and (1 − cos(2k))/2.
losses. We expect that this is a generic effect caused by The black dashed line is the long time behavior of the mean density
predicted by Eq. (50). The orange dashed curve corresponds to the
K-body losses for any K. In Fig. 8, we observe that
long time behavior of the mean density computed from an initial
the mean density decays as t−1/2 for an initial infinite rapidity distribution with non-vanishing first Fourier mode. The
temperature state, while for any non-zero initial tem- gray dotted curve shows the long time behavior of the mean density
perature it crosses over to a t−α decay at long times for an initial rapidity distribution with no first Fourier mode.
with an exponent α ≃ 0.21. This exponent seems to
be independent of the initial temperature as long as it is
non-zero, see Fig. 8. However, for an initial rapidity dis- that the density also decays as a power-law at late time,
tribution that is far from thermal, such as for instance although with a different exponent α: the exponent is
ρ(k, t = 0) = (1 − cos k)/2 or (1 − cos(2k))/2, we find close to 0.21 for ρ(k, t = 0) = (1 − cos k)/2, and close to
11

ρ(k, t)/n(t) 10 a) mode of ρ(k) is tuned to zero) goes to a ,low-density, low-


t=0
Γt = 2.4 × 107
temperature thermal state for K = 2, while for K = 1 it
never does. In Fig. 9 we display this ratio at late time for
5 K = 3. Fig. 9.(a) corresponds to a thermal initial rapid-
ity distribution, and Figs. 9.(b)-(c) to non-thermal ini-
0 tial rapidity distributions ρ(k, t = 0) = (1 − cos k)/2 and
−π − π2 0 π
2 π
(1−cos(2k))/2 respectively. We observe that the rescaled
b) rapidity distribution concentrates around the maxima of
ρ(k, t)/n(t)

t=0
10 Γt = 2.4 × 107 the initial rapidity distribution at long times. Even for
an initial thermal distribution (Fig. 9 (a)), the long time
5 behavior of the density profile can not be described by
a Boltzmann distribution, as it looks like a bell-shaped
0
−π − π2 0 π
π distribution that has a small dip at k = 0. A similar
2
conclusion holds for Fig. 9.(b). Finally Fig. 9.(c) shows
15 c) the emergence of peaks localised at k = ±π/2. We con-
ρ(k, t)/n(t)

t=0
Γt = 2.4 × 107 clude that, in contrast with the K = 2 case, the late-time
10
rapidity distribution is generically non-thermal.
5
0
−π − π2 0 π
2 π
V. HARMONICALLY TRAPPED GAS
Rapidity k

Figure 9. Long time behavior of different rapidity distributions In many cold atom experiments, the gas lies in a longi-
rescaled by the corresponding density under three-body losses. For tudinal trapping potential. This prompts us to study the
the subplots a, b and c, the blue curve represents the initial rapid- influence of the trapping potential on the dynamics of
ity distribution while the red curve is the rapidity distribution at our lossy lattice hard-core gas. For simplicity we restrict
long time. a) The initial distribution is a Fermi-Dirac distributions
ρ0 (k) = (1 + exp(− cos(k)/T ))−1 with T =0.1. b) The initial ra- ω 2 x2
to a harmonic potential V (x) = .
pidity distribution is (1 − cos(k))/2 which has a non-vanishing first 2
Fourier mode. c) The initial rapididty distribution is (1−cos(2k))/2 We adopt a coarse-grained perspective of the gas:
which has no first Fourier mode. we assume that the gas can be divided into fluid cells
which contain a large number of bosons, and that the
state of the gas within each fluid cell [x, x + dx] is
0.38 for ρ(k, t = 0) = (1 − cos(2k))/2.
a certain macrostate represented by the local density
We have not been able to analytically derive the ob-
of rapidities ρ(x, k). Such coarse-grained descriptions
served generic power-law decay for K = 3 or for higher
have been very successful lately in describing the out-
K, beyond the special case of the initial infinite temper-
of-equilibrium quantum many-body dynamics of nearly
ature state. The latter case is easily understood because,
integrable gases [52, 55–57, 62–64]. Here we investigate
for an infinite temperature state the rapidity distribu-
the effect of losses on our lattice hard-core gas within
tion is constant, ρ(k) = n, and the equation (15) can be
that coarse-grained description.
solved analytically. Then the determinant of the matrix
B is equal to the determinant of C, and the matrices A The equation satisfied by the position-dependent ra-
reduce to identical and diagonal matrices. Therefore the pidity distribution is
loss functional is simply given by
∂t ρ(x, k, t) + sin(k) ∂x ρ(x, k, t) − ω 2 x ∂k ρ(x, k, t)
K
FK [ρ](k) = Kn , (56) = −ΓF [ρ(x, ., t)](k). (58)
which is the result expected from the mean-field ap-
In the first line, the term ∂x sin(k)ρ(x, k) corresponds to
proach. Then the solution of the evolution equation (15)
gives the mean density the gradient of the current of quasi-particles with rapid-
ity k, j(x, k) = sin(k)ρ(x, k). Here sin(k) is the group
n(0) velocity of quasi-particles with lattice dispersion relation
n(t) = . (57) ε(k) = − cos(k). The term −ω 2 x ∂k ρ(x, k) in Eq. (58)
(1 + n(0)K−1 K(K − 1) Γt)1/(K−1)
corresponds to Newton’s second law, and encodes the
Beyond that simple case, we have not been able to express fact that the quasi-particles feel the harmonic potential
the loss functional in a simple form so as to derive the and are accelerated according to k̇ = −∂x V (x) = −ω 2 x.
long-time decay of the mean density. Finally, the r.h.s of Eq. (58) is the loss term at position x,
Similarly to the K = 1 and K = 2 cases, we have which follows from the assumption that the gas is locally
investigated the behavior of the rescaled rapidity distri- homogeneous so that we can apply the formalism devel-
bution ρ(k, t)/n(t) at late times. Recall that this ratio re- oped in previous sections, this time within each fluid cell
veals that the gas generically (i.e. unless the first Fourier [x, x + dx].
12

