• Non ci sono risultati.

Dynamical quantum phase transitions in systems with continuous symmetry breaking

N/A
N/A
Protected

Academic year: 2021

Condividi "Dynamical quantum phase transitions in systems with continuous symmetry breaking"

Copied!
8
0
0

Testo completo

(1)

Simon A. Weidinger,1,∗ Markus Heyl,2, 1 Alessandro Silva,3 and Michael Knap1

1

Department of Physics and Institute for Advanced Study, Technical University of Munich, 85748 Garching, Germany

2

Max-Planck-Institut f¨ur Physik komplexer Systeme, 01187 Dresden, Germany

3

SISSA - International School for Advanced Studies, via Bonomea 265, 34136 Trieste, Italy (Dated: September 5, 2017)

Interacting many-body systems that are driven far away from equilibrium can exhibit phase transitions be-tween dynamically emerging quantum phases, which manifest as singularities in the Loschmidt echo. Whether and under which conditions such dynamical transitions occur in higher-dimensional systems with spontaneously broken continuous symmetries is largely elusive thus far. Here, we study the dynamics of the Loschmidt echo in the three dimensional O(N) model following a quantum quench from a symmetry breaking initial state. The O(N) model exhibits a dynamical transition in the asymptotic steady state, separating two phases with a finite and vanishing order parameter, that is associated with the broken symmetry. We analytically calculate the rate function of the Loschmidt echo and find that it exhibits periodic kink singularities when this dynamical steady-state transition is crossed. The singularities arise exactly at the zero-crossings of the oscillating order parameter. As a consequence, the appearance of the kink singularities in the transient dynamics is directly linked to a dy-namical transition in the order parameter. Furthermore, we argue, that our results for dydy-namical quantum phase transitions in the O(N) model are general and apply to generic systems with continuous symmetry breaking.

I. INTRODUCTION

In recent years, synthetic quantum matter such as ultra-cold atoms, polar molecules, and trapped ions have demon-strated their capabilities to experimentally study nonequi-librium quantum states far beyond the regime of linear re-sponse and thus far beyond a thermodynamic description. Due to the isolation from the environment and the high level of control, experiments with synthetic quantum matter have shown that inherently dynamical phenomena can be real-ized and probed, ranging from many-body localization,1–7,

prethermalization,8,9 discrete time crystals,10,11 the particle-antiparticle production in the Schwinger model,12 to emer-gent Bloch oscillations.13 In addition, not only the

dynam-ical phases themselves have become accessible in experi-ments, but also the associated dynamical transitions between the phases.14–16

Current experimental platforms for studying dynamics are often focusing on one- and two-dimensional systems. Yet, a future prospect concerns extensions toward the realization of non-equilibrium many-body states in three spatial dimensions, where new physical phenomena become accessible. This in-cludes, for example, the possibility of spontaneously broken continuous symmetries at nonzero temperatures, which is ex-cluded for lower dimensions due to the Mermin-Wagner the-orem in systems with short range interactions.

In this work, we study the quantum dynamics of an in-teracting many-body system in three dimensions which ex-hibits such a spontaneously broken symmetry. Specifically, we calculate the unitary real-time evolution of the O(N) vec-tor model following a quantum quench of the mass, with an initial state that breaks the continuous symmetry of our sys-tem, see Fig.1. We approach the problem fully analytically via the large-N limit, where the dynamics can be solved ex-actly.

The O(N) model exhibits a dynamical quantum phase tran-sition in the asymptotic steady state, separating two

dynami-bare mass (a)

(b) (c)

FIG. 1. Dynamical criticality in the Loschmidt echo for systems with spontaneous symmetry breaking. (a) We study dynamical quantum phase transitions of the O(N) model following quantum quenches from an initial bare mass r0

i to a final bare mass rf0. The

initial state is chosen to break the continuous symmetry of the O(N) model and hence is described by a finite order parameter. Our system exhibits a steady-state dynamical phase transition at rdyn

c , which

sep-arates the dynamically ordered phase in which the long-time average of the order parameter ¯φ remains finite, from the disordered phase in which ¯φ vanishes. We analytically calculate the Loschmidt echo and find that the associated rate function remains smooth for quenches within the dynamically ordered phase (b) but exhibits nonanalytic kink singularities when crossing the dynamical critical point rcdyn(c).

cal phases with finite and vanishing order parameter, respec-tively.17 Here, we show that in addition to the dynamical steady-state transition of the order parameter, the O(N) model exhibits a critical dynamical phenomenon on transient time scales. In particular, non-analyticities appear in the Loschmidt echo periodically in time when the dynamical transition is crossed by the quantum quench (Fig.1). We show that in the O(N) model these singularities contribute only subexten-sively to the rate function associated with the Loschmidt echo. Making use of the analogy between the Loschmidt echo and the boundary partition function, this effect is reminiscent of

(2)

surface phase transitions in equilibrium systems, which also contribute only subextensively to the free energy.18

Further-more, we find that the dynamical critical point obtained from the order parameter coincides with the one obtained from the Loschmidt echo. These different concepts of dynamical criti-cality are further linked by the fact that the non-analyticities in the Loschmidt echo occur at times when the order parameter crosses zero. A similar relation has been found in the long-range transverse-field Ising model.19We argue that our results

are not specific to the O(N) model or the large-N limit, and hence apply to generic systems with a spontaneously broken continuous symmetry.

