You are on page 1of 13

Emergence of dark soliton signatures in a one-dimensional unpolarized attractive

Fermi gas on a ring


Andrzej Syrwid,1 Dominique Delande,2 and Krzysztof Sacha1, 3
1
Instytut Fizyki imienia Mariana Smoluchowskiego, Uniwersytet Jagielloński,
ulica Profesora Stanisława Łojasiewicza 11 PL-30-348 Kraków, Poland
2
Laboratoire Kastler Brossel, Sorbonne Universités, CNRS,
ENS-Université PSL, Collège de France, 4 Place Jussieu, 75005 Paris, France
3
Mark Kac Complex Systems Research Center, Uniwersytet Jagielloński,
ulica Profesora Stanisława Łojasiewicza 11 PL-30-348 Kraków, Poland
arXiv:1805.12103v1 [cond-mat.quant-gas] 30 May 2018

The two-component Fermi gas with contact attractive interactions between different spin compo-
nents can be described by the Yang-Gaudin model. Applying the Bethe ansatz approach, one finds
analytical formulae for the system eigenstates that are uniquely parametrized by the solutions of the
corresponding Bethe equations. The lowest energy egenstates at a given non-zero total momentum
subspaces called yrast states were recently analyzed for periodic boundary conditions. The numerical
studies show that the yrast spectrum of the Yang-Gaudin model resembles yrast dispersion relation
of the Lieb-Liniger model which in turn matches the dark soliton dispersion obtained within the
nonlinear Schrödinger equation. It was shown that such conjecture in the case of the Lieb-Liniger
model was not accidental and that dark soliton features emerged in the course of particle positions
measurement, when the system was initially prepared in an yrast eigenstate. Following the idea of
particle detection, we employ the Bethe ansatz approach as well as numerical diagonalization to
show that, starting from yrast eigenstates, the key soliton signatures, like density notch and phase
flip, are visible in the wave function when all fermions, but two belonging to different components
have been detected. With the help of the Bethe ansatz, we study soliton signatures in a wide
range of interaction strength. The discussion is supplemented by the analysis of the formation of
two-particle bound states built up by fermions with opposite spins.

PACS numbers: 67.85.Lm, 03.75.Ss, 03.75.Lm

I. INTRODUCTION tems provoked the investigations of similar structures in


Fermi systems. The two-component Fermi gas with at-
tractive interactions between fermions with different in-
Non-linear wave equations can possess solitonic solu- ternal degrees of freedom can form a superfluid state
tions that propagate without any change of their shape. which can be described by a set of non-linear Bogoliubov-
These extraordinary structures appear in a wide range de Gennes equations in the Baarden-Cooper-Schrieffer
of physical systems and may be formed by electromag- (BCS) regime [2]. Although this approach is dedicated to
netic waves in non-linear optics [1], as well as mat- the determination of ground state properties, it can also
ter waves particularly investigated in ultra-cold atomic be used to describe dark soliton solutions, where particle
gases. Bose-Einstein condensates (BEC), made up by a densities are very similar to those of the ground state,
single-component Bose gas cooled to nearly absolute zero but where the BCS pairing function reveals signatures of
temperature, turn out to be excellent playgrounds for the a dark soliton [43–45]. Passing the BCS-BEC crossover,
investigation of matter wave solitons [3–12]. In the mean- the BCS pairing function, with dark soliton signatures,
field description, we assume that every single atom in a becomes the dark soliton wavefunction of a molecular
BEC experiences an effective average potential and occu- BEC [45, 46].
pies exactly the same single-particle state. Such a case is
described by the Gross-Pitaevskii equation (GPE) that The experimental realization of dark solitons in a BEC
possesses bright and dark soliton solutions in one dimen- is based on the phase imprinting method [3, 4, 7, 8, 47–
sional (1D) space for attractive and repulsive interparti- 49]. The phase of the condensate can be modified by ap-
cle interactions, respectively [2]. The experimental real- plication of a short laser pulse whose intensity varies over
ization of both kinds of solitons confirmed the theoreti- the atomic cloud. In particular, it is possible to carve a
cal predictions obtained within the GPE [3–10, 13]. The dark soliton notch so that a half of a condensate cloud ac-
observation of the quantum nature of solitons, i.e. many- quires a π phase. It turns out that, using the same phase
body effects that go beyond the mean-field GPE descrip- engineering technique, one can observe the generation of
tion [14–42], is still very challenging from an experimen- pairs of dark and bright soliton-like states in noninteract-
tal point of view. Nevertheless, the rapid development of ing single-component fermionic system, where the Pauli
laboratory techniques devoted to investigations of ultra- blocking plays the role of interparticle repulsion [50, 51].
cold atomic gases gives an opportunity to study systems The same idea was applied to create a dark soliton in a su-
dominated by quantum many-body effects [13]. perfluid Fermi system [52]. However, the resulting state
The experimental observation of solitons in Bose sys- quickly decayed to a vortex that has been displayed in
2

numerical simulations [53, 54] and observed in the subse- the mean pair separation. Tightly bound pairs can be
quent experiment [55]. The analysis of the nature of the observed in the second case when the attraction is very
dark soliton BCS pairing function suggests that, in order strong. The thermodynamic description reveals that the
to excite a dark soliton in a superfluid Fermi system, only strongly attractive Yang-Gaudin model is closely related
one fermion of a Cooper pair has to undergo the phase to a strongly interacting gas of bosonic dimers described
imprinting procedure [56]. by the Lieb-Liniger model. That is, the ground state
In general, our understanding of quantum many-body energy of tightly bound pairs of fermions coincides with
systems is very limited. Fortunately, there are many- the energy of the attractive Bose gas, described by the
body systems in lower dimensions for which the bril- Lieb-Liniger model, which forms a highly excited super
liant method of Bethe ansatz is applicable [57]. This is Tonks-Girardeau phase [73–75, 88–91]. The latter can be
exactly the case of one-dimensional nonrelativistic Bose described by a system of attractive hard rods [73]. In the
and Fermi gases with particle interactions described by limit of infinitely strong interactions, the energy of the
point-like contact potentials [58–60]. In comparison to super Tonks-Girardeau phase matches the ground state
the case of identical bosons (Lieb-Liniger model [61, 62]), energy of the Tonks-Girardeau gas described by the Lieb-
the problem of the multi-component Fermi gas with con- Liniger model of strongly repulsive bosons [73]. Note that
tact interactions requires a generalization of the Bethe the pairing phenomenon in similar systems confined in a
ansatz procedure. The model of a Fermi gas consisting harmonic trap was meticulously analyzed in Ref. [92].
of arbitrary numbers of fermions with two internal de-
grees of freedom has been solved analytically by Yang
and Gaudin [63, 64]. Such an ultracold two-component
Fermi gas has been the subject of extensive studies [65– Although the link between the repulsive Lieb-Liniger
75]. gas and the attractive Fermi system described by the
Yang-Gaudin Hamiltonian is not entirely understood, it
The second branch of elementary excitations (the type is expected that the yrast excitations of the Fermi gas in
II excitations) of the Lieb-Liniger model with periodic question may correspond to dark solitons. The suppo-
boundary conditions corresponds to the so-called yrast sition is additionally supported by recent results show-
states, referring to the lowest energy at a given non-zero ing that the spectrum of yrast excitations in the Yang-
total momentum. For weak repulsive interactions, the Gaudin model is very similar to the corresponding type
type II eigenstates have been associated with dark soli- II spectrum of the Lieb-Liniger model which, in turn,
tons due to the coincidence between the yrast spectrum matches the dispersion relation of dark solitons in the
of the Lieb-Liniger model and the dark soliton dispersion weak interaction limit [75]. In analogy with the Bose
relation obtained within the mean field approach [76, 77]. case, eigenstates of the Yang-Gaudin system are trans-
The conjecture was underpinned by other strong argu- lationally invariant when we impose periodic boundary
ments presented in many publications [78–87]. The direct conditions [75]. Hence, we may expect that dark soliton
observation of the emergence of dark soliton signatures signatures are hidden in the translationally symmetric
during the particles’ positions measurement process, for yrast states and may be observed only by the analysis of
a system initially prepared in a type II eigenstate, has higher order correlation functions.
been reported recently [38, 39]. A similar analysis led to
the identification of dark soliton-like eigenstates of the
Lieb-Liniger model in the presence of an infinite square
well potential [41]. The present paper is devoted to the analysis of attrac-
For periodic boundary conditions, all energy eigen- tively interacting unpolarized systems of spin– 12 fermions
states are invariant under translations of all particles by with attractive contact interactions between different
the same distance. Therefore, the reduced single par- spin components, confined in a ring geometry. By ap-
ticle density is uniform and cannot display any soliton plying the Bethe ansatz approach, we investigate the for-
signature. Such a feature of the type II eigenstates was mation of pairs of ↓–↑ fermions and determine their size
the main impediment during investigation of their soli- in a wide range of interaction strength, when the sys-
ton character. The unequivocal connection between dark tem is prepared either in the ground state or in an yrast
solitons and yrast excitations in the Lieb-Liniger model state. As pointed out in [75], we observe the crossover
resulted in a broader examination of the solitonic na- between two significantly different physical regimes cor-
ture of yrast states. An ultracold balanced (unpolar- responding to a BCS-like gas and a gas of impenetrable
ized) gas of spin– 12 fermions can be described within bosonic dimers, when the relevant dimensionless inter-
the Yang-Gaudin model [60, 63–65, 67–75]. In the pres- action parameter γ ≈ −1. Following the Monte Carlo
ence of attractive interaction, two fermions with oppo- method [38, 39, 41] we repeatedly perform the successive
site spins tend to create a two-particle bound state. In measurement of particles’ positions, revealing the dark
the many-body case, there are two physically different soliton signatures, i.e. a density notch and a phase flip
regimes corresponding to weak and strong interaction in the wave function of the last anticipated pair of ↓–
limits. The first one refers to the BCS-like Cooper pair ↑ fermions. In addition, we analyze how the increasing
formation, for which the size of the pairs is larger than number of particles affects the soliton structure.
3