K=1 K=1 K=1 K=1 K=1 K=1 K=1 K=1 K=1 K=1
π 0τ 1τ 2τ 3τ 4τ 5τ 6τ 7τ 8τ 9τ
ρ(x, k)
0
k

0 % lost 10 % lost 20 % lost 30 % lost 40 % lost 50 % lost 60 % lost 70 % lost 80 % lost 90 % lost
−π
1.0
n(x, t)/n0

0.5 0.8

0.0 K=2 K=2 K=2 K=2 K=2 K=2 K=2 K=2 K=2 K=2
π 0τ 12 τ 24 τ 42 τ 66 τ 102 τ 150 τ 228 τ 390 τ 810 τ
0.6
0
k

0 % lost 10 % lost 20 % lost 30 % lost 40 % lost 50 % lost 60 % lost 70 % lost 80 % lost 90 % lost
−π
1.0
n(x, t)/n0

0.4
0.5

0.0 K=3 K=3 K=3 K=3 K=3 K=3 K=3 K=3 K=3 K=3
π 0τ 66 τ 154 τ 286 τ 484 τ 836 τ 1584 τ 3300 τ 16000 τ 38400 τ 0.2
0
k

0 % lost 10 % lost 20 % lost 30 % lost 40 % lost 50 % lost 60 % lost 70 % lost 80 % lost 90 % lost
−π
1.0 0.0
n(x, t)/n0

0.5

0.0
−3 0 3 −3 0 3 −3 0 3 −3 0 3 −3 0 3 −3 0 3 −3 0 3 −3 0 3 −3 0 3 −3 0 3
x x x x x x x x x x

Figure 10. Evolution of the position-dependent distribution ρ(x, k) in a harmonic trap and under K-body losses. The initial
distribution is 1/(1 + exp (− cos(k) + 0.5ω 2 x2 )/T with the harmonic trap frequency ω=1/2 and the temperature T =0.1. The

loss rate is Γ=0.1 and the time step τ = 0.125/ω. The x and k-axis are discretized with 300 points on each axis. Below each
phase portraits we add the corresponding density profiles n(x, t) obtained by integrating ρ(x, k, t) over k. The density profiles
are rescaled by the initial density at the trap center n0 = n(x = 0, t = 0) = 0.5. The two first rows show the effects of one-body
losses (K=1) on the distribution in phase-space and the corresponding density profiles. The two next rows show the effects of
two-body losses (K=2) and the last two rows show the effect of three-body losses (K=3).

A. Numerical method where y(τ, k) is the solution of the differential equation


∂τ y(τ, k) = −ΓF [y(τ, .)](k), y(0, k) = ρ′t+∆t,x . (61)
Our main goal in this section is to solve numerically
the evolution equation (58). For this we use a split- The rapidity distribution ρ(x, k) is discretized on a reg-
step method. Assuming that we know the rapidity dis- ular grid in phase space, and the two steps in Eq. (59)
tribution ρt (x, k), from time t to t + ∆t we first com- and Eq. (60) are implemented as follows.
pute the new rapidity distribution ρ′t+∆t (x, k) generated For the transport step in (59), we use the method of
by the transport of quasi-particles, and then compute characteristics, which here is very simple since the un-
ρt+∆t (x, k) from ρ′t+∆t (x, k) by implementing localized derlying dynamics is the one of non-interacting quasi-
lossy evolution during a time step ∆t. This gives the particles. Each quasi-particle at position (x, k) in phase
following scheme, with first step space evolves according to:
 dx
 dt = sin(k)

ρ′t+∆t (x, k) = ρt (x, k) − ∆t sin(k)∂x ρt (x, k)

(62)
+ ∆t ω 2 x ∂k ρt (x, k), (59)  dk = −ω 2 x.


dt
and second step We start from the values ρt (x, k) on the regular grid in
phase space, and we move each node of that grid accord-
ρt+∆t (x, k) = y(∆t, k) (60) ing to x → x + ∆t sin k and k → k − ∆t ω 2 x. This gives
13

us the new rapidity distribution after transport over a 100


time ∆t, which however is no longer defined on the ini-
tial regular grid. To get the new rapidity distribution a) K = 2
ρ′t+∆t (x, k) on the initial regular grid, we use linear in-

N (t)/N (0)
terpolation. As a benchmark, we have checked that this
method gives excellent numerical precision for the simu-
lated transport in the absence of losses.
10−1
The second step consists in solving numerically the dif- ω = 0.5, µ = −0.77
ferential equation (61) for each column of the grid in ω = 1.0, µ = −0.52
phase-space. The solution of each differential equation ω = 2.0, µ = −0.07
is obtained by the Runge-Kutta method with the initial ∼ 1/t
condition y(0, k) = ρ′t+∆t,x . We do this for each column
of the grid, and we thus get the new space-dependent 100 101 102
rapidity distribution ρt+∆t (x, k) according to Eq. (60). 100
The combination of the two steps allows us to go from
a rapidity distribution ρt (x, k) to the new distribution
ρt+∆t (x, k), both defined on the same phase-space grid. b) K = 3

N (t)/N (0)
We then repeat this procedure many times with a small
time step ∆t to simulate the lossy evolution of the gas in
the trap. ω = 0.5, µ = −0.77
ω = 1.0, µ = −0.52
ω = 2.0, µ = −0.07