This work is organized as follows. In Sec.II we discuss two different notions of dynamical quantum phase transitions; one associated with the time evolution of the order parame-ter, and the other one exploiting the similarity between the Loschmidt echo and a classical partition function. In Sec.III

we review the far from equilibrium dynamics of the O(N ) model in the large N approximation to leading order. The time evolved state of the O(N ) model is calculated in Sec.IV A. In Sec.IV Bwe derive the return probability of the time evolved state to the ground state manifold and in Sec.IV Cwe analyze the corresponding rate function and show that it exhibits non-analyticities, when the model is quenched across the dynami-cal critidynami-cal point. Finally, in Sec.Vwe conclude our findings and discuss potential extensions of our work.

II. DYNAMICAL QUANTUM PHASE TRANSITIONS

We investigate two notions of dynamical quantum phase transitions. The first one, is associated with the time evolution of the order parameter.17,19–26 The dynamical quantum phase

transition is then characterized by a critical point which sepa-rates regimes where the long-time average of the order param-eter ¯φ is either finite or zero. Close to this dynamical critical point the long-time average ¯φ exhibits scaling relations with critical exponents.17,26 However, the location of the

dynami-cal critidynami-cal point can in general differ from the equilibrium one and might also depend on the initial state, due to a dynamical renormalization of parameters.17,26

A second approach to study the nonequilibrium dynamical criticality is to exploit the formal similarity between the equi-librium partition function Z = tr[e−β ˆH] and the Loschmidt amplitude hψ0|e−i ˆHt|ψ0i.27,28The equilibrium partition

func-tion becomes non-analytic at a convenfunc-tional phase transifunc-tion as a function of the control parameter such as temperature or pressure. It turns out, that the Loschmidt amplitude can also exhibit nonanalyticities, but as a function of time rather than a control parameter. Indeed it has been shown that the rate function, which is obtained from taking the logarithm of the Loschmidt amplitude, exhibits nonanalyticities when the system is quenched across a quantum critical point whereas it remains smooth for quenches within the same dynamical phase.19,27–38. Recently, it became also possible to measure Loschmidt amplitudes in various experimental settings.14,39

So far the Loschmidt amplitude has mostly been studied for one dimensional systems with discrete Z2 symmetries (see,

however, Refs.32,34, and38). In this work, we look at a three dimensional model with a continuous O(N) symmetry: the O(N) vector model. This model provides a universal de-scription for many systems close to their critical point and is well established in the study of (non-equilibrium) quan-tum phase transitions.17,26,40–43 For example, the equilibrium Mott-insulator to superfluid transition in the Bose-Hubbard model falls into the universality class of the O(2) model and the Heisenberg antiferromagnet can be described by an O(3) model.

We propose the following generalization of the Loschmidt echo to systems with a continuously broken symmetry

L(t) = Z

{|χ|=φ0}

dNχ |hχ|Ψ(t)i|2. (1)

Here, |Ψ(t)i = ˆU (t)|ψ0i is the time evolved state after the quench and the integral is taken over the full set of symmetry-broken ground states |χi, which can be pictured as a sphere within an N -dimensional space. The radius φ0 is set by the order parameter in the initial state. Below we will analyze the dynamics of the rate function associated with the Loschmidt echo

R(t) = − 1

LdN log L(t), (2)

which shows nonanalytic behavior for quantum quenches from the dynamically ordered to the disordered phase.

III. THE O(N) MODEL FAR FROM EQUILIBRIUM

The quantum O(N) model consists of N real scalar fields ˆ

Φa, a = 1, . . . , N and conjugate momenta ˆΠa in d spatial dimensions. The corresponding Hamiltonian is

ˆ H = Z x  1 2 ˆ Π2a+1 2(∇ ˆΦa) 2+r0 2 ˆ Φ2a+ λ 4!N( ˆΦa ˆ Φa)2  , (3) where r0is the square of the bare mass and λ is the interaction strength. The fields obey the canonical commutation relation [ ˆΦa(x), ˆΠb(x0)] = iδabδ(x − x0). We assume, that repeated indices are summed over.

In the following, we consider the limit of infinitely many scalar fields, N → ∞. In that limit, the interaction of strength λ solely renormalizes the bare mass r0as follows

r = r0+ λ 6Nh ˆΦ

2

ai. (4)

The large-N approximation relies on the factorization of the expectation value hΦˆaΦˆa

N Φbi = hˆ ˆ ΦaΦˆa

N ih ˆΦbi + O(1/N ) to leading order in 1/N .40 Therefore, there are no interactions

between excitations and the model possesses an infinite num-ber of conserved quantities and is non-ergodic.41 As a con-sequence it does not thermalize. Only next-to-leading order terms introduce scattering between quasi-particle excitations and may ultimately enable thermalization.44–46 In the present

(3)

FIG. 2. Dynamical phase diagram of the O(N) model in three spa-tial dimensions. The system is prepared in the equilibrium symme-try broken phase at zero temperature. For quenches to a point inside of the dynamically symmetry-broken phase, rf0< rdync , the order

pa-rameter relaxes to a finite value ¯φ (red line) and the effective mass rf remains zero, indicating the presence of gapless excitations in the

steady state. For quenches into the symmetric phase, r0f > rcdyn, the

long-time average of the order parameter is zero ¯φ = 0 and the effec-tive mass rf becomes finite (blue dashed line). Close to the critical

point rdyn

c the long-time average ¯φ vanishes as (rcdyn− rf0)1/4and the

effective mass as rf∼ (r0f − r dyn c ).17

work we are not interested in the late-time thermalization physics, but rather in the transient prethermal regime after the quench; accordingly a leading order analysis is sufficient.