II. YANG-GAUDIN MODEL The summation in (3) is taken over all permutations π
of the permutation group SN↑ . The parity of permu-
A nonrelativistic ultracold gas of spin– 21 fermions in- tation π is extracted by sgn(π) = ±1, while, for real
teracting via point-like δ inter-components potential in x, the function sign(x) = x/|x|. The eigenstates (3)
1D can be described by the Yang-Gaudin model [59, 60, are uniquely parametrized by the sets of quasi-momenta
63, 64, 74, 93, 94]. Assuming that we deal with a system {kj }j=1,...,N↓ +N↑ and spin-roots {Λs }s=1,...,N↑ . The lat-
at zero temperature containing N↓ ≥ N↑ spin-down and ter quantities are auxiliary and appear due to the exis-
spin-up particles of equal masses m = m↓ = m↑ = 21 , the tence of two internal degrees of freedom interpreted as
Hamiltonian reads opposite spin directions. Since the Hamiltonian (1) com-
mutes with the total momentum operator
N↓ N↑ N↓ N↑
∂2 ∂2
δ(x↓j − x↑s ), (1)
X X X X
H=− − + 2c N↓ N↑
∂x↓2 ∂x↑2
s
X ∂ X ∂
j=1 j s=1 j=1 s=1 P = −i −i , (7)
j=1 ∂x↓j s=1 ∂x↑s
where the units have been chosen such that ~ = 2m = 1.
The number of particles in each single component N↓,↑
the states Ψ simultaneously satisfy the two eigenequa-
is a conserved quantity. Note that we can distinguish
tions
particles belonging to different spin components because
there is no spin flipping term in the Hamiltonian (1).
The interaction strength is measured by the following HΨ{k} = E{k} Ψ{k} , PΨ{k} = P{k} Ψ{k} , (8)
dimensionless parameter
where the eigenvalues E{k} and P{k} are simply given by
c the quasi-momenta kj [93, 94]
γ= , (2)
n
N↓ +N↑ N↓ +N↑
N +N
with n = ↓ L ↑ denoting the average particle density in
X X
E{k} = kj2 , P{k} = kj . (9)
the system of size L. j=1 j=1
While the Bethe ansatz formulation of the problem is
very simple, i.e. the eigenstates are superpositions of The symmetry properties of the wave functions Ψ,
plane waves, the structure of the resulting wave func-
tions is very cumbersome [59, 60, 63, 64, 72, 74, 93, 94]
Ψ(ρσ {xσ }, {k}, {Λ}) = sgn(ρσ )Ψ({xσ }, {k}, {Λ}), (10)
and hard to use in numerical calculations. Fortunately,
solutions of Yang-Gaudin model can be rewritten in the
determinant form [93, 94]
Ψ({xσ }, τ {k}, {Λ}) = sgn(τ )Ψ({xσ }, {k}, {Λ}), (11)
X
Ψ({x↓ }, {x↑ }, {k}, {Λ}) ∝ sgn(π) Wπ,↑ det Φ, (3)
π∈SN↑
Ψ({xσ }, {k}, η{Λ}) = sgn(η)Ψ({xσ }, {k}, {Λ}), (12)
where
for arbitrary permutations ρσ=↓,↑ ∈ SN↓,↑ , τ ∈ SN↓ and
N↑ h i η ∈ SN↑ , are discussed in details in Refs. [93, 94].
+ c sign(x↑l − x↑j ) ,
Y 
Wπ,↑ = i Λπ(j) − Λπ(l) (4)
Imposing periodic boundary conditions, i.e.
j<l

and the (N↓ + N↑ ) × (N↓ + N↑ ) matrix Φ, represented by ∀ : Ψ(. . . , x↓,↑ ↓,↑


j + L, . . .) = Ψ(. . . , xj , . . .), (13)
j=1,...,N↓,↑
two rectangular matrices separated by the vertical bar,
reads we obtain the following set of Bethe ansatz equations for
 
N↑ quasi-momenta {k} and spin-roots {Λ} [60, 63, 64, 69,

Aj (Λπ(s) , x↓l − x↑s )eikj xl 
Y
Φ =  72, 74, 75]
s=1
N↑

Y kj − Λn + i 2c
exp (ikj L) = , (14)
k − Λn − i 2c
 
N↑
n=1 j j=1,...,N↓ +N↑
ikj x↑
Y
 Aj (Λπ(s) , x↑m − x↑s )e m  ,
j=1,...,N↑ +N↓
s6=m
l=1,...,N↓
m=1,...,N↑
N↓ +N↑ N↑

(5) Y Λm − kj + i c Y Λ m − Λ n + ic
2
= .

with c
Λm − kj − i 2 Λ − Λn − ic

j=1 n=1 m m=1,...,N↑
c n6=m
Aj (Λ, x) = i(kj − Λ) + sign(x). (6) (15)
2
4

III. NUMERICAL METHOD IV. ATTRACTIVE INTERACTIONS: THE


GROUND STATE
Despite the fact that the many-body eigenstates Ψ can
be cast into a superposition of determinants, the analy- We start our considerations with the ground state in
sis of their properties is very burdensome. In fact, it is the presence of attractive interactions between different
intractable to extract valuable physical information from spin components (c < 0). Additionally, we restrict the
every single determinant of the Φ matrix. Moreover, the deliberation to the unpolarized system for which N↓ =
number of terms in the summation (3) dramatically pro- N↑ = N < ∞. It has been shown that, in such a case,
liferates with N↑ . In order to investigate the features of fermions with opposite spins tend to form bound state
Ψ, we should examine the corresponding correlation func- pairs what is reflected by the appearance of conjugate
tions. For this purpose, we have decided to numerically pairs of quasi-momementa kj,± = κj ± iµj , where κj =
simulate the measurement of particle positions. <(kj,± ), µj = |=(kj,± )| (see for example [60, 69, 72, 74]).
In general, the aforementioned simulations are based The ground state solutions of the Bethe equations (14)
on a one-by-one process of particle detection. Such an and (15) in the weakly (c → 0− ) and strongly (c → −∞)
approach requires the calculation of conditional prob- interacting limits are schematically depicted in Fig. 1.
ability densities for measurements of consecutive parti- c → 0− c → −∞
cles [21, 23, 38, 95–97]. Numerically, this method is ex- quasimomenta c/2
tremely expensive in the considered system. The result

Imaginary Part

Imaginary Part
spin-roots
of the measurement of M = N↓ + N↑ particles can also c
L
be obtained by another method, i.e. by a direct sampling c
0 0
of the corresponding M -particle probability density em- L

ploying the Monte Carlo algorithm of Metropolis et al.