B. Results ∼ 1/ t
∼ 1/t0.6
10−1
We have performed numerical simulations of the evo- 100 101 102
lution of the position-dependent rapidity distribution Γt
ρ(x, k) under K-body losses using the algorithm pre-
sented in the previous section, see Fig. 10. For the initial Figure 11. (a) The time evolution of the total particle
state, we use a thermal (Fermi-Dirac) rapidity distribu- number under two-body losses. The initial distribution is
tion ρ(x, k) = 1/(1 + exp (− cos(k) + ω 2 x2 /2 − µ)/T ), 1/(1 + exp (− cos(k) + 0.5ω 2 x2 − µ)/T , where T = 0.1 and

where µ is the chemical potential and T the temperature. µ, the chemical potential, is chosen to fix the initial mean
Our numerical study allows us to make the following particle density to 0.5 and N (0) represents the initial num-
general observations, illustrated in Fig. 10. For all K ber of particle. The loss rate is Γ=0.1 and the time step
dt=1.5. The red dashed line is the long time behavior of
(we have simulated K = 1, 2, 3) the distribution typi-
the total particle number for the homogeneous case with an
cally spreads in phase-space, similarly to the homoge- initial rapidity distribution which has no first Fourier mode
neous case. However, while the number of particles de- (see Fig.(4)). (b) The time evolution of the total particle
cays exponentially for K = 1, the loss dynamics is much number under three-body losses. The initial distribution is
slower for higher K, and this has visible effects on the 1/(1 + exp (− cos(k) + 0.5ω 2 x2 − µ)/T , where T = 0.1 and
distribution of rapidities after a given percentage of lost µ, the chemical potential,is chosen to fix the initial mean par-
atoms. ticle density to 0.5 and N (0) represents the initial number of
For K = 1 we observe that the edges of the phase- particle. The loss rate is Γ=0.1 and the time step dt=1.5.
space distribution get depopulated very fast and quickly The black dashed line represents the asymptotic behavior of
results in a halo around the origin. The spreading is also the mean density and the red dashed line is the mean-field
prediction.
clearly visible
´ in real space in the particle density profile
n(x) = ρ(x, k)dk/(2π), see the second line of Fig. 10.
For K = 2, the situation is a little bit different, see
Fig. 10. Until ∼20% of the atoms have been lost, the For K = 3, we also observe a small spiral appearing
dynamics is similar to the one for K = 1, but after that at the center of the distribution after ∼30% of the atoms
we observe the formation of spirals in the bulk of the have been lost. This spiral remains localised at the center
phase-space distribution. The spirals become more visi- during the dynamics. Interestingly, as one can see from
ble as the ratio Γ/ω is increased, see Appendix D. Com- the fifth line in Fig. 10, this time it is the center of the
pared to K = 1, the bulk of the distribution also gets phase-space distribution that decreases faster compared
depopulated, leading to an approximately uniform cir- to the edges. After ∼60% of atoms lost, a hole starts
cular droplet in phase-space, with a density that decays developing at the center of the phase-space distribution,
with time. This is visible in Fig. 10 after ∼50% of the and the latter looks more and more like a ring. This is
atoms have been lost; we note that, compared to the case a clear signature of a strongly out-of-equilibrium gas in
of one-body losses, the distribution spreads less signifi- the trap, with a population inversion: the higher-energy
cantly in phase-space. single-particle orbitals get more populated than the low-
14

energy ones. This effect is reflected in the corresponding not drive the gas to a low-density thermal equilibrium
real-space particle density n(x) then acquires a doubly- state at long times. In the case of two-body losses, our
peaked shape, with a local minimum at x = 0, similarly formula (49) gives an implicit expression for the rapidity
to what we observed also in the homogeneous case. distribution and using a similar method as in Ref. [25],
Finally, having simulated the dynamics of the position- we were able to investigate the long time behavior of the
dependent rapidity distribution ρ(x, k), one can easily rapidity distribution and the mean particle density. √In
get the evolution of the total particle number N (t) by particular, we found that it decays generically as ∼ 1/ t,
integrating ρ(x, k, t) over x and k at fixed time t. For except when the first Fourier mode of the initial distri-
one-body losses, one always finds an exponential decay bution vanishes; in that case the particle density decays
N (t) = e−Γt N (0). For two-body and three-body losses as ∼ 1/t. A similar conclusion was drawn for a different
the result is more interesting. In Fig. 11.(a), we show the loss process in Refs.[25, 43].
evolution of the mean density under two-body losses for Finally, we considered the inhomogeneous system con-
different trap frequencies ω. Starting with a thermal dis- sisting in a lattice hardcore bosons gas in harmonic po-
tribution at temperature T and chemical potential µ, we tential. We provided a numerical method to solve the
observe that the mean density decreases at long times as dynamics combining the effects of the losses and of the
∼ 1/t, a result that coincides with the mean-field (or infi- trapping potential. We observed that for K ≥ 2 the trap
nite temperature) decay for the homogeneous gas. More- generically speeds up the decay of the total particle num-
over, we see that the higher the trap frequency the faster ber. We also found that the gas typically evolves towards
the decay of the total particle number. a highly non-thermal state; in particular for K = 3 we ob-
For three-body losses, we see in Fig. 11.(b) that, like in serve a striking ring-shaped distribution in phase space,
the two-body case, a stronger confinement speeds up the which signals a inversion of population (i.e. higher en-
decrease of total particle number. However, this time we ergy single-particle orbitals get more populated than the
do not observe a clear convergence towards the expecta- lower energy ones), see Fig. 10. Further investigations
tion from the mean-field (or infinite temperature) result are needed to draw more quantitative conclusions about
for the homogeneous case, which would be N (t) ∼ t−1/2 . the dynamics in the trap.
We observe a number of particles that decreases approx-
imately as a power-law ∼ t−α with an exponent α ≃ 0.6. Note: while we were finishing this paper, a preprint
Our numerics does not allow us to draw a clear conclu- by Perfetto, Carollo, Garrahan and Lesanovsky appeared
sion as to whether or not this would go to the mean-field [65], where a similar model of spinless fermions with
exponent 1/2 at longer times. Nevertheless, let us stress K = 2, 3, 4-body losses is studied within the context of
that, qualitatively, the effect of three-body losses is the quantum reaction-diffusion dynamics of annihilation pro-
same as in the two-body case: compared to the homoge- cesses. Our hard-core model coincides with their model
neous case, the trap dramatically speeds up the losses. for K even, but not for K odd because of the Jordan-
Wigner mapping, as we explained in detail in Sec. III.
Perfetto et al. do not focus on the effect of losses on
VI. CONCLUSION the rapidity distributions of the gas, but rather on the
evolution of the number of particles in the homogeneous
We have studied the effects of K-body losses on a gas setting. For the case of K even, where our models co-
of lattice hardcore bosons, in particular their effect on incide, their findings about the evolution of the number
the thermalisation of the gas at late time. For this, we of particles in the homogeneous setting are in agreement
have relied on the hypothesis of adiabatic losses used pre- with ours.
viously in Refs. [25, 40, 58–60]. We derived analytical re-
sults for the loss functional for any integer K in the form
of a small finite determinant, and closed expressions in ACKNOWLEDGMENTS
the cases K = 1, 2.
For K = 1 and K = 2, we solved analytically the time We thank Alberto Biella, Isabelle Bouchoule, Mario
evolution equation of the rapidity distribution of the spa- Collura and Leonardo Mazza for very useful discus-
tially homogeneous gas. In the case of one-body losses, sions and for joint work on closely related topics. The
our formula (45) shows that the loss functional is in gen- work of JD, FR and DK is supported by the Agence
eral non-linear and non-local in rapidity space, as already Nationale de la Recherche through ANR-20-CE30-0017-
observed for the continuous Lieb-Liniger gas in Ref. [40]. 01 project ‘QUADY’ and ANR-22-CE30-0004-01 project
After investigating the long time behavior of the rapid- ‘UNIOPEN’. L.R. acknowledges hospitality from LPCT
ity distribution, we concluded that one-body losses do during the completion of this work.
15