In equilibrium, the O(N) model hosts two different phases: a disordered phase with finite effective mass r > 0 and an ordered phase, in which the system spontaneously breaks the continuous O(N) symmetry by developing a finite order pa-rameter hΦi 6= 0. In the ordered phase the mass gap van-ishes r = 0. The equilibrium critical point is given by reqc = −12λ

R p

1

|p|, which is finite for d > 1. In d > 2 the ordered phase extends to finite temperatures. In the rest of the paper we will focus on three spatial dimensions, d = 3.

Let us assume in the following, that the system has been prepared in the symmetry-broken ground state |Ψ0i at r0

i, with the order parameter h ˆΦai = δ1,aφ0pointing along the a = 1 direction. The value of φ0is given by

(φ0)2= −6r 0 i λ − (N − 1)h ˆΦ2 ˆ Φ2i, (5)

which follows directly from the initial mass being zero. Here, we also used, that there is a remaining O(N − 1)-symmetry for the a ≥ 2 components. We then suddenly change the mass to the final value r0= rf0and let the system evolve in time. If the final value r0f is smaller than the dynamical critical value rdyn

c , the system reaches an ordered steady state characterized by rf= 0 and ¯φ = limT →∞T1

R∞

0 dtφ(t) > 0.

17On the other

hand, if r0f > rcdynthe order is melted. Therefore, the effective mass rf > 0 and the order parameter ¯φ = 0, as illustrated in the dynamical phase diagram for d = 3 in Fig.2.

To obtain the equations of motion at N → ∞, we treat the a = 1 component of the field as a classical variable, ˆ

Φ1(t) → φ(t) ∈ R, and expand the a ≥ 2 components into creation and annihilation operators that diagonalize the initial Hamiltonian26 ˆ Φa≥2(p, t) = fp(t)ˆb(a)p + fp∗(t)ˆb (a)† −p , (6) where ˆΦa(x, t) = V−1/2P pΦa(p, t)eˆ

ipx. Note, that due to the O(N − 1) symmetry of the remaining a ≥ 2 components, the time dependence is identical for all of them and hence the mode functions fp(t) in Eq. (6) do not carry a field component index.

Using the Heisenberg equations of motions, we obtain ¨

fp(t) + [p2+ r(t)]fp(t) = 0 ¨

φ(t) + r(t)φ(t) = 0, (7)

with the time-dependent effective mass

r(t) = rf0+ λ 6N  φ2(t) + (N − 1) Z p |fp(t)|2  . (8)

It is important to notice, that φ(t) ∼ √N . Therefore, both terms in the parenthesis in Eq. (8) scale linearly with N and contribute to the effective mass.

The initial conditions of Eq.’s (7) are fp(0) = 1/p2|p|, ˙

fp(0) = −ip|p|/2, which follow from requiring that ˆbp, ˆb†p diagonalize the initial Hamiltonian and r(t = 0) = 0. Fur-thermore we have φ(0) = φ0 and ˙φ(0) = 0, with φ0 given by Eq. (5). To regularize the infrared divergence of fp(0), we introduce an cut-off p0 = 2π/L, with L being the linear ex-tension of the system. This amounts to placing the field theory in a finite box with volume Ld. Eventual UV divergencies are regularized with a finite cut-off Λ in momentum space.

IV. RESULTS A. Time evolved state

In order to calculate the return probability to the ground-state manifold, we need to know the time evolved ground-state |Ψ(t)i = ˆU (t)|Ψ0i. In the N → ∞ limit the state |Ψ(t)i fac-torizes in the field components due to the effectively quadratic Hamiltonian at leading order.26In the a ≥ 2 components there

is a squeezed state |ψsq(t)i and in the ”classical” a = 1 com-ponent a coherent state |φ(t)i ,

|Ψ(t)i = |φ(t)i ⊗ |ψsq(t)i |φ(t)i = e−12γ 2φ2(t) eγφ(t)ˆb(1)†p0 |0i |ψsq(t)i =Y p>0 a≥2 1 p|αp(t)|exp  β∗ p(t) 2α∗ p(t) (ˆb(a)†p )2  |0i, (9)

(4)

where αp(t) = fp(t) q |p| 2 + i ˙ fp(t) √ 2|p|, βp(t) = fp(t) q |p| 2 − if˙p(t) 2|p| and γ = L d−1 2 √ dπ 2 12

. The coherent state contribu-tion gives rise to a finite order parameter hΨ(t)| ˆΦ1|Ψ(t)i = φ(t).