−c/2
[98]. By using the analytical expression for M -particle −4π/L −2π/L 0 2π/L 4π/L −2π/L −π/L 0 π/L 2π/L
probability distribution |Ψ(r1 , . . . , rM )|2 and following Real Part Real Part
Refs. [39, 41, 99], we perform a so-called Markovian walk
in the configuration space, generating a sequence of sam- FIG. 1: Sketch of the solutions of the Bethe equations (14)
ples R = {r1 , . . . , rM } called a Markov chain. In our case, and (15) in the complex plane, for the balanced ground state
with N↓ = N↑ = N = 5. Left (Right) panel corresponds to
we assume that rj = x↓j for j = 1, . . . , N↓ and rN↓ +j = x↑j the regime of weak (strong) attraction c → 0− (c → −∞). The
for j = 1, . . . , N↑ . Technically speaking, if R is the last resulting quasi-momenta form conjugate pairs kj,± = κj ±iµj .
element of the Markov chain, the next randomly chosen In both panels, the spin-roots Λj are equal to κj .
set of particle positions R0 is accepted with probability
p = min(1, |Ψ(R0 )|2 /|Ψ(R)|). If R0 is not accepted, we The expansion of the Bethe equations leads to the ob-
again append the set R at the end of the Markov chain. servation that in the two regimes in question, the ground
The Metropolis procedure increases significantly the state solutions take the following forms
numerical efficiency but still allows for the studies r
of few body systems only. In order to investi- 2π c→0− |c|
gate system containing more particles, one can em- lim κj = lim Λj = nj , µj ≈ ,
c→0− c→0− L L
ploy numerical diagonalization of the Hamiltonian (1) (17)
in a truncated Hilbert space. Eigenstates of the π c→−∞ |c|
system can be represented in the Fock state basis lim κj = lim Λj = nj , µj ≈ ,
hQ Ei hQ

Ei

c→−∞ c→−∞ L 2
smax smax
j=smin mj j=smin mj , where the j-th single
with nj ∈ − N 2−1 , . . . , N 2−3 , N 2−1 [60, 74]. The solu-

particle mode φj (x) = L−1/2 exp[i2πjx/L] is occupied by tions can be interpreted as filling a "Fermi sphere" with
mj↓,↑ = 0, 1 particles. The numbers smin and smax deter- the "Fermi surface" referring to the "Fermi momentum"
mine the modes taken into account and have to be ad- ±maxj (Λj ). Note that the correspondingP binding ener-
justed to reproduce the examined eigenstate accurately. gies per ↓–↑ pair, defined as εB = − N1 j [=(kj )]2 =
The Hamiltonian (1) commutes with P, Eq. (7), so in − N1 j µ2j , are the following
P
the chosen basis H is partitioned into blocks referring to
different values of the total momentum P . Note that the c→0− |c| c→−∞ c2
following constraints have to be satisfied εB ≈ −2 , εB ≈ − . (18)
L 2
sX
max sX
max In general, the Bethe equations are very difficult to
X 2π
mσj = Nσ , P = mσj = J , (16) solve for arbitrary values of c < 0. Therefore, it is conve-
L
j=smin σ={↓,↑} j=smin nient to start with one of the considered limits and, by
employing a simple linear approximation, consecutively
where J ∈ Z. By definition the yrast states correspond to increase or decrease the coupling strength [75]. The re-
the lowest energy eigenvalue for a given total momentum sults of this procedure applied to the 5+5 particle ground
(in fact, given by J ). state (N = 5) are presented in Fig. 2. In the strongly
5

attractive case, we can apply an additional approxima- taken in order to form pairs and calculate distances (20).
tion and simplify the Bethe equations. That is, one can In other words, the obtained histograms will not change
replace kj,± by Λj ± 2c and by simple algebraic manipu- if we randomly permute particles. This is the reason why
lations obtain [75, 100] the background density ≈ 2 appear in the histograms in-
N   dependently on the attraction strength (see Fig. 3).
X Λm − Λn
2Λm L = 2πlm −2 arctan . (19)
c

n=1

m=1,...,N
7
where the soultions are determined by distinct quantum c = − 35 (γ = − 3.5)
6
numbers lm . By substituting lm = nm , cf. (17), we get c = − 25 (γ = − 2.5)

normalized histogram
the parametrization of the ground state of the balanced 5 c = − 10 (γ = − 1.0)
gas of fermions. c = − 5 (γ = − 0.5)
4
4π/L 50
3π/L
2π/L 25 3
Re[k ]

Im[k ]
π/L
0 0
-π/L
-2π/L -25 2
-3π/L
-4π/L -50
4π/L 1
8x10-10
3π/L
2π/L 4x10-10
Re[Λ]

π/L 0
0 0 Im[Λ] 0 0.1 0.2 0.3 0.4 0.5
-π/L
-10 relative distance
-2π/L -4x10
-3π/L
-4π/L -8x10-10 FIG. 3: Histograms of the relative distance ∆n j [see (20)] be-
-100 -75 -50 -25 0 -75 -50 -25 0
tween fermions with opposite spins for different values of the
cL = 10γ cL = 10γ
coupling constant c. The considered 5+5 particles system of
size L = 1 was prepared initially in the ground state. The
FIG. 2: Solutions of the Bethe equations (14) and (15) for measurement of the particle positions was performed with
the ground state with N↑ = N↓ = N = 5 versus the cou- the help of the Metropolis algorithm, as described in Sec. III.
pling constant c (γ = cL/2N ). While upper panels represent Increasing the inter-components attraction above c = −10
the real and imaginary parts of the resulting quasi-momenta (γ = c/2N = −1) one observes an escalation of small-sized
kj , lower panels refer to the spin-roots Λj solutions. Note ↓–↑ pairs occurrence. That is, when γ < −1, the effective
that, in the two limiting cases of weak and strong interac- attraction causes the creation of dimers of size smaller than
tion strength, both quasi-momenta and spin-roots follow the their mean separation. The background density ≈ 2 domi-
predictions (17) depicted in Fig. 1. nates in all cases for distances larger than the average dis-
tance between particles belonging to the same spin compo-
Let us now analyze in details the problem of pair- nent δ̄ = L/N = 0.2. The histograms have been prepared
ing of fermions belonging to different components. For from the data collected from several millions of measurement
this purpose, we have decided to investigate histograms realizations.
of the relative distance between particles that are ob-
tained in many measurement realizations. Dealing with The average distance between particles possessing the
the N↓ = N↑ = N = 5 system and employing the same spin δ̄ = L/N (δ̄ = 0.2 for L = 1 and N = 5) is a
Bethe ansatz solution, we perform numerical simulations reference quantity. When γ = cL/2N . −1 the ↓–↑ pair-
of the particle measurement process with the help of the ing becomes visible in Fig. 3. That is, if the attraction
Metropolis routine (see Sec. III). In every single j-th re- is strong enough, the size of the pairs is smaller than δ̄.
alization of the detection procedure,
 we collect two sets In this way we enter the regime of tightly bound pairs
of particle positions i.e. Xjσ = xσj,1 , xσj,2 , . . . , xσj,5 with of fermions with opposite spins. On the other hand, we
σ =↓, ↑. The relative distance on a ring of size L can be expect that formation of Cooper-like pairs of size greater
defined as follows than δ̄ when γ & −1 [75]. A careful analysis of Fig. 3 re-
 
∆nj = min x↓j,n − x↑j,n , L − x↓j,n − x↑j,n , (20)
veals oscillations in the profiles of the distributions with
period ≈ δ̄. Moreover, in the strongly attractive case
where the j and n indices refer to the measurement re- c = −35 (γ = −3.5), one notices a density dip near
alization and to the particle number, respectively. By the relative distance 0.08. It is clear that, in such a
collecting all ∆nj distances (for all j and n) after many regime, fermions are tightly bound and we deal with a
realizations of the particle detection process, one can pre- gas of impenetrable bosonic dimers. The particles coming
pare a histogram of relative distances between spin-up from the same spin component feel the Pauli exclusion.
and spin-down fermions. In Fig. 3, we compare such his- Hence, the ↓–↑ molecules tend to distribute themselves
tograms for different strengths of inter-components at- uniformly in space. This simple mechanism is respon-
traction. We stress that, if one collects many measure- sible for the oscillating behavior visible in Fig. 3. The
ment realizations, it does not matter which positions are same features can be observed within the BCS approach
6