Appendix A: Deriving solution (45)

For one-body losses, the time evolution of the rapidity distribution ρ(t, k) is given by
∂t ρ = −Γ ρ − (ρ2 − H(ρ)2 − n2 (t)) + 2n(t)H′ (ρ) ,

(A1)
´π
where Γ is the loss rate. We introduced the Hilbert transform H(f (x)) = 2π 1
dy tanf (y) with f (x) a periodic
−π
2 )
( x−y
−Γt
function. The mean density n(t) is known: n(t) = n0 e .
Here the rapidity distribution is a 2π-periodic real-valued function. From the rapidity distribution, we can construct
a complex-valued function whose the imaginary part is the Hilbert transform of the real part: Q = ρ(k) + iH(ρ(k)).
Such a function is called an analytic signal and can be analytically continued to the upper half-plane: the function
´ π dq
Q(z) = 2π −π tan ρ(q)
i
is well-defined for Im(z) > 0 and Re(z) ∈ [−π, π] and reduces to ρ(k) on the real axis.
2 )
( z−q
Taking the Hilbert transform of (A1)
∂t H(ρ) = −Γ H(ρ) − H(ρ2 − H(ρ)2 ) + 2n(t)H(H′ (ρ))

(A2)
and adding (A1) and i (A2), one has
∂τ Q(τ, z) = −(Q(τ, z) − i2n∂z Q(τ, z) − Q2 (τ, z) + n2 (τ )). (A3)
We used some properties of the Hilbert transform: i) H(H(f )) = −f , ii) H′ (f ) = H(f ′ ). Moreover, since Q2 (z) is
analytic for Im(z) > 0, the function ρ2 − H(ρ)2 + i2ρH(ρ) is an analytic signal if and only if H(ρ2 − H(ρ)2 ) = 2ρH(ρ).
Introducing the function Y (τ, z) = Q(τ, z + i2n(τ )), one gets
∂τ Y (τ, z) = Y 2 (τ, z) − Y (τ, z) − n2 (τ ) (A4)
This equation can be solved if one assumes Y (τ, z) = α(τ, z) e−τ . Indeed, thanks to this trick the above equation
reduces to
∂τ α(τ, z) = (α2 (τ, z) − n20 ) e−τ (A5)
Putting all terms depending on α in the left-hand side, one has

ˆ
= −e−τ + C1 , (A6)
α − n20
2

which leads to
α(τ, z) = n0 tanh n0 e−τ + C2 .

(A7)

The initial condition Y (0, z) = Y0 sets the constant: C2 = tanh−1 (Y0 /n0 ) − n0 . Thus, one can write
Y (τ, z) = n(τ ) tanh n0 (e−τ − 1) + tanh−1 (Y0 /n0 )


tanh(n0 (e−τ − 1)) + Y0 /n0


 
= n(τ ) . (A8)
1 + tanh(n0 (e−τ − 1))Y0 /n0
Finally, the rapidity distribution reads
 
−Γt
 i
´π ρ0 (q)
 tanh n0 (e − 1) + 2πn0 −π dq   
 tan k−q
2 + in0 (1 − e−Γt ) 
ρ(t, k) = n0 e−Γt Re  . (A9)
 
i −Γt − 1)) π dq
´ ρ 0 (q)
 1 + 2πn tanh(n (e
 
0
0 −π
  
tan k−q
2 + in 0 (1 − e −Γt )

Appendix B: Derivation of Eq. (49) of the main text

In this section we present the main steps to derive the exact expression of the rapidity distribution in the homoge-
neous case for K = 2. We consider the time evolution of n(t) written as:
ˆ π
1
∂t n(t) = ∂t ρ(k, t)dk. (B1)
2π −π
16

We now insert the evolution equation (48) obtaining:


2
2Γn(t) π
ˆ π
Γ
ˆ
∂t n(t) = − ρ(k, t)dk + 2 cos (k)ρ(k, t)dk
2π −π 2π −π
ˆ π 2
2 Γ
= − 2Γn(t) + 2 cos (k)ρ(k, t)dk (B2)
2π −π
By inverting the latter relation, one obtains:
ˆ π r
2
cos (k)ρ(k, t)dk = π
(∂t n(t) + 2Γn(t)2 ), (B3)