B. Return probability to the groundstate manifold

An arbitrary state in the groundstate manifold of (3) in the symmetry-broken phase can be written as

|χi = e−12γ 2χ2 eγχTˆb†p0|0i, (10) where χ = (χ1, . . . , χN), |χ|2 = φ20 and ˆbp = (ˆb(1)p , . . . , ˆb (N )

p ). The expectation of the field-operator in this state is given by hχ| ˆΦa|χi = χa. The overlap hχ|Ψ(t)i fac-torizes into a product over the field components. For a = 1 we get a scalar product of two coherent states and for a ≥ 2 we have scalar products of a coherent and a squeezed state, which we calculate by expanding the exponentials. For the return probability to a specific initial state, we obtain

|hχ|Ψ(t)i|2= exp ( −LdN Z p log |αp(t)| − Ld−1 √ dπ 2 [φ 2 (t) + φ2(0) − 2χ1φ(t) +X a≥2 χ2a] ) . (11) In deriving this formula we also made use of the fact, that for large systems, L  1, i.e., small p0 = 2π/L, the ratio βp0(t)/αp0(t) approaches 1.

The overlap |hχ|Ψ(t)i|2 is rotational invariant around the a = 1 axis. Hence, we use spherical coordinates (see Fig.3) to calculate the integral over the groundstate manifold as re-quired in Eq. (1). Defining θ ∈ [0, π] as the angle between the vector χ and the a = 1 axis, i.e., cos θ = χ1/φ0, we can write

|hχ|Ψ(t)i|2= N sq(t) exp ( −Ld−1 √ dπ 2 φ 2(0) × " 1 + φ(t) φ(0) 2 + sin2θ − 2 φ(t) φ(0)  cos θ #) . (12)

Here, we introduced the abbreviation Nsq(t) = exp[−LdNR

plog |αp(t)|]. The integration element can be written as dNχ δ(|χ| − φ0) = SN −1(φ0sin θ)φ0dθ/SN(φ0), where Sn(r) = 2πn

2Γ(n/2)−1rn−1 is the surface of the

n-sphere (see Fig.3for a graphical interpretation). Exponen-tiating the sin θ - term, we obtain the return probability to the ground state manifold

L(t) = A Nsq(t) π Z

0

dθe−Ld−1N F (θ,φ(t)), (13)

FIG. 3. Definition of the angle θ and surface of element of the N-dimensional sphere. The ground state manifold of the O(N ) model can be pictured as a sphere with radius φ0in a N -dimensional space.

The return probability L(t) to the ground state manifold is obtained from the integral of the overlap |hχ|Ψ(t)i|2over this sphere. Defin-ing θ as the angle between the vector χ and the initial order parameter (φ0, 0, . . . , 0) and making use of the rotational symmetry around the

(a = 1)-axis, one can write the integration element dNχ δ(|χ| − φ0)

as the product of the arc length φ0dθ and the surface area of the

sphere in N − 1 dimensions S(N −1)

0sin θ) generated by rotating

χ around the a = 1 - axis with θ fixed. To obtain a probability mea-sure, we finally divide the integration element by the total available surface area SN(φ0). with F (θ, φ) =r π 2 φ2 0 N " 1 + φ φ0 2 − 2 φ φ0  cos θ + sin2θ # −N − 2 N L −d+1log sin θ (14) and a constant A = π−1/2Γ(N 2)/Γ( N −1 2 ). We will refer to L(t) also as Loschmidt Echo. Eq. (13) can be interpreted as a classical partition function of the angular variable θ mov-ing in an order parameter landscape F (θ, φ(t)), with Ld−1N playing the role of inverse temperature. The energy land-scape, Eq. (14), has the shape of a double well potential, where the order parameter φ is acting as an external field tilt-ing the two wells against each other, see Fig.4. The two wells are energetically equivalent, when the external field vanishes (φ(t) = 0).The larger L, the more the two wells move out-wards to 0 and π . Nevertheless, the log sin θ term is im-portant, because it is responsible for creating the double well landscape.

In the thermodynamic limit L  1, we can evaluate the in-tegral in Eq. (13) using a saddle point approximation. Taking this limit corresponds to very low temperatures in the classi-cal partition function and the variable θ will pick the minimum energy well L(t) ' L1Nsq(t) exp  −Ld−1N min θ∈[0,π]F (θ, φ(t))  . (15)

(5)

(a) (b) (c)

FIG. 4. Order parameter landscape for the angle θ: The return probability, Eq. (13), can be interpreted as a classical partition function for the variable θ ∈ [0, π] moving in an effective free energy landscape F (θ, φ). The landscape has the form of a double well potential, where the order parameter φ(t) is acting as an external field, shifting the two wells against each other. For (a) φ(t) > 0 the left minimum is energetically more favorable, while for (c) φ(t) < 0 the situation is reversed. As the system size L is increased, the left (right) potential well is shifted toward θ = 0 (θ = π), which correspond to the states having an order parameter parallel (anti-parallel) to the initial state. (b) When φ(t) changes sign the most relevant value of θ jumps from one well to the other, which gives rise to the kinks in the Loschmidt rate function R(t).

For L → ∞, the last term in Eq. (14) vanishes and the min-imum is at θmin = 0 (θmin = π) for φ(t) > 0 (φ(t) < 0), meaning that χ is parallel (antiparallel) to the order parameter of the initial state. Therefore, only two states from the con-tinuous ground state manifold contribute significantly to the Loschmidt echo: L(t) ∼ (|h+φ0|Ψ(t)i|2+ |h−φ0|Ψ(t)i|2). This can be interpreted as follows: the order-parameter os-cillates only along a fixed axis due to the symmetry of the Hamiltonian and cannot explore the whole ground state man-ifold.