(see [101]) if we analyze the following correlation function as


ψ̂↓ (x)ψ̂↑† (y)ψ̂↑ (y)ψ̂↓ (x) , where ψ̂σ (x) are the canonical

 
e n = min x↓ − x↑ ↓ ↑
∆ j,τ (n) , L− xj,n − xj,τ (n) , (22)

field operators and the average h.i is taken in the BCS j j,n
ground state.
The above discussion concerning the creation of tightly and measures the relative distance between paired
fermions. As before, we collect all the distances ∆e n and
bound molecules when γ . −1 stays in a very good j
agreement with the results which can be obtained for the prepare histograms corresponding to the distributions of
two-body problem. The relative distribution of two dis- relative distances between fermions with opposite spins
tinguishable particles interacting via a contact attractive that are paired. It turns out that the obtained results
potential is well known, [74, 102, 103] match quite well the normalized 2-particle solutions (21).
Such an agreement confirms that the system is dominated
|ψ(r)|2 ∝ e−|c|r , (21) by two-body physics for γ . −1. The many-body numer-
ical outcomes and the above mentioned two-body results
are presented in Fig. 4. Note that the same analysis in the

where r = x − x↑ is the relative distance of the two
particles. By rewriting c = presence of weak attraction cannot be performed. That
 2γN/L = 2γ/δ̄, one obtains is, when the size of the anticipated ↓–↑ pairs is larger
|ψ(r)|2 ∝ exp − 2|γ|r/δ̄ . Then the molecule size is
than the average dimer separation δ̄, we do not have any
given by δ̄/|γ|. We will see that this simple two-body
practical tool to determine which fermions in the consid-
result matches the results obtained in the many-body
ered gas are paired.
simulations. Furthermore, it is now straightforward that
the state (21) fits in the system of size L = N δ̄ for γ .
−1. V. ATTRACTIVE CASE: YRAST EXCITATION
12
c = −10 (γ = −1)
30 c = −30 (γ = −3) Let us now consider an yrast eigenstate of the balanced
2

10
normalized histogram & 2−body |ψ|

25
8 many−body system containing N↓ = N↑ = N = 5 particles. It turns
20
2−body |ψ|2 out that, in the present case, the weak and strong in-
6 15
4 10 teraction regimes are separated by a bifurcation of the
2 5 solutions of the Bethe ansatz equations. Following the
0 0 discussion presented in [75], we have chosen the yrast ex-
0 0.1 0.2 0.3 0.4 0.5 0 0.06 0.12 0.18 0.24
citation corresponding to the total momentum P = 6π L .
50 70
c = −50 (γ = −5) 60 c = −70 (γ = −7) In the small |c| limit, where the binding energy (18)
40 changes linearly with |c|, one immediately notices that,
50
30 40 in order to excite the lowest energy eigenstate belonging
20 30 to the subspace of P = 6π L , it is energetically favorable
20 p
10 to break the pair of quasi-momenta k = ±i |c|/L from
10
0 0 the ground state parametrization (see Fig. 1) and set one
0 0.03 0.06 0.09 0 0.03 0.06 0.09 of them to k = 0 and the other one to 6π L . Such maneu-
relative distance relative distance
ver fulfills the total momentum requirement but still it is
FIG. 4: Distributions of the relative distance between paired not clear what is the structure of the corresponding spin-
fermions belonging to different spin components. Normalized roots. As before, by numerical investigations and the ex-
histograms represent the results obtained from many-body pansion of the Bethe equations (14), (15) we easily find
simulations that are based on the ∆ en
j quantities defined in the limiting values of quasi-momenta and spin-roots for
Eq. (22). Red dashed lines depict the corresponding two- the yrast state in question (see the left panel of Fig. 5).
body probability densities (21). Note that the many-body Note that the scheme of the yrast excitation resembles
results follow quite well the simple two-body solutions indicat- the collective excitation of a single component Fermi gas
ing domination of the two-body physics for γ . −1. The sys- discussed in [50, 51, 104]. Indeed, for the very weakly in-
tem was prepared initially in the ground state. The size of the teracting 5+5 particle system, the considered yrast state
1D space is L = 1 and the number of particles N↓ = N↑ = 5.
is represented by the following superposition
√ 
Armed with this knowledge, we can investigate the 2 
|Ψi ≈ |{y}i↓ |{g}i↑ + |{g}i↓ |{y}i↑ , (23)
pairing phenomenon in the strongly attractive regime 2
(when γ . −1) in details. In order to check how the pair-
where the Fock states (σ =↓, ↑)
ing depends on the attraction strength, we have decided
to determine which fermions belonging to the measure- |{g}iσ = |. . . , 0−3 , 1−2 , 1−1 , 10 , 11 , 12 , 03 , . . .i ,
ment collections {Xj↓ } and {Xj↑ } collections are actually |{y}iσ = |. . . , 0−3 , 1−2 , 1−1 , 00 , 11 , 12 , 13 , 04 , . . .i ,
(24)
paired. For this purpose, for every single j-th realization
of the measurement process, we found the permutation describe the occupation (i.e. 0j or 1j ) of single particle
P en en
τ ∈ SN↑ minimizing the sum n ∆ j , where ∆j is defined momentum states ∝ exp[i2πjx/L]. This result stays in
7

agreement with the BCS prediction that, in the weakly P = 6πL in a wide range of attraction strength (see Fig. 6).
interacting case, only one component of the Fermi gas has The resulting quasi-momenta reveal a bifurcation around
to be collectively excited in order to reproduce dark soli- cL ≈ −9.05 (γ ≈ −0.905), where we observe a transi-
ton features [56]. We stress that the BCS regime, where tion between two different parametrization scenarios of
Cooper pairs are much larger than the average interparti- the same yrast eigenstate. The bifurcation point coin-
cle distance, is very hard to simulate numerically within cides with the interaction strength where the pairing of
the Bethe approach. Indeed, in order to deal with such fermions starts to be dominated by two-body physics. In-
Cooper pairs the system must be much larger, i.e. the deed, we have justified in Sec. IV that the size of dimers is
total number of particles N  5, that is not attainable comparable to the mean separation between fermions of
with the current computer resources. the same kind when γ ≈ −1 and the pairing is dominated
On the other hand, in the strongly attractive limit, by two-body physics for γ . −1.
the binding energy increases very quickly ∼ c2 . Hence, to
6π/L 10
deal with the yrast state one cannot break any pair of the 5π/L
4π/L 5
ground state quasi-momenta. This case is closely related

Im[k]
3π/L 50

Re[k ]

Im[k ]
2π/L 0
to the type II excitations known from the Lieb-Liniger π/L
-50 0
0 -100 0
model [38, 39, 58, 59, 61, 62, 84–86]. It was also pointed -π/L cL=10γ
-2π/L -5
out in Ref. [75] that, in the c → −∞ limit, the yrast exci- -3π/L
-4π/L
tation relies on the shift of a pair of quasi-momenta just -20 -15 -10 -5
-10
4π/L 8x10-10
above the Fermi surface. The same thing has to be done 3π/L
2π/L 4x10-10

Re[Λ]
with the corresponding spin-root. In our case, the pair π/L
0 0 Im[Λ]
k = ±i 2c and the spin-root Λ = 0, related to the ground -π/L -10
-2π/L -4x10
state, have to be moved to the values k = 3π c
L ± i 2 and -3π/L -10
3π -4π/L -8x10
Λ = L (see the right panel of Fig. 5). Such an excita- -50 -40 -30 -20 -10 0 -40 -30 -20 -10 0
tion scenario is very similar to the one-hole excitation in cL = 10γ cL = 10γ
the Lieb-Liniger model that reveals totally dark soliton
structures [38, 39]. That is, when the momentum per bo- FIG. 6: Quasi-momenta and spin-roots solutions of the Bethe
π
son of an yrast exictation approaches ± L , one expects a equations (14) and (15) corresponding to the unpolarized 5+5
single point in configuration space where the wave func- particles yrast state with total momentum P = 6π L
. Left
tion of the last boson reveals a phase flip by π indicating (Right) panel refers to real (imaginary) parts of kj and Λj
the presence of a dark soliton density notch. Therefore, versus the coupling constant c. Note that the results approach
we have chosen the total momentum 5L 6π
per dimer in the to the limiting solutions shown in Fig. 5. Vertical dashed lines
indicate the interaction strength [cL ≈ −9.05 (γ ≈ −0.905)]
hope of observing clearly visible dark soliton signatures
where the bifurcation takes place.
in the two-component Fermi gas.

c → 0− c → −∞
We expect that tightly bound ↓–↑ molecules appear for
c/2
γ . −1. For the considered yrast state, we can carry out
quasimomenta
the approach used for the ground state, i.e. by collecting
Imaginary Part