−π Γ
where |•| denotes
´π the absolute value. At this point of the derivation it is useful to introduce the following variable:
σ(t) = sgn( −π cos (k)ρ(k, t)dk), with sgn(x) being the sign function. We now claim that the function σ(t) is solely
determined by its value at initial time, i.e. σ(t) = σ(0) + σ0 , the argument goes as follows. (i) If the first Fourier
mode vanishes at some time t, then it must vanish also at any later time. This follows from Eq. (48). (ii) This implies
that the sign of the first Fourier mode is continuous in time. Since it can take only discrete values, it is in fact a
constant.
By inserting the latter equation in Eq. (48) the time evolution of the rapidity distribution can be then recasted into
the following form:
r
2
ρ̇(k, t) = −2Γ ρ(k) n(t) + Γσ0 cos (k)ρ(k) (∂t n(t) + 2Γn(t)2 ). (B4)
Γ
We now divide both sides by ρ(k, t) and then integrate:
  ˆ t ˆ tr
ρ(k, t) ′ ′ 2
ln = −2Γ n(t )dt + Γσ0 cos (k) (∂t′ n(t′ ) + 2Γn(t′ )2 )dt′ . (B5)
ρ0 (k) 0 0 Γ
By exponentiating the latter equation we get:
 s 
ˆ t ˆ t ′ ′ 2
 
 2 ∂t′ n(t ) + 2Γn(t )
ρ(k, t) = ρ0 (k) exp −2Γ n(t′ )dt′ + σ0 cos (k) dt′ , (B6)
 0 0 Γ 

this concludes the derivation of Eq. (49).

Appendix C: Additional data for the homogeneous K = 2 case

100
1.5 ρ0 (k, T = 0.1) = 1
1+ecos (k)/T

ρ0 (k) = 21 (1 − cos (k)))


10−1 ∼ log(t)
f (t) − g(t)
n(t)/n(0)

1.0
10−2

ρ0 (k) = 12 (1 − cos (k)) 0.5


10−3 ρ0 (k) = 21 (1 − cos (5k))
Eq.(51)
∼ t−1/2
10 −4
0.0
0 1 2 3 4
10 −2
10 −1
10 10 10 10 10 10−2 10−1 100 101 102 103 104
Γt Γt

Figure 12. Left panel: the mean density under two-body losses for two different non-thermal rapidity distributions. Solid
colored curves are obtained by solving numerically the time evolution equation of ρ(k) for the loss functional (31) with a time
step dt=0.05 and a loss rate Γ =0.1. The red dashed line is the mean density given by Eq. (50). The long time behavior for non
vanishing first Fourier mode distributions is presented by the black dashed line (see Eq.(??)). Right panel: difference between
the two function f (t) and g(t) defined in the main text for a thermal and non-thermal distributions.
17

In this section we present additional data concerning the homogeneous K = 2 case. In particular, we show in
Fig. 12 (left panel) the dynamics of the mean density for two different initial rapidity distributions which are not
1
thermal. Firstly, we consider an initial distribution given by ρ0 (k) = (1 − cos (k)), which has first Fourier mode
2
1
different from zero. Secondly, we consider ρ0 (k) = (1 − cos (5k)), whose first Fourier mode vanishes. We see that
2 √
the dynamics induced by the former distribution has a longtime behaviour guven by ∼ 1/ t, whereas the latter,
due to its vanishining first Fourier mode, is described by Eq. (50). This corroborates our findings for initial thermal
distributions presented in the main text. Moreover, we show in Fig. 12 (right panel) the quantity f (t) − g(t), where
´t ´tq ∂τ n(τ )
g(t) = 0 n(τ ) dτ and f (t) = 0 1 + 2Γ n(τ )2 n(τ )dτ for two different rapidity distributions. In the main text we
took the first order expansion of f (t) resulting in a logarthmic growth for the quantity f (t) − g(t), which is thus
corroborated by the numerical data here presented. As such, given an initial rapidity distribution
√ whose first Fourier
mode is non-vanishing, one has a longtime decay of the mean density given by n(t) ∼ 1/ t.

Appendix D: Brief discussion on the spiral in figure Fig. 10

During the evolution of the position-dependent rapidity distribution in phase space (see Fig. 10) , the distribution
exhibits a spiral, which is clearly visible after 40% of atoms lost for two-body losses. In principle, in a regime where
the trap frequency ω is highly dominating the loss rate Γ, one expects that the distribution remains rotation invariant
at all times. This implies that the spiral vanishes for ω ≫ Γ. To check this statement we compare the quantity ρ(x, k)
for two distincts values of ω (see the figure below). In figure Fig. 13 We can see that the spiral appearing for ω = 5Γ
covers entirely the distribution, while for ω = 20Γ the spiral is localised at the distribution’s center. Moreover, in the
case ω = 20Γ we observe small oscillations between the edges and the center of the distribution. The frequency of
these oscillations is high compared to the center of the distribution.

ρ(x, k)
3 3
ω = 5Γ ω = 20 Γ
2 2 0.8

1 1
0.6
0 0
k

0.4
−1 −1

−2 −2 0.2
40 % lost 40 % lost
−3 −3 0.0
−4 −2 0 2 4 −1.0 −0.5 0.0 0.5 1.0
x x

Figure 13. Phase portraits showing the spiral for two different values of ω. The initial distribution is identical to the one used
in figure Fig. 10 and here we only consider the distribution after 40% of atoms lost.

[1] I. Bloch, J. Dalibard, and W. Zwerger, “Many-body sate,” Phys. Rev. A, vol. 85, p. 025602, Feb 2012. doi:
physics with ultracold gases,” Rev. Mod. Phys., vol. 80, 10.1103/PhysRevA.85.025602.
pp. 885–964, Jul 2008. doi:10.1103/RevModPhys.80.885. [3] Browaeys, A., Poupard, J., Robert, A., Nowak, S.,
[2] S. Knoop, J. S. Borbely, R. van Rooij, and W. Vassen, Rooijakkers, W., Arimondo, E., Marcassa, L., Boiron,
“Nonexponential one-body loss in a bose-einstein conden- D., Westbrook, C. I., and Aspect, A., “Two body loss
18