Our result for the coherent state contribution to the Loschmidt rate function scales subextensively with system size as ∼ Ld−1, see the prefactor of F (θ, φ(t)) in Eq. (15). This is a consequence of the infrared divergence of the initial mode function fp(0) due to the spontaneously broken symme-try, which leads to the scaling of γ ∼ L(d−1)/2in the coherent state, Eq. (9). From that, the wavefunction overlap hχ|Ψ(t)i of the time evolved state and an arbitrary state in the ground state manifold contains terms, that scale subextensively ∼ Ld−1. We emphasize that the subextensive scaling shows up only in the wave function overlap but not in expectation values of observables. Examples include the order parameter and the work performed in a quench. The latter shows a normal exten-sive scaling ∼ Ldwith system size. The average work h ˆHfi is given by the expectation value of the post-quench Hamilto-nian in the initial state, h ˆHfi = Ld r0f−r0i

2 (N R p 1 2|p|+ φ 2 0). All higher cumulants of the work distribution function vanish in our leading order approximation. Generally, the logarithm of the Loschmidt amplitude acts as the generating function for cumulants of the work-distribution.47–49 We also find in our model that to leading order in N , the Loschmidt echo repro-duces exactly the cumulants of the work.

C. Rate function

Calculating the rate function R(t) = −L−dN−1log L(t) from Eq.(15), we find that

R(t) = Rsq(t) + 1 LRcoh(t) Rsq(t) = Z p log |αp(t)| Rcoh(t) =r π 2N −1φ2 0 " 1 + φ(t) φ0 2 − 2 φ(t) φ0 # . (16)

The contribution from the squeezed state Rsqis obtained from Nsq, and the coherent state contribution is obtained by explic-itly calculating the minimum in Eq. (15). The rate function Rsqis a smooth function of time, since |αp(t)| is smooth and bounded from below by 1. Rcoh on the other hand exhibits kinks at zero crossings of φ(t) due to the absolute value in the last term of Eq. (16). As discussed above, the coherent state contribution is suppressed by a factor of L−1. However, the squeezed-state part of the rate function Rsq(t) relaxes to a constant value on a much shorter time-scale than the order parameter φ(t), because of an integral over momenta. There-fore, the non-analyticities in Rcoh(t) can be identified for in-stance in the second derivative ¨R(t) of the rate function. For the squeezed state, ¨Rsq(t) ≈ 0, whereas the coherent state retains prominent δ-peaks ¨Rcoh(t) ∼P

Tkinkδ(t − Tkink), as

illustrated in Fig.5.

The coherent state contribution to the Loschmidt rate func-tion Rcoh(t) exhibits kinks at the zero crossings of the order parameter, φ(Tkink) = 0, see Fig.6. From the numerical solu-tion of the equasolu-tions of mosolu-tion (7) we also find that the order parameter relaxes to a non-zero value for quenches inside the dynamical symmetry-broken phase (r0f < rdync ). In this case there are no zero crossings of φ(t) and hence we do not find

(6)

(a) (b)

FIG. 5. Loschmidt rate function. The Loschmidt rate function R(t) of the return probability to the groundstate manifold contains a squeezed state contribution Rsq(t), which scales extensively with system size, and a coherent state contribution Rcoh(t) that scales subextensively.

Whereas Rsq(t) is a smooth function of time, Rcoh(t) shows kinks, when the system is quenched across the dynamical quantum phase

transition. Due to the subextensive scaling of the coherent state contribution Rcoh, the non-analytic behavior is not visible in the full

rate-function R(t) of the return probability, (a). Nevertheless the non-analytic behavior is clearly observable in the second derivative ¨R(t) as δ-peaks, since the squeezed state contribution relaxes on a much shorter time scale than the one of the coherent state, (b). The system parameters are L = 2.5 × 104, λ = 1.0, ri0= −1.0 and r

0 f = 0.0.

FIG. 6. Coherent state contribution to the rate function and or-der parameter dynamics. The coherent state contribution Rcohto

the rate function exhibits kinks at the zero-crossings of the order pa-rameter φ(t). The kinks appear periodically and the time between them ∆Tkink is determined by final effective mass rf: ∆Tkink ∼

(r0 f − rdync )

−1

2. The data is evaluated for the same parameters as in

Fig.5

any non-analyticities in Rcoh. By contrast, for quenches to the symmetric phase (r0f > rdyn

c ), the order parameter os-cillates around zero and approaches ¯φ = 0 and Rcohexhibits kinks. As a consequence, there is an intimate relation between the dynamical phase transition of the order parameter and the kinks in the Loschmidt rate function of the return probability to the groundstate manifold.

Following a quench to the symmetric phase, the effective mass r(t), Eq. (8), attains a finite average value rf, which feeds back into the equations of motion, Eq. (7), as frequency squared of φ(t). Accordingly, the kinks in Rcoh appear at equidistantly spaced times Tkinkand the time ∆Tkinkbetween

two kinks is uniquely determined by rf. The effective mass after a quench to the symmetric phase scales linearly with the distance of the final bare mass r0

f from the dynamical critical point rdyn

c , rf ∼ r0f−r dyn

c , as depicted in Fig.2. We therefore find ∆Tkink= π √ rf ∼ (r 0 f − r dyn c )− 1 2. (17)

Therefore, the time between the kinks diverges with the same critical exponent upon approaching the dynamical critical point as the correlation length in equilibrium, which is a man-ifestation of the O(N) model being a relativistic field theory in which time and space scale in the same way.