Imaginary Part

spin-roots
c ∆,
e defined in Eq. (22), in many realizations, we can com-
L

c
0 0 pute the distributions of relative distance between paired
L fermions. The dimer size is determined by the expecta-
e 2 . That is, the quantity
tion value of ∆
−c/2
−6π/L−4π/L −2π/L 0 2π/L 4π/L 6π/L −3π/L−2π/L −π/L 0 π/L 2π/L 3π/L
v
u1 n 2
u X
Real Part Real Part
ξ = 2t ∆e , (25)
n j=1 j
FIG. 5: Scheme of the limiting solutions of the Bethe equa-
tions (14)-(15) corresponding to the yrast state with total mo-
mentum P = 6π L
obtained for an unpolarized system contain- where n denotes the number of collected ∆ e distances, is
ing N = 5 particles in each component. Left (Right) plots related to the width of the two-body probability distribu-
in the complex plane refer to the weakly (strongly) attractive tion (21). The quantity ξ has been calculated in a wide
limit c → 0− (c → −∞). Filled circles and diamonds show range of interaction strength for the ground state and
the resulting quasi-momenta and spin-roots, respectively. In for the yrast state with P = 6π L . It turns out that, for
comparison to the ground state solutions (see Fig. 1), just a
c & −25 (γ & −2.5), the results for both states overlap
few values (indicated by empty symbols) have been modified.
as one can see in Fig. 7. For c . −25 (γ . −2.5), we
enter the stronger interaction regime where the pairing is
The structure of the Bethe solutions for the yrast eigen- definitely dominated by two-body physics. Then, the av-
state changes dramatically between weakly and strongly erage dimer size ξ is slightly larger for the yrast state than
attractive regimes. Following the step by step proce- for the ground state. Such a behavior can be attributed
dure [75] mentioned in Sec. IV, we obtained the yrast to the fact that there is much more kinetic energy in the
solutions of the equations (14)-(15) for N = 5 and excited eigenstate than in the ground state.
8

0.25 and the wavefunction of the last two fermions reads


N = 5, L = 1
0.2 x↓1,...,N −1 , x↓ }, {e
Ψ2 (x↓ , x↑ ) = Ψ({e x↑1,...,N −1 , x↑ }). (26)

In the following we analyze properties of (26). We con-


pair size ξ

0.15 sider two different ways of measuring the positions of


2N − 2 particles:
0.1
1. We can assume that the measurement takes place
mean dimmer separation L/N when two fermions with spin up and down are de-
0.05 ground state tected at the same positions, i.e. x↓s = x↑s for s =
yrast state 1, 2, . . . , N − 1. It resembles the rapid ramp tech-
nique used in experiments where a sweep across a
0 Feshbach resonance leads to the creation of tightly
-80 -70 -60 -50 -40 -30 -20 -10 0
c = 10γ bound molecules [105–115] and a subsequent mea-
surement of the molecular density is performed. We
FIG. 7: Dependence of the pair size ξ (25) versus the interac- will call this measurement "zero-size" because in
tion strength c = 10γ in the unpolarized N↑ = N↓ = N = 5 this case the size of pairs of ↓–↑ fermions is zero.
particles system of size L = 1 prepared in the ground state
(red curve with filled circles) and in the yrast state (blue curve 2. We can perform the measurement of particle posi-
with filled diamonds) with P = 6π. Note that both curves tions by sampling the many-body probability den-
overlap for c ≥ −25 (γ ≥ −2.5). For stronger interactions, sity without any constraint. In other words, we do
the pair size is slightly larger for the yrast eigenstate than
not measure the pairs but single particles. This
for the ground state. The average distance between dimers
δ̄ = L/N = 0.2 is indicated by the black dashed line. kind of the measurement will be dubbed "any-size"
because we assume that pairs of fermions can have
any size.
The very strongly attractive regime is very difficult Let us start with an analysis of a single realization of
to study numerically. To satisfy the periodic boundary the zero-size measurement process. We consider an unpo-
conditions, we have to operate with quadruple precision larized system of N↓ = N↑ = N = 5 particles, L = 1 and
which turns out to be insufficient when γ < −7. Such a wide range of interaction strength, i.e. −70 ≤ c ≤ −0.1
requirements come from the fact that in the analytical (−7 ≤ γ ≤ −0.01). The system is prepared initially in
expressionσ (3) of the wave function Ψ, there are exponen- the yrast state with total momentum P = 6π which is an-
tials eikj xn : for complex quasi-momenta =(kj ) ≈ ±i 2c alyzed in Sec. V. The structures of the observed two-body
with a large c, the quadruple precision is insufficient. wave function (26) depend on the positions where the first
Moreover, in order to properly reproduce the investigated 2N − 2 particles are measured. For a small particle num-
distributions by means of the Metropolis procedure, one ber, a single realization of the measurement process may
needs more steps of a Markovian walk for strong attrac- result in a two-particle wave function (26) which does
tion. Hence, we restrict our further studies to γ ≥ −7. not clearly show dark soliton signatures. Therefore, we
have decided to choose optimal positions of the measured
pairs of fermions, that is a configuration of the pairs that
corresponds to the maximal value of the probability den-
VI. YRAST STATE: EMERGENCE OF DARK
SOLITON SIGNATURES
sity, e.g. x e↓j = x
e↑j = 51 (j − 1) for j = 1, 2, 3, 4. This
choice means that the last remaining pair of fermions
is likely to be detected between x↓ = x↑ = 0.6 and 1
Yrast states in the Yang-Gaudin model are expected where there is the largest free space interval. The mod-
to be strictly connected with dark solitons [75]. Because ulus and phase of the resulting two-body wave functions
of the ring geometry of the system, the eigenstates are Ψ2 (x↓ , x↑ ) = |Ψ2 |eiφ2 (26) are presented in Fig. 8. It
translationally invariant. Therefore, the corresponding turns out that, in such a case, the amplitudes |Ψ2 | and
reduced single particle probability density cannot possess phase distributions φ2 reveal clearly visible a density
any soliton-like features. We expect that dark soliton notch and a phase jump localized around x↓ ≈ x↑ ≈ 0.8
structures can emerge due to the spontaneous breaking both for weak and strong attraction. Such dark soliton
of the translational symmetry induced by measurements signatures do not emerge in the course of the particle
of particle positions, like in the Lieb-Liniger model [38, detection process when we start with the ground state
39, 41]. Starting with the yrast eigenstate of the balanced (for comparison, see top panels of Fig. 8). Note that,
N↓ = N↑ = N system, we have performed numerical as expected, the stronger interactions we deal with, the
simulations of the measurement of positions of N↑ − 1 more dominant the diagonal elements of the density are.
spin-up fermions and of N↓ −1 spin-down fermions. Then, Furthermore, thanks to the Pauli exclusion principle, in
we know the positions of 2N − 2 particles, x e↓,↑
j=1,...,N −1 , the plots of the phase distributions, one notices a nodal
9

structure at the positions of the initially measured zero-


sized pairs of ↓–↑ fermions. Such nodal structures are
present for any interaction strength, both for the ground
state and the yrast eigenstate and can be explained by a
simple reasoning. That is, the wave function describing
two identical noninteracting fermions can be cast into the
form
α
ϕ(x, x + ε) ∝ eiαx ei 2 ε sin (βε) , α ∈ R, β ∈ R+ , (27)

that simply reveals the π-phase flip at x when ε passes


through zero.
Figure 9 are cuts along the diagonals of the two-
dimensional plots shown in Fig. 8, i.e. we present the
probability density and the phase of Ψ2 (x, x). We ob-
serve a dark soliton like density notch and a phase flip
at x = 0.8, and also notice that the distance between the
two main peaks of probability density around x = 0.8
slightly increases when γ becomes more negative. Note
that the yrast state for very weak interactions corre-
sponds to a single fermion excitation, while, in the strong
attraction limit, it is related to the excitation of a pair
of fermions. In the former case, the momentum of a
single fermion has to be twice larger than the momen-
tum of each fermion in the latter case. Consequently, we
deal with longer wavelengths for strong interaction which
can be responsible for the observed increase of the dis-
tance between the two peaks around the density notch.
Since one deals with a Fermi system, the wave function
Ψ2 (x, x) vanishes at the positions of the initially mea-
sured dimers. For comparison, we also depict the results
for the ground state in the weakly interacting case, for
γ = −0.01. One immediately notices that no soliton sig-
nature around x = 0.8 can be observed in this case .
So far we have analyzed dark soliton signatures in the
wave function of the last pair of fermions when the N − 1
pairs of zero-size are assumed to be measured at the
equidistant positions. Such a configuration allows us to
observe clearly the dark soliton signatures. However, in
experiments pairs of fermions or fermions themselves are
detected at random positions according to the probabil-
ity density of the yrast state. Now, preparing the system
initially in the same yrast state as before, we investi-
gate the diagonal probability density and the phase dis-
tribution for the last pair of ↓–↑ fermions averaged over FIG. 8: Wave function Ψ2 (x↓ , x↑ ) = |Ψ2 |eiφ2 for the last two
many realizations of the particle detection process. By fermions with opposite spins in the 5+5 particles system on
employing the Metropolis routine and the Bethe ansatz a ring of length L = 1. We consider the yrast state corre-
approach, we have performed numerical simulations of sponding to total momentum P = 6π for different coupling
the measurement of N − 1 = 4 dimers assuming that strengths, i.e. −70 ≤ c ≤ −0.1 (−7 ≤ γ ≤ −0.01). It is
assumed that four pairs of spin up and down fermions (of
they have either zero-size or any-size (i.e. measurement
zero size) have been measured at the positions corresponding
of particles without any additional restrictions). By the
to the maximal probability density, e.g. x e↓j = x
e↑j = j−1
5
for
fact that we deal with periodic boundary conditions the j = 1, 2, 3, 4. The resulting amplitudes (left panels) and phase
position of the phase flip indicating the soliton-like struc- (right panels) of the wave function of the last two fermions
ture varies randomly from one measurement realization show a density notch and a phase flip along the diagonal, re-
to another one. In order to determine average distribu- spectively. Such dark soliton-like signatures appear around
tions we shift all the results so that the corresponding x↓ ≈ x↑ ≈ 0.8, i.e. exactly between most distant dimers
phase flip is always located at L2 = 0.5. In every single that have been measured. For comparison, the ground state
realization of the detection process, the diagonal phase wave function Ψ2 (x↓ , x↑ ), which can be chosen as a real-valued
distribution φ2 (x, x) of the last ↓–↑ pair reveals a flip. function, is depicted in the first row.
10