rate in a magneto-optical trap of metastable he,” Eur. [16] D. M. Harber, J. M. McGuirk, J. M. Obrecht, and
Phys. J. D, vol. 8, no. 2, pp. 199–203, (2000). doi: E. A. Cornell, “Thermally Induced Losses in Ultra-Cold
10.1007/s100530050027. Atoms Magnetically Trapped Near Room-Temperature
[4] L. Franchi, L. F. Livi, G. Cappellini, G. Binella, M. In- Surfaces,” Journal of Low Temperature Physics, vol. 133,
guscio, J. Catani, and L. Fallani, “State-dependent inter- pp. 229–238, Nov. (2003). doi:10.1023/A:1026084606385.
actions in ultracold 174yb probed by optical clock spec- [17] A. Johnson, S. S. Szigeti, M. Schemmer, and I. Bou-
troscopy,” New Journal of Physics, vol. 19, p. 103037, choule, “Long-lived nonthermal states realized by atom
nov 2017. doi:10.1088/1367-2630/aa8fb4. losses in one-dimensional quasicondensates,” Physical
[5] T. Tomita, S. Nakajima, I. Danshita, Y. Takasu, and Review A, vol. 96, p. 013623, July (2017). doi:
Y. Takahashi, “Observation of the mott insulator to su- 10.1103/PhysRevA.96.013623.
perfluid crossover of a driven-dissipative bose-hubbard [18] I. Bouchoule, M. Schemmer, and C. Henkel, “Cooling
system,” Science Advances, vol. 3, no. 12, p. e1701513, phonon modes of a Bose condensate with uniform few
2017. doi:10.1126/sciadv.1701513. body losses,” SciPost Phys., vol. 5, p. 043, 2018. doi:
[6] N. Syassen, D. M. Bauer, M. Lettner, T. Volz, D. Dietze, 10.21468/SciPostPhys.5.5.043.
J. J. García-Ripoll, J. I. Cirac, G. Rempe, and S. Dürr, [19] L. H. Dogra, J. A. P. Glidden, T. A. Hilker, C. Eigen,
“Strong dissipation inhibits losses and induces correla- E. A. Cornell, R. P. Smith, and Z. Hadzibabic, “Can
tions in cold molecular gases,” Science, vol. 320, no. 5881, three-body recombination purify a quantum gas?,” Phys.
pp. 1329–1331, 2008. doi:10.1126/science.1155309. Rev. Lett., vol. 123, p. 020405, Jul 2019. doi:
[7] T. Tomita, S. Nakajima, Y. Takasu, and Y. Takahashi, 10.1103/PhysRevLett.123.020405.
“Dissipative bose-hubbard system with intrinsic two- [20] P. Grišins, B. Rauer, T. Langen, J. Schmiedmayer,
body loss,” Phys. Rev. A, vol. 99, p. 031601, Mar 2019. and I. E. Mazets, “Degenerate bose gases with uniform
doi:10.1103/PhysRevA.99.031601. loss,” Phys. Rev. A, vol. 93, p. 033634, Mar 2016. doi:
[8] B. Yan, S. A. Moses, B. Gadway, J. P. Covey, K. R. A. 10.1103/PhysRevA.93.033634.
Hazzard, A. M. Rey, D. S. Jin, and J. Ye, “Observation of [21] B. Rauer, P. Grišins, I. E. Mazets, T. Schweigler,
dipolar spin-exchange interactions with lattice-confined W. Rohringer, R. Geiger, T. Langen, and J. Schmied-
polar molecules,” Nature, vol. 501, pp. 521–525, Sep 2013. mayer, “Cooling of a one-dimensional bose gas,” Phys.
doi:10.1038/nature12483. Rev. Lett., vol. 116, p. 030402, Jan 2016. doi:
[9] K. Sponselee, L. Freystatzky, B. Abeln, M. Diem, 10.1103/PhysRevLett.116.030402.
B. Hundt, A. Kochanke, T. Ponath, B. Santra, [22] M. Schemmer, A. Johnson, R. Photopoulos, and I. Bou-
L. Mathey, K. Sengstock, and C. Becker, “Dynamics of choule, “Monte carlo wave-function description of losses
ultracold quantum gases in the dissipative fermi–hubbard in a one-dimensional bose gas and cooling to the ground
model,” Quantum Science and Technology, vol. 4, state by quantum feedback,” Phys. Rev. A, vol. 95,
p. 014002, sep 2018. doi:10.1088/2058-9565/aadccd. p. 043641, Apr 2017. doi:10.1103/PhysRevA.95.043641.
[10] K. Honda, S. Taie, Y. Takasu, N. Nishizawa, M. Naka- [23] M. Schemmer and I. Bouchoule, “Cooling a bose gas by
gawa, and Y. Takahashi, “Observation of the sign reversal three-body losses,” Phys. Rev. Lett., vol. 121, p. 200401,
of the magnetic correlation in a driven-dissipative fermi Nov 2018. doi:10.1103/PhysRevLett.121.200401.
gas in double wells,” Phys. Rev. Lett., vol. 130, p. 063001, [24] I. Bouchoule and M. Schemmer, “Asymptotic tempera-
Feb 2023. doi:10.1103/PhysRevLett.130.063001. ture of a lossy condensate,” SciPost Phys., vol. 8, p. 060,
[11] J. Söding, D. Guéry-Odelin, P. Desbiolles, F. Chevy, 2020. doi:10.21468/SciPostPhys.8.4.060.
H. Inamori, and J. Dalibard, “Three-body decay [25] D. Rossini, A. Ghermaoui, M. B. Aguilera, R. Va-
of a rubidium bose–einstein condensate,” Applied tré, R. Bouganne, J. Beugnon, F. Gerbier, and
Physics B, vol. 69, pp. 257–261, Oct 1999. doi: L. Mazza, “Strong correlations in lossy one-dimensional
10.1007/s003400050805. quantum gases: from the quantum Zeno effect
[12] T. Weber, J. Herbig, M. Mark, H.-C. Nägerl, and to the generalized Gibbs ensemble,” Physical Re-
R. Grimm, “Three-body recombination at large scat- view A, vol. 103, p. L060201, June (2021). doi:
tering lengths in an ultracold atomic gas,” Phys. 10.1103/PhysRevA.103.L060201.
Rev. Lett., vol. 91, p. 123201, Sep 2003. doi: [26] J. J. García-Ripoll, S. Dürr, N. Syassen, D. M. Bauer,
10.1103/PhysRevLett.91.123201. M. Lettner, G. Rempe, and J. I. Cirac, “Dissipation-
[13] B. L. Tolra, K. M. O’Hara, J. H. Huckans, W. D. Phillips, induced hard-core boson gas in an optical lattice,” New
S. L. Rolston, and J. V. Porto, “Observation of reduced Journal of Physics, vol. 11, p. 013053, Jan. (2009). doi:
three-body recombination in a correlated 1d degenerate 10.1088/1367-2630/11/1/013053.
bose gas,” Phys. Rev. Lett., vol. 92, p. 190401, May 2004. [27] E. Misra and C. Sudarshan, “The zeno’s paradox in quan-
doi:10.1103/PhysRevLett.92.190401. tum theory,” J. Math. Phys., vol. 18, p. 756, 1977.
[14] F. Ferlaino, S. Knoop, M. Berninger, W. Harm, J. P. [28] W. M. Itano, D. J. Heinzen, J. J. Bollinger, and D. J.
D’Incao, H.-C. Nägerl, and R. Grimm, “Evidence for Wineland, “Quantum zeno effect,” Phys. Rev. A, vol. 41,
universal four-body states tied to an efimov trimer,” pp. 2295–2300, 1990. doi:10.1103/PhysRevA.41.2295.
Phys. Rev. Lett., vol. 102, p. 140401, Apr 2009. doi: [29] A. Beige, D. Braun, and P. L. Knight, “Driving atoms
10.1103/PhysRevLett.102.140401. into decoherence-free states,” New J. Phys., vol. 2,
[15] J. H. Gurian, P. Cheinet, P. Huillery, A. Fioretti, J. Zhao, pp. 22–22, sep 2000. doi:10.1088/1367-2630/2/1/322.
P. L. Gould, D. Comparat, and P. Pillet, “Observation of [30] A. Beige, D. Braun, B. Tregenna, and P. L.
a resonant four-body interaction in cold cesium rydberg Knight, “Quantum computing using dissipation to re-
atoms,” Phys. Rev. Lett., vol. 108, p. 023005, Jan 2012. main in a decoherence-free subspace,” Phys. Rev.
doi:10.1103/PhysRevLett.108.023005. Lett., vol. 85, pp. 1762–1765, Aug 2000. doi:
10.1103/PhysRevLett.85.1762.
19