V. CONCLUSION AND OUTLOOK

We have studied the rate function of the return probabil-ity to the ground state manifold in the O(N) model following a quantum quench from a symmetry breaking initial state to the symmetric phase. The rate function exhibits kinks, which are located at the zero crossings of the order parameter φ(t) and are equally spaced with a period ∆Tkinkdetermined by the final effective mass. In our model, the non-analytic con-tribution to the return probability scales subextensively with system size. Such a subextensive contribution can also ap-pear in equilibrium whenever a system undergoes a surface or impurity phase transition.

For quenches from the symmetric to the symmetry-broken phase kinks are absent, since the closing of the gap leads to a divergent time scale between kinks. Also, due to the absence of explicit symmetry-breaking terms in the Hamiltonian, no finite order parameter can be ever generated.

Our results for the non-equilibrium dynamics are obtained fully analytically to leading order in the number of compo-nents N of the field theory. We point out that the saddle-point approximation, which we employ in the calculation of

(7)

the return probability, only relies on the thermodynamic limit L → ∞ and not on N being large. Furthermore the presence of kinks in the rate function R(t) hinges on the presence of the coherent state, i.e., a finite order parameter φ(t). Next-to-leading order corrections would modify the time evolution of the order parameter and the quantum fluctuations in the time evolved state, but would not destroy the symmetry-broken phase, i.e. the coherent contribution to the time evolved state. Therefore, we argue, that our results remain valid beyond a leading order approximation in 1/N . Moreover, due to the universality of the O(N) model, we expect our results to be generic for dynamical critical points in models with continu-ous symmetries. In particular, the return probability should be dominated by the states parallel and anti-parallel to the initial state, leading to non-analytic behavior of the rate function for quenches from the symmetry-broken to the symmetric phase.

Moreover, the zero crossings of the order parameter should determine the times at which nonanalyticities appear in the Loschmidt echo. It would be intriguing to explore these find-ings in other models with continuous symmetry breaking.

ACKNOWLEDGMENTS

Acknowledgments.—We acknowledge support from the Deutsche Forschungsgemeinschaft via the Gottfried Wilhelm Leibniz Prize program (MH), the Technical University of Mu-nich - Institute for Advanced Study, funded by the German Excellence Initiative and the European Union FP7 under grant agreement 291763 (SW, MK), and from the DFG grant No. KN 1254/1-1 (SW, MK).

simon.weidinger@tum.de 1

Michael Schreiber, Sean S. Hodgman, Pranjal Bordia, Henrik P. L¨uschen, Mark H. Fischer, Ronen Vosk, Ehud Altman, Ulrich Schneider, and Immanuel Bloch, “Observation of many-body lo-calization of interacting fermions in a quasirandom optical lat-tice,”Science 349, 842–845 (2015).

2

Jae-yoon Choi, Sebsastian Hild, Johannes Zeiher, Peter Schauß, Antonio Rubio-Abadal, Tarik Yefash, Vedika Khemani, David A. Huse, Immanuel Bloc, and Christian Gross, “Exploring the many-body localization transition in two dimensions,” Science 352, 1547–1552 (2016).

3

J. Smith, A. Lee, P. Richerme, B. Neyenhuis, P. W. Hess, P. Hauke, M. Heyl, D. A. Huse, and C. Monroe, “Many-body localization in a quantum simulator with programmable random disorder,”Nat. Phys. 12, 907–911 (2016).

4 Pranjal Bordia, Henrik P. L¨uschen, Sean S. Hodgman, Michael

Schreiber, Immanuel Bloch, and Ulrich Schneider, “Coupling Identical one-dimensional Many-Body Localized Systems,”Phys. Rev. Lett. 116, 140401 (2016).

5

Pranjal Bordia, Henrik P. L¨uschen, Ulrich Schneider, Michael Knap, and Immanuel Bloch, “Periodically driving a many-body localized quantum system,”Nat. Phys. 13, 460–464 (2017).

6

Henrik P. L¨uschen, Pranjal Bordia, Sebastion Scherg, Fabian Alet, Ehud Altman, Ulrich Schneider, and Immanuel Bloch, “Evidence for Griffiths-Type Dynamics near the Many-Body Localization Transition in Quasi-Periodic Systems,”arXiv:1612.07173 (2016).

7

Pranjal Bordia, Henrik L¨uschen, Sebastian Scherg, Sarang Gopalakrishnan, Michael Knap, Ulrich Scheider, and Immanuel Bloch, “Probing Slow Relaxation and Many-Body Localization in Two-Dimensional Quasi-Periodic Systems,”arXiv:1704.03063 (2017).

8 M. Gring, M. Kuhnert, T. Langen, T. Kitagawa, B. Rauer,

M. Schreitl, I. Mazets, D. Adu Smith, E. Demler, and J. Schmied-mayer, “Relaxation and prethermalization in an isolated quantum system,”Science 337, 1318–1322 (2012).

9

Tim Langen, Sebastian Erne, Remi Geiger, Bernhard Rauer, Thomas Schweigler, Maximilian Kuhnert, Wolfgang Rohringer, Igor E. Mazets, Thomas Gasenzer, and J¨org Schmiedmayer, “Ex-perimental observation of a generalized Gibbs ensemble,”Science 348, 207–211 (2015).

10

J. Zhang, P. W. Hess, A. Kyprianidis, P. Becker, A. Lee, J. Smith, G. Pagano, I. D. Potirniche, A. C. Potter, A. Vishwanath, N. Y.