ground state c = − 0.1 (γ = − 0.01) zero-size pair any-size pair


yrast state c = − 0.1 (γ = − 0.01) 3
c = −0.01 (γ = −0.001)

probability density
yrast state c = − 10 (γ = − 1.0) 2.5 c = −10 (γ = −1.0)
yrast state c = − 20 (γ = − 2.0) 2 c = −70 (γ = −7.0)
6 yrast state c = − 35 (γ = − 3.5)
1.5
yrast state c = − 70 (γ = − 7.0)
probability density

1
π/2 ground state
0.5
yrast state
phase

4 0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 1
π
position position
− π/2
0 0.2 0.4 0.6 0.8 1 c = − 0.01 (γ = −0.001) c = − 10 (γ = −1.0) c = − 70 (γ = −7.0)
position any-size pair zero-size pair
2 π

phase
0

0
0 0.2 0.4 0.6 0.8 1 −π
position 0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 1
position position position
FIG. 9: Diagonal part of |Ψ2 (x↓ , x↑ )|2 shown in Fig. 8. The
inset presents the corresponding diagonal phase distribution FIG. 10: Diagonal probability densities and phase distribu-
φ2 which is identical for all coupling constants c < 0. The tions for the last pair of ↓–↑ fermions averaged over many real-
density notch is clearly visible around x = 0.8 and its position izations of the detection process for a system of N↓ = N↑ = 5
coincides with the position of the phase flip. For comparison, fermions prepared in the yrast state with total momentum
the same numerical experiment was performed for the ground P = 6π (the system size L = 1). Upper panels show the aver-
state for weak attraction c = −0.1 (γ = −0.01) when no aged densities for different attraction strengths obtained us-
soliton-like structure can be observed around x = 0.8. ing two alternative initial measurement schemes: left (right)
panel corresponds to zero-size (any-size) of fermionic pairs.
Lower panels display the averaged phases obtained within dif-
ferent detection schemes and for different attraction strengths.
To satisfy the periodic boundary conditions the relation Note that they are almost identical independently of γ and of
φ2 (L, L) − φ2 (0, 0) = 2πJ, where in general J ∈ Z, has to the applied detection scheme. They reveal down (J = 0) and
be fulfilled. In the limiting case where the soliton is com- up (J = 1) phase flips.
pletely dark (i.e. when the density drops to zero like in
Fig. 9), the phase flip occurs abruptly at a single point,
i.e. lim→0 [φ2 (xS +, xS +) − φ2 (xS −, xS −)] = ±π average phase distribution.
modulo 2π, where xS is the soliton position. Then, all The last thing we would like to consider is the influ-
J ∈ Z become equivalent and cannot be distinguished. ence of the particle number N = N↓ = N↑ on the soli-
Therefore, in such a case we show the phase plot corre- ton structures. For weak interactions, systems contain-
sponding to J = 0 only, see in Fig. 9. ing more than N = 5 particles in each component can be
The results presented in the upper panels of Fig. 10 studied by numerical diagonalization of the Hamiltonian
show the density notches in the diagonal probability (1) (see Sec. III). In order to compare N > 5 particles
density for the last fermions |Ψ2 (x, x)|2 (26), averaged systems with the N = 5 results explored so far in this
over many realizations of the measurement of 2N − 2 paper, we have to choose exactly the same type of an
fermions. The notches are clearly visible for all interac- yrast excitation. For this purpose, we investigate odd
tion strengths independently of the measurement scheme numbers of particles in each component up to N = 13.
(zero-size or any-size) of the detected fermionic pairs. The yrast excitation corresponds to the total momentum
Moreover, we always observe phase flips of two types: P = π(N + 1) for the system of size L = 1. Converged
facing up (J = 1) and down (J = 0) that are collected results can be obtained only for weak attraction. By
separately in lower panels of Fig. 10. In contrast to the comparing with results obtained using the Bethe ansatz,
case of the equidistant measurement of zero-size dimers we found that current computer resources allows us to
– cf. Figs. 8-9 – the average density notches do not drop study N = 5 systems via numerical diagonalization up
to zero, i.e. the soliton structure is not completely dark. to γ & −0.2 only. By investigating the wave function
The shape of the average density notch depends on the for the last pair of ↓–↑ fermions we, in fact, calculate
measurement scheme: the density notch is sharper with higher order correlation functions. If so, the eigenstate
the zero-size scheme than for the any-size one. Like in corresponding to the yrast state has to be determined
Fig. 9, the distance between the two main peaks around very accurately to avoid significant numerical errors in
the density notch increases with the interaction strength. the final results. The number of basis states that have
This is related to a decrease of accessible momenta in the to be taken into account for fixed γ, proliferates dramat-
yrast state. Suprisingly, the measurement scheme and the ically with the increase of N . Such demand eliminates
interaction strength almost do not affect the shape of the the possibility of analysis for interactions stronger than
11

1.8 N = 5, P = 6π The system is described by the Yang-Gaudin Hamilto-


N = 9, P = 10π
N = 13, P = 14π
nian (1) that can be solved analytically with the help of
1.5 the Bethe ansatz. Since the Hamiltonian is invariant un-
der spatial translation of all particles, one cannot observe
probability density

1.2 any feature of the eigenstates by looking at the reduced


single particle density only. Therefore, by employing the
0.9
π Metropolis algorithm, we performed numerical simula-
phase tions of the measurement of particle positions. Starting
0
with the unpolarized 5+5 particles system in the ground
0.6
state, we investigated the formation of dimers in a wide
range of attraction strength. The analysis showed that
0.3 −π
0 0.25 0.5 0.75 1 the average size of pairs of ↓–↑ fermions becomes smaller
position than the mean dimer separation δ̄ when the effective di-
0 mensionless interaction parameter γ . −1. In such a
0 0.2 0.4 0.6 0.8 1
position regime, the many-body distribution of relative distance
between fermions follows the two-body prediction given
FIG. 11: Average probability density for the last zero-size pair by (21). When the attraction between fermions with op-
of ↓–↑ fermions for different initial number of particles N = posite spins is weak (i.e. when γ & −1), one enters the
N↓ = N↑ = 5, 9, 13. The system of size L = 1 was initially regime where the size of the Cooper-like pairs is larger
prepared in the yrast state with total momentum P = π(N +
than δ̄.
1). The more particles in the system, the narrower the density
notch visible at the center of the plot. The inset displays
The key element of this paper is an analysis of yrast
the corresponding (nearly identical for all N considered here) eigenstates in the context of the anticipated emergence of
average phase distributions of two kinds: facing up (with J = soliton signatures. For this purpose, we studied a partic-
1) and down (J = 0). All the calculations were performed in ular yrast state with total momentum P = π(N + 1)/L.
the presence of weak attraction γ = −0.01. By successive particle detections, we analyzed the wave
function for the last pair of fermions. Here, we decided
to examine two different schemes of initial N − 1 pairs
γ ≈ −0.2. Therefore, we restrict ourselves to γ = −0.01. measurements: either detection of zero-size dimers or
We have decided to apply only the any-size measurement without restriction on measured positions of spin up and
scheme because, for weak interactions, the average size down fermions. The results clearly show the dark soli-
of a ↓–↑ pair is larger than the mean interparticle separa- ton signatures (density notches and phase flips) in the
tion. In the weak coupling regime, in order to create the wave function of the last remaining pair of fermions, for
yrast excitation, we need to break the pair of the quasi- all interaction strengths and for both detection schemes.
momenta with zero real part and translate one of those However, the choice of the detection schemes has an in-
quasi-momenta just above the Fermi surface. The Fermi fluence on the shape of the average probability densities.
momentum increases with N , hence, the yrast excitation Surprisingly, the interaction strength and the detection
requires the "injection" of a larger momentum for more schemes almost does not affect the shape of the average
particles in the system. It is consistent with the results phase distributions, revealing a conspicuous phase flip.
presented in Fig. 11 where the width of the density notch Such a resistance of the phase flip to parameter changes
decreases with increasing N . We also observe the phase resembles the behavior observed for a Bose gas described
flips with winding numbers J = 0 and J = 1. by the Lieb-Liniger model [38, 39, 41].