[31] J. Kempe, D. Bacon, D. A. Lidar, and K. B. Wha- body losses: Weak dissipation and spin conservation,”
ley, “Theory of decoherence-free fault-tolerant universal Phys. Rev. A, vol. 104, p. 053305, Nov 2021. doi:
quantum computation,” Phys. Rev. A, vol. 63, p. 042307, 10.1103/PhysRevA.104.053305.
Mar 2001. doi:10.1103/PhysRevA.63.042307. [47] L. Rosso, L. Mazza, and A. Biella, “Eightfold way to
[32] P. Facchi, H. Nakazato, and S. Pascazio, “From the quan- dark states in su(3) cold gases with two-body losses,”
tum zeno to the inverse quantum zeno effect,” Phys. Phys. Rev. A, vol. 105, p. L051302, May 2022. doi:
Rev. Lett., vol. 86, pp. 2699–2703, Mar 2001. doi: 10.1103/PhysRevA.105.L051302.
10.1103/PhysRevLett.86.2699. [48] L. Rosso, A. Biella, J. De Nardis, and L. Mazza, “Dy-
[33] P. Facchi and S. Pascazio, “Quantum zeno subspaces,” namical theory for one-dimensional fermions with strong
Phys. Rev. Lett., vol. 89, p. 080401, Aug 2002. doi: two-body losses: Universal non-hermitian zeno physics
10.1103/PhysRevLett.89.080401. and spin-charge separation,” Phys. Rev. A, vol. 107,
[34] R. Schützhold and G. Gnanapragasam, “Quantum zeno p. 013303, Jan 2023. doi:10.1103/PhysRevA.107.013303.
suppression of three-body losses in bose-einstein conden- [49] M. Nakagawa, N. Tsuji, N. Kawakami, and
sates,” Phys. Rev. A, vol. 82, p. 022120, Aug 2010. doi: M. Ueda, “Dynamical sign reversal of magnetic
10.1103/PhysRevA.82.022120. correlations in dissipative hubbard models,” Phys.
[35] K. Stannigel, P. Hauke, D. Marcos, M. Hafezi, S. Diehl, Rev. Lett., vol. 124, p. 147203, Apr 2020. doi:
M. Dalmonte, and P. Zoller, “Constrained dynamics 10.1103/PhysRevLett.124.147203.
via the zeno effect in quantum simulation: Implement- [50] M. Nakagawa, N. Kawakami, and M. Ueda, “Exact liou-
ing non-abelian lattice gauge theories with cold atoms,” villian spectrum of a one-dimensional dissipative hubbard
Phys. Rev. Lett., vol. 112, p. 120406, Mar 2014. doi: model,” Phys. Rev. Lett., vol. 126, p. 110404, Mar 2021.
10.1103/PhysRevLett.112.120406. doi:10.1103/PhysRevLett.126.110404.
[36] Z. Gong, S. Higashikawa, and M. Ueda, “Zeno hall effect,” [51] G. Perfetto, F. Carollo, J. P. Garrahan, and
Phys. Rev. Lett., vol. 118, p. 200401, May 2017. doi: I. Lesanovsky, “Reaction-limited quantum reaction-
10.1103/PhysRevLett.118.200401. diffusion dynamics,” Phys. Rev. Lett., vol. 130, p. 210402,
[37] H. Fröml, C. Muckel, C. Kollath, A. Chiocchetta, and May 2023. doi:10.1103/PhysRevLett.130.210402.
S. Diehl, “Ultracold quantum wires with localized losses: [52] B. Doyon, “Lecture notes on Generalised Hydrodynam-
Many-body quantum zeno effect,” Phys. Rev. B, vol. 101, ics,” SciPost Physics Lecture Notes, p. 18, Aug. (2020).
p. 144301, Apr 2020. doi:10.1103/PhysRevB.101.144301. doi:10.21468/SciPostPhysLectNotes.18.
[38] K. Snizhko, P. Kumar, and A. Romito, “Quan- [53] A. Hutsalyuk and B. Pozsgay, “Integrability breaking in
tum zeno effect appears in stages,” Phys. Rev. the one dimensional Bose gas: Atomic losses and en-
Research, vol. 2, p. 033512, Sep 2020. doi: ergy loss,” Physical Review E, vol. 103, p. 042121, Apr.
10.1103/PhysRevResearch.2.033512. (2021). doi:10.1103/PhysRevE.103.042121.
[39] A. Biella and M. Schiró, “Many-Body Quantum Zeno Ef- [54] B. Bertini, F. H. Essler, S. Groha, and N. J. Robin-
fect and Measurement-Induced Subradiance Transition,” son, “Prethermalization and Thermalization in Mod-
Quantum, vol. 5, p. 528, Aug. 2021. doi:10.22331/q-2021- els with Weak Integrability Breaking,” Physical Re-
08-19-528. view Letters, vol. 115, p. 180601, Oct. (2015). doi:
[40] I. Bouchoule, B. Doyon, and J. Dubail, “The effect 10.1103/PhysRevLett.115.180601.
of atom losses on the distribution of rapidities in the [55] O. A. Castro-Alvaredo, B. Doyon, and T. Yoshimura,
one-dimensional Bose gas,” SciPost Phys. 9, 044, June “Emergent Hydrodynamics in Integrable Quantum Sys-
(2020). doi:10.21468/SciPostPhys.9.4.044. tems Out of Equilibrium,” Physical Review X, vol. 6,
[41] I. Bouchoule, L. Dubois, and L.-P. Barbier, “Losses in p. 041065, Dec. (2016). doi:10.1103/PhysRevX.6.041065.
interacting quantum gases: Ultraviolet divergence and [56] B. Bertini, M. Collura, J. De Nardis, and M. Fagotti,
its regularization,” Phys. Rev. A, vol. 104, p. L031304, “Transport in Out-of-Equilibrium X X Z Chains: Ex-
Sep 2021. doi:10.1103/PhysRevA.104.L031304. act Profiles of Charges and Currents,” Physical Re-
[42] F. Minganti, V. Savona, and A. Biella, “Dissipative phase view Letters, vol. 117, p. 207201, Nov. (2016). doi:
transitions in n-photon driven quantum nonlinear res- 10.1103/PhysRevLett.117.207201.
onators,” 2023. doi:10.48550/arXiv.2303.03355. [57] I. Bouchoule and J. Dubail, “Generalized hydrodynam-
[43] L. Rosso, A. Biella, and L. Mazza, “The one-dimensional ics in the one-dimensional bose gas: theory and ex-
Bose gas with strong two-body losses: the effect of the periments,” Journal of Statistical Mechanics: Theory
harmonic confinement,” SciPost Physics, vol. 12, p. 044, and Experiment, vol. 2022, p. 014003, jan 2022. doi:
Jan. (2022). doi:10.21468/SciPostPhys.12.1.044. 10.1088/1742-5468/ac3659.
[44] C.-H. Huang, T. Giamarchi, and M. A. Cazalilla, “Mod- [58] F. Lange, Z. Lenarčič, and A. Rosch, “Time-dependent
eling particle loss in open systems using keldysh path in- generalized Gibbs ensembles in open quantum systems,”
tegral and second order cumulant expansion,” 2023. doi: Physical Review B, vol. 97, p. 165138, Apr. (2018). doi:
10.48550/arXiv.2305.13090. 10.1103/PhysRevB.97.165138.
[45] B. Zhu, B. Gadway, M. Foss-Feig, J. Schachenmayer, [59] F. Lange, Z. Lenarčič, and A. Rosch, “Pumping approx-
M. L. Wall, K. R. A. Hazzard, B. Yan, S. A. imately integrable systems,” Nature Communications,
Moses, J. P. Covey, D. S. Jin, J. Ye, M. Holland, vol. 8, p. 15767, Jun 2017. doi:10.1038/ncomms15767.
and A. M. Rey, “Suppressing the loss of ultracold [60] Z. Lenarčič, F. Lange, and A. Rosch, “Perturba-
molecules via the continuous quantum zeno effect,” tive approach to weakly driven many-particle systems
Phys. Rev. Lett., vol. 112, p. 070404, Feb 2014. doi: in the presence of approximate conservation laws,”
10.1103/PhysRevLett.112.070404. Phys. Rev. B, vol. 97, p. 024302, Jan 2018. doi:
[46] L. Rosso, D. Rossini, A. Biella, and L. Mazza, 10.1103/PhysRevB.97.024302.
“One-dimensional spin-1/2 fermionic gases with two-
20

