Next Article in Journal
Maximum Entropy Exploration in Contextual Bandits with Neural Networks and Energy Based Models
Next Article in Special Issue
Coarse-Grained Effective Hamiltonian via the Magnus Expansion for a Three-Level System
Previous Article in Journal
Parameterization of the Stochastic Model for Evaluating Variable Small Data in the Shannon Entropy Basis
Previous Article in Special Issue
On the Fidelity Robustness of CHSH–Bell Inequality via Filtered Random States
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Superradiant Quantum Phase Transition for an Exactly Solvable Two-Qubit Spin-Boson Model

by
Roberto Grimaudo
1,*,†,
Davide Valenti
1,†,
Alessandro Sergi
2,3,† and
Antonino Messina
4,†
1
Dipartimento di Fisica e Chimica “Emilio Segrè”, Group of Interdisciplinary Theoretical Physics, Università degli Studi di Palermo, Viale delle Scienze Ed. 18, 90128 Palermo, Italy
2
Dipartimento di Scienze Matematiche e Informatiche, Scienze Fisiche e Scienze della Terra, Università degli Studi di Messina, Viale F. Stagno d’Alcontres 31, 98166 Messina, Italy
3
Institute of Systems Science, Durban University of Technology, P.O. Box 1334, Durban 4000, South Africa
4
Dipartimento di Matematica ed Informatica, Università degli Studi di Palermo, Via Archirafi 34, 90123 Palermo, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Entropy 2023, 25(2), 187; https://doi.org/10.3390/e25020187
Submission received: 5 December 2022 / Revised: 12 January 2023 / Accepted: 16 January 2023 / Published: 17 January 2023

Abstract

:
A spin-boson-like model with two interacting qubits is analysed. The model turns out to be exactly solvable since it is characterized by the exchange symmetry between the two spins. The explicit expressions of eigenstates and eigenenergies make it possible to analytically unveil the occurrence of first-order quantum phase transitions. The latter are physically relevant since they are characterized by abrupt changes in the two-spin subsystem concurrence, in the net spin magnetization and in the mean photon number.

1. Introduction

Every realistic quantum physical system is unavoidably coupled to sources of decoherence and/or dissipation [1]. Open quantum systems are particularly intriguing since their dynamic behaviour provides the platform to study the quantum-classical border.
A basic model exhibiting the quantum dissipation phenomenon is the spin-boson model (SBM), which describes a single spin-1/2 coupled to a quantized bosonic field [2]. The SBM has been deeply investigated through several methods and techniques since the 1980s [2,3,4,5,6,7,8,9,10].
Because of zero-point fluctuations rather than thermal ones [11,12,13], the SBM exhibits quantum phase transitions (QPTs) with respect to both the system-bath coupling and the transverse field strength [4,6,7,8,14,15,16,17,18,19,20]. Despite of its apparent simplicity, the SBM