Yao, and C. Monroe, “Observation of a discrete time crystal,”

Nature 543, 217–220 (2017).

11 Soonwon Choi, Joonhee Choi, Renate Landig, Georg Kucsko,

Hengyun Zhou, Junichi Isoya, Fedor Jelezko, Shinobu Onoda, Hi-toshi Sumiya, Vedika Khemani, Curt von Keyserlingk, Norman Y. Yao, Eugene Demler, and Mikhail D. Lukin, “Observation of dicrete time-crystalline order in a disordered dipolar many-body system,”Nature 543, 221–225 (2017).

12 Esteban A. Martinez, Christine A. Muschik, Philip Schindler,

Daniel Nigg, Alexander Erhard, Markus Heyl, Philipp Hauke, Marcello Dalmonte, Thomas Monz, Peter Zoller, and Rainer Blatt, “Real-time dynamics of lattice gauge theories with a few-qubit quantum computer,”Nature 534, 516–519 (2015).

13

Florian Meinert, Michael Knap, Emil Kirilov, Katharina Jag-Lauber, Mikhail B. Zvonarev, Eugene Demler, and Hanns-Christoph N¨agerl, “Bloch oscillations in the absence of a lattice,”

Science 356, 945–948 (2017).

14

C. Jurcevic, H. Shen, P. Hauke, C. Maier, T. Brydges, C. Hempel, B. P. Lanyon, M. Heyl, R. Blatt, and C. F. Roos, “Direct Obser-vation of Dynamical Quantum Phase Transitions in an Interacting Many-Body System,”Phys. Rev. Lett 119, 080501 (2017).

15 Nick Fl¨aschner, Dominik Vogel, Matthias Tarnowski, Benno S.

Rem, Dirk-S¨oeren Luehmann, Markus Heyl, Jan Carl Budich, Ludwig MAthey, Klaus Sengstock, and Christof Weitern-berg, “Observation of a dynamical topological phase transition,”

arXiv:1608.05616 (2016).

16

J. Zhang, G. Pagano, P. W. Hess, A. Kyprianidis, P. Becker, H. Ka-plan, A. V. Gorshkov, Z.-X. Gong, and C. Monroe, “Observa-tion of a Many-body Dynamical Phase Transi“Observa-tion with a 53-Qubit Quantum Simulator,”arXiv:1708.01044 (2017).

17

Bruno Sciolla and Giulio Biroli, “Quantum quenches, dynamical transitions, and off-equilibrium quantum criticality,”Phys. Rev. B 88, 201110(R) (2013).

18

H. W. Diehl, “The theory of boundary critical phenomena,” Inter-national Journal of Modern Physics B 11, 3503–3523 (1997).

19

Bojan Zunkovic, Markus Heyl, Michael Knap, and Alessandro Silva, “Dynamical Quantum Phase Transitions in Spin Chains with Long-Range Interactions: Merging different concepts of non-equilibrium criticality,”arXiv:1609.08482 (2016).

20

Emil A. Yuzbashyan, Oleksandr Tsyplyatyev, and Boris L. Alt-shuler, “Relaxation and Persistent Oscillations of the Order Pa-rameter in Fermionic Condensates,”Phys. Rev. Lett. 96, 097005

(8)

(2006).

21

Martin Eckstein, Marcus Kollar, and Philipp Werner, “Interaction quench in the hubbard model: Relaxation of spectral function and the optical conductivity,”Phys. Rev. B 81, 115131 (2010).

22

Andrea Gambassi and Pasquale Calabrese, “Quantum quenches as classical critical films,”Europhys. Lett 95, 6 (2010).

23

Marco Schiro and Michele Fabrizio, “Time-Dependent Mean Field Theory for Quench Dynamics in Correlated Electron Sys-tems,”Phys. Rev. Lett. 105, 076401 (2010).

24

Bruno Sciolla and Giulio Biroli, “Dynamical transitions and quan-tum quenches in mean-field models,”J. Stat. Mech.: Theor. and Exper. 11, P11003 (2011).

25 Simone A. Hamerla and Gotz S. Uhrig, “Dynamical transition

in interaction quenches of the one-dimensional Hubbard model,”

Phys. Rev. B 87, 064304 (2013).

26 Pietro Smacchia, Michael Knap, Eugene Demler, and Alessandro

Silva, “Exploring dynamical phase transitions and prethermaliza-tion with quantum noise of excitaprethermaliza-tions,”Phys. Rev. B 91, 205136 (2015).

27

Markus Heyl, Anatoli Polkovnikov, and Stefan Kehrein, “Dy-namical Quantum Phase Transitions in the Transverse-Field Ising Model,”Phys. Rev. Lett. 113, 135704 (2013).

28

Markus Heyl, “Dynamical Quantum Phase Transitions in Systems with Broken-Symmetry Phases,”Phys. Rev. Lett. 113, 205701 (2014).

29 C. Karrasch and D. Schuricht, “Dynamical phase tran- sitions

af-ter quenches in non-integrable models,”Phys. Rev. B 87, 195104 (2013).

30 F. Andraschko and J. Sirker, “Dynamical quantum phase

transi-tions and the Loschmidt echo: A transfer matrix approach,”Phys. Rev. B 89, 125120 (2014).