[61] P. L. Butzer and R. J. Nessel, Fourier analysis and ap- pp. 1129–1133, 2021. doi:10.1126/science.abf0147.
proximation. Lehrbücher und Monographien aus dem Ge- [64] M. Coppola and D. Karevski, “Some speculations about
biete der exakten Wissenschaften. Mathematische Reihe, local thermalization of nonequilibrium extended quan-
Bd. 40, Basel, Stuttgart: Birkhäuser, 1971. tum systems,” Condensed Matter Physics, vol. 26, no. 1,
[62] M. Schemmer, I. Bouchoule, B. Doyon, and J. Dubail, p. 13502, (2023). doi:10.5488/CMP.26.13502.
“Generalized hydrodynamics on an atom chip,” Phys. [65] G. Perfetto, F. Carollo, J. P. Garrahan, and
Rev. Lett., vol. 122, p. 090601, Mar 2019. doi: I. Lesanovsky, “Quantum reaction-limited reaction-
10.1103/PhysRevLett.122.090601. diffusion dynamics of annihilation processes,”
[63] N. Malvania, Y. Zhang, Y. Le, J. Dubail, M. Rigol, and arXiv preprint arXiv:2305.06944, 2023. doi:
D. S. Weiss, “Generalized hydrodynamics in strongly in- 10.48550/arXiv.2305.06944.
teracting 1d bose gases,” Science, vol. 373, no. 6559,

You might also like