VII. CONCLUSIONS Acknowledgments

We have considered a one-dimensional two-component Support of the National Science Centre, Poland
gas of ultracold fermions interacting via an attractive via projects No.2016/21/B/ST2/01095 (A.S.) and No.
Dirac-delta potential with periodic boundary conditions. 2015/19/B/ST2/01028 (K.S.) is acknowledged.

[1] Y. S. Kivshar and G. P. Agrawal, Optical Solitons, Aca- stock, A. Sanpera, G. V. Shlyapnikov, and M. Lewen-
demic Press, An imprint of Elsevier Science, San Diego, stein, Phys. Rev. Lett. 83, 5198 (1999).
California, 2003. [4] J. Denschlag, J. E. Simsarian, D. L. Feder, Charles W.
[2] C. Pethick and H. Smith, Bose-Eistein condensation in Clark, L. A. Collins, J. Cubizolles, L. Deng, E. W. Ha-
dilute gases (Cambridge University Press, Cambridge, gley, K. Helmerson, W. P. Reinhardt, S. L. Rolston, B.
England, 2002). I. Schneider, and W. D. Phillips, Science 287, 97 (2000).
[3] S. Burger, K. Bongs, S. Dettmer, W. Ertmer, K. Seng- [5] K. E. Strecker, G. B. Partridge, A. G. Truscott, and R.
12

G. Hulet, Nature 417, 150 (2002). [35] D. Delande and K. Sacha, Phys. Rev. Lett. 112, 040402
[6] L. Khaykovich, F. Schreck, G. Ferrari, T. Bourdel, J. (2014).
Cubizolles, L. D. Carr, Y. Castin, and C. Salomon, Sci- [36] S. Krönke and P. Schmelcher, Phys. Rev. A 91, 053614
ence 296, 1290 (2002). (2015).
[7] C. Becker, S. Stellmer, P. Soltan-Panahi, S. Dörscher, [37] I. Hans, J. Stockhofe, and P. Schmelcher, Phys. Rev. A
M. Baumert, E.-M. Richter, J. Kronjäger, K. Bongs, 92, 013627 (2015).
and K. Sengstock, Nature Physics 4, 496 (2008). [38] A. Syrwid and K. Sacha, Phys. Rev. A 92, 032110
[8] S. Stellmer, C. Becker, P. Soltan-Panahi, E.-M. Richter, (2015).
S. Dörscher, M. Baumert, J. Kronjäger, K. Bongs, and [39] A. Syrwid, M. Brewczyk, M. Gajda, and K. Sacha,
K. Sengstock, Phys. Rev. Lett. 101, 120406 (2008). Phys. Rev. A 94, 023623 (2016).
[9] A. Weller, J. P. Ronzheimer, C. Gross, J. Esteve, M. K. [40] A. L. Marchant, T. P. Billam, M. M. H. Yu, A. Rakon-
Oberthaler, D. J. Frantzeskakis, G. Theocharis, and P. jac, J. L. Helm, J. Polo, C. Weiss, S. A. Gardiner, and
G. Kevrekidis, Phys. Rev. Lett. 101, 130401 (2008). S. L. Cornish, Phys. Rev. A 93, 021604(R) (2016).
[10] G. Theocharis, A. Weller, J. P. Ronzheimer, C. [41] A. Syrwid and K. Sacha, Phys. Rev. A 96, 043602
Gross, M. K. Oberthaler, P. G. Kevrekidis, and D. J. (2017).
Frantzeskakis, Phys. Rev. A 81, 063604 (2010). [42] R. Ołdziejewski, W. Górecki, K. Pawłowski, and K.
[11] K. Gawryluk, M. Brewczyk, M. Gajda, and J. Rzążewski, preprint arXiv:1803.11042 (2018).
Mostowski, J. Phys. B: At. Mol. Opt. Phys. 39, L1-L7 [43] J. Dziarmaga and K. Sacha, preprint arXiv:cond-
(2006). mat/0407585 (2004).
[12] T. Pawlowski and K. Rzążewski, New J. Phys. 17, [44] J. Dziarmaga and K. Sacha, Laser Phys. 15, 674 (2005).
105006 (2015) [45] M. Antezza, F. Dalfovo, L. P. Pitaevskii, and S.
[13] A. Boissé, G. Berthet, L. Fouché, G. Salomon, S. As- Stringari, Phys. Rev. A 76, 043610 (2007).
pect, S. Lepoutre, and T. Bourdel, EPL 117, 10007 [46] P. Pieri and G. C. Strinati, Phys. Rev. Lett. 91, 030401
(2017). (2003).
[14] Y. Lai and H. A. Haus, Phys. Rev. A 40, 844 (1989). [47] Ł. Dobrek, M. Gajda, M. Lewenstein, K. Sengstock,
[15] Y. Lai and H. A. Haus, Phys. Rev. A 40, 854 (1989). G. Birkl, and W. Ertmer Phys. Rev. A 60, R3381(R)
[16] J. F. Corney, P. D. Drummond, and A. Liebman, Opt. (1999).
Commun. 140, 211 (1997). [48] G. Andrelczyk, M. Brewczyk, Ł. Dobrek, M. Gajda, and
[17] Y. Castin, in Les Houches Session LXXII, Coherent M. Lewenstein Phys. Rev. A 64, 043601 (2001).
atomic matter waves 1999, edited by R. Kaiser, C. West- [49] L. D. Carr, J. Brand, S. Burger, and A. Sanpera, Phys.
brook and F. David, (Springer-Verlag Berlin Heilder- Rev. A 63, 051601 (2001).
berg New York 2001). [50] T. Karpiuk, M. Brewczyk, Ł. Dobrek, M. A. Baranov,
[18] J. F. Corney and P. D. Drummond, J. Opt. Soc. Am. B M. Lewenstein, and K. Rzążewski, Phys. Rev. A 66,
18, 153 (2001). 023612 (2002).
[19] J. Dziarmaga, and K. Sacha, Phys. Rev. A 66, 043620 [51] T. Karpiuk, M. Brewczyk, and K. Rzążewski, J. Phys.
(2002). B: At. Mol. Opt. Phys. 35, L315-L321 (2002).
[20] C. K. Law, Phys. Rev. A 68 015602 (2003). [52] T. Yefsah, A. T. Sommer, M. J. H. Ku, L. W. Cheuk,
[21] J. Dziarmaga, Z. P. Karkuszewski, and K. Sacha, J. W. Ji, W. S. Bakr, and M. Zwierlein, Nature 499, 426
Phys. B, 36, 1217 (2003). (2013).
[22] J. Dziarmaga, Phys. Rev A. 70, 063616 (2004). [53] A. Bulgac, M. M. Forbes, M. M. Kelley, K. J. Roche,
[23] J. Dziarmaga and K. Sacha, J. Phys. B, 39, 57 (2006). and G. Wlazłowski, Phys. Rev. Lett. 112, 025301
[24] A. I. Streltsov, O. E. Alon, and L. S. Cederbaum, Phys. (2014).
Rev. Lett. 100, 130401 (2008). [54] P. Scherpelz, K. Padavić, A. Rançon, A. Glatz, I. S.
[25] C. Weiss and Y. Castin, Phys. Rev. Lett. 102, 010403 Aranson, and K. Levin, Phys. Rev. Lett. 113, 125301
(2009). (2014).
[26] R. V. Mishmash and L. D. Carr, Phys. Rev. Lett. 103, [55] M. J. H. Ku, W. Ji, B. Mukherjee, E. Guardado-
140403 (2009). Sanchez, L. W. Cheuk, T. Yefsah, and M. W. Zwierlein,
[27] R. V. Mishmash, I. Danshita, Charles W. Clark, and L. Phys. Rev. Lett. 113, 065301 (2014)
D. Carr, Phys. Rev. A 80, 053612 (2009). [56] K. Sacha and D. Delande, Phys. Rev. A 90, 021604(R)
[28] K. Sacha, C. A. Müller, D. Delande, and J. Zakrzewski, (2014).
Phys. Rev. Lett. 103, 210402 (2009). [57] H. Bethe, Z. Physik 71, 205 (1931).
[29] J. Dziarmaga, P. Deuar, and K. Sacha, Phys. Rev. Lett. [58] V. E. Korepin, N. M. Bogoliubov, and A. G. Izer-
105, 018903 (2010). gin, Quantum Inverse Scattering Method and Corre-
[30] R. V. Mishmash and L. D. Carr, Phys. Rev. Lett. 105, lation Functions (Cambridge University Press, Cam-
018904 (2010). bridge, 1993).
[31] A. D. Martin and J. Ruostekoski, New J. Phys. 12, [59] M. Gaudin, The Bethe wavefunction, Cambridge Uni-
055018 (2010). versity Press, 2014.
[32] B. Gertjerenken, T. P. Billam, L. Khaykovich, and C. [60] N. Oelkers, M. T. Batchelor, M. Bortz, and X. W. Guan,
Weiss, Phys. Rev. A 86, 033608 (2012). J. Phys. A: Math. Gen. 39, 1073 (2006).
[33] B. Gertjerenken, T. P. Billam, C. L. Blackley, C. Ruth [61] E. H. Lieb and W. Liniger, Phys. Rev. 130, 1605 (1963).
Le Sueur, L. Khaykovich, S. L. Cornish, and C. Weiss, [62] E. H. Lieb, Phys. Rev. 130, 1616 (1963).
Phys. Rev. Lett. 111, 100406 (2013). [63] C. N. Yang, Phys. Rev. Lett. 19, 1312 (1967).
[34] D. Delande, K. Sacha, M. Płodzień, S. K. Avazbaev, [64] M. Gaudin, Phys. Lett. 24A, 55 (1967).
and J. Zakrzewski, New J. Phys. 15, 045021 (2013). [65] J. N. Fuchs, A. Recati, and W. Zwerger, Phys. Rev.
13