31 J. Kriel, C. Karrasch, and S. Kehrein, “Dynamical quantum phase

transitions in the axial next-nearest-neighbor Ising chain,”Phys. Rev. B 90, 125106 (2014).

32 Elena Canovi, Philipp Werner, and Martin Eckstein, “First-order

dynamical phase transitions,”Phys. Rev. Lett 113, 265702 (2014).

33

Szabolcs Vajna and Bal´azs D´ora, “Disentangling dynamical phase transitions from equilibrium phase transitions,”Phys. Rev. B 89, 161105 (2014).

34

Bal´azs D´ora, Frank Pollmann, J´ozsef Fort´agh, and Gergely Zar´and, “Loschmidt Echo and the Many-body Orthogonality Catastrophe in a Qubit-Coupled Luttinger Liquid,” Phys. Rev. Lett. 111, 046402 (2013).

35

Markus Heyl, “Scaling and Universality at Dynamical Quantum Phase Transitions,”Phys. Rev. Lett. 115, 140602 (2015).

36

Shradda Sharma, Sei Suzuki, and Amit Dutta, “Quenches and

dynamical phase transitions in a nonintegrable quantum Ising model,”Phys. Rev. B 92, 104306 (2015).

37 Markus Heyl, “Quenching a quantum critical state by the order

parameter: Dynamical quantum phase transitions and quantum speed limits,”Phys. Rev. B 95, 060504(R) (2017).

38 Thom´as Fogarty, Ayaka Usui, Thomas Busch, Alessandro Silva,

and John Goold, “Dynamical phase transitions, temporal orthog-onality and the dynamics of observables in one dimensional ultra-cold quantum gases: from the continuum to the lattice,”

arXiv:1704.07659 (2017).

39

Marko Cetina, Michael Jag, Rianne S. Lous, Isabella Fritsche, Jook T. M. Walraven, Rudolf Grimm, Jesper Levinsen, Meera M. Parish, Richard Schmidt, Michael Knap, and Eugene Demler, “Ultrafast many-body interferometry of impurities coupled to a fermi sea,”Science 354, 96–99.

40 Moshe Moshe and Jean Zinn-Justin, “Quantum field theory in the

large N limit: a review,”Phys. Rep. 385, 385 (2003).

41

Anushya Chandran, Arun Nanduri, S. S. Gubser, and S. L. Sondhi, “Equilibration and coarsening in the quantum O(N ) model at infinite N ,”Phys. Rev. B 88, 024306 (2013).

42

Alessio Chiocchetta, Marco Tavora, Andrea Gambassi, and Aditi Mitra, “Short-time universal scaling in an isolated quantum sys-tem after a quench,”Phys. Rev. B 91, 220302(R) (2015).

43

Anna Maraga, Alessio Chiocchetta, Aditi Mitra, and Andrea Gambassi, “Aging and coarsening in isolated quantum systems after a quench: Exact results for the quantum O(N ) model with N → ∞,”Phys. Rev. E 92, 042151 (2015).

44

Gert Aarts, Daria Ahrensmeier, Rudolf Baier, J¨urgen Berges, and Julien Serreau, “Far-from-equilibrium dynamics with broken symmetries from the 1/N expansion of the 2PI effective action,”

Phys. Rev. D 66, 045008 (2002).

45 J¨urgen Berges, “Controlled nonperturbative dynamics of quantum

fields out of equilibrium,” Nucl. Phys. A 699, 847–886 (2002).

46

Simon A. Weidinger and Michael Knap, “Floquet prethermaliza-tion and regimes of heating in a periodically driven, interacting quantum system,”Sci. Rep. 7, 45382 (2017).

47

Peter Talkner, Eric Lutz, and Peter H¨anggi, “Fluctuation theo-rems: Work is not an observable,”Phys. Rev. E 75, 050102(R) (2007).

48

Alessandro Silva, “Statistics of Work Done on a Quantum Critical System by Quenching a Control Parameter,”Phys. Rev. Lett. 101, 120603 (2008).

49

Michele Campisi, Peter H¨anggi, and Peter Talkner, “Colloquium: Quantum fluctuation relations: Foundations and applications,”

Riferimenti

Documenti correlati

The first Chapter goes through Alfred Wallace’s biography, with particular emphasis to the paragraph named “The origins of the doubt”, referring to the debate between

Lo scopo di questo studio `e lo sviluppo di una nuova tecnica di imaging a riso- nanza magnetica che fornisca una mappa quantitativa della perfusione cerebrale, ovvero

Riassunto: L’obiettivo del presente lavoro di ricerca è stato quello di effettuare un’indagine riguardante la tecnica di decontaminazione elettrocinetica mirata ad

co dell'impianto filosofico. Come dirà nella Prefazione al primo volume degli Opuscoli filosofici nel '27, occorre saper unire virtù e verità, che rappresentano gli

Objectives: In the present study, we aimed to study interhemispheric connectivity in a sample of schizophrenia patients and healthy controls using functional

Hierarchical Cluster Analysis (HCA), was applied to several spectra of individual MEG-01 blasts that had been ob- tained by SR microFT-IR absorbance spectroscopy per- formed at

4 give the amplitude es- timates for each component measured from the data, while the coloured and grey histograms give the distribution of these quan- tities for the 143 ×

A Permutational Multivariate Analysis of Variance (PERMANOVA) was performed to explore the effects and significance of several variable such as sample Source (bark or grape),