Lett. 93, 090408 (2004). [92] T. Sowiński, M. Gajda, and K. Rzążewski, EPL 109,
[66] I. V. Tokatly, Phys. Rev. Lett. 93, 090405 (2004). 265005 (2015).
[67] C. Mora, A. Komnik, R. Egger, and A. O. Gogolin, [93] C. Recher, Dissertation: Highly imbalanced Fermion-
Phys. Rev. Lett. 95, 080403 (2005). Fermion mixtures in one dimension, Duisburg Essen
[68] T. Iida and M. Wadati, J. Phys. Soc. Jpn. 74, 1724 Universitätsbibliothek, Duisburg-Essen (2013).
(2005). [94] C. Recher, H. Kohler, preprint arXiv:1306.6377 (2013).
[69] M. T. Batchelor et al., J. Phys. Conf. Ser. 42, 5 (2006). [95] J. Javanainen and S. M. Yoo, Phys. Rev. Lett. 76, 161
[70] G. Orso, Phys. Rev. Lett. 98, 070402 (2007). (1996).
[71] H. Hu, X.-J. Liu, and P. D. Drummond, Phys. Rev. [96] D. Dagnino, N. Barberán, and M. Lewenstein, Phys.
Lett. 98, 070403 (2007). Rev. A 80, 053611 (2009).
[72] X.-W. Guan, M. T. Batchelor, C. Lee, and M. Bortz, [97] K. Sakmann and M. Kasevich, Nature Physics 12, 451-
Phys. Rev. B 76, 085120 (2007). 454 (2016).
[73] S. Chen, X.-W. Guan, X. Yin, L. Guan, and M. T. [98] M. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth,
Batchelor, Phys. Rev. A 81, 031608(R) (2010). A. H. Teller, and E. Teller, J. Chem. Phys. 21, 1087
[74] X.-W. Guan, M. T. Batchelor, C. Lee, Rev. Mod. Phys. (1953).
85, 1633 (2013). [99] M. Gajda, J. Mostowski, T. Sowiński, and M. Załuska-
[75] S. S. Shamailov, J, Brand, New J. Phys. 18, 075004 Kotur, EPL 115, 20012 (2016).
(2016) [100] B.Sutherland, Beautiful Models: 70 Years of Exactly
[76] P. P. Kulish, S. V. Manakov, and L. D. Faddeev, Theor. Solved Quantum Many-Body Problems (World Scien-
Math. Phys. 28, 615 (1976). tific, Singapore, 2004).
[77] M. Ishikawa and H. Takayama, J. Phys. Soc. Jpn. 49, [101] W. Ketterle and M. W. Zwierlein, Riv Nuovo Cimento
1242 (1980). 31, 247 (2008).
[78] S. Komineas and N. Papanicolaou, Phys. Rev. Lett. 89, [102] J. B. McGuire, J. Math. Phys. 5, 622 (1964).
070402 (2002). [103] Y. Castin, C. Herzog, C. R. Acad. Sci. Paris 2, série IV,
[79] A. D. Jackson and G. M. Kavoulakis, Phys. Rev. Lett. pp. 419-443 (2001).
89, 070403 (2002). [104] B. Damski, K. Sacha, and J. Zakrzewski, J. Phys. H:
[80] R. Kanamoto, L. D. Carr, and M. Ueda, Phys. Rev. At. Mol. Opt. Phys. 35, L153-L159 (2002).
Lett. 100, 060401 (2008). [105] F. H. Mies, E. Tiesinga, and P. S. Julienne, Phys. Rev.
[81] R. Kanamoto, L. D. Carr, and M. Ueda, Phys. Rev. A 61, 022721 (2000)
A 81, 023625 (2010); Erratum Phys. Rev. A 81, [106] F. A. van Abeelen and B. J. Verhaar, Phys. Rev. Lett.
049903(E) (2010). 83, 1550 (1999).
[82] T. Karpiuk, P. Deuar, P. Bienias, E. Witkowska, K. [107] V. A. Yurovsky, A. Ben-Reuven, P. S. Julienne, and C.
Pawłowski, M. Gajda, K. Rzążewski, and M. Brewczyk, J. Williams, Phys. Rev. A 60, R765(R) (1999.)
Phys. Rev. Lett. 109, 205302 (2012). [108] C. A. Regal, C. Ticknor, J. L. Bohn, and D. S. Jin,
[83] T. Karpiuk, T. Sowiński, M. Gajda, K. Rzążewski, and Nature 424, 47 (2003).
M. Brewczyk, Phys. Rev. A 91, 013621 (2015). [109] J. Cubizolles, T. Bourdel, S. J. J. M. F. Kokkelmans,
[84] J. Sato, R. Kanamoto, E. Kaminishi, and T. Deguchi, G. V. Shlyapnikov, and C. Salomon, Phys. Rev. Lett.
Phys. Rev. Lett. 108, 110401 (2012). 91, 240401 (2003)
[85] J. Sato, R. Kanamoto, E. Kaminishi, and T. Deguchi, [110] S. Jochim, M. Bartenstein, A. Altmeyer, G. Hendl, C.
preprint arXiv:1204.3960. Chin, J. H. Denschlag, and R. Grimm, Phys. Rev. Lett.
[86] J. Sato, R. Kanamoto, E. Kaminishi, and T. Deguchi, 91, 240402 (2003).
New J. Phys. 18, 075008 (2016). [111] K. E. Strecker, G. B. Partridge, and R. G. Hulet, Phys.
[87] K. Gawryluk, M. Brewczyk, and K. Rzążewski, Phys. Rev. Lett. 91, 080406 (2003).
Rev. A 95, 043612 (2017). [112] R. B. Diener and T.-L. Ho, preprint arXiv:cond-
[88] G. E. Astrakharchik, D. Blume, S. Giorgini, and B. E. mat/0404517 (2004).
Granger, Phys. Rev. Lett. 92, 030402 (2004). [113] A. Perali, P. Pieri, and G.C. Strinati, Phys. Rev. Lett.
[89] G. E. Astrakharchik, J. Boronat, J. Casulleras, and S. 95, 010407 (2005).
Giorgini, Phys. Rev. Lett. 95, 190407 (2005). [114] E. Altman and A. Vishwanath, Phys. Rev. Lett. 95,
[90] M. T. Batchelor, M. Bortz, S.-W. Guan, and N. Oelkers 110404 (2005).
J. Stat. Mech. 2005, L10001. [115] E. A. Yuzbashyan, B. L. Altshuler, V. B. Kuznetsov,
[91] E. Haller, M. Gustavsson, M. J. Mark, J. G. Danzl, R. and V. Z. Enolskii, Phys. Rev. B 72, 220503 (2005).
Hart, G. Pupillo, and H. C. Nägerl, Science 325, 1224
(2009).

You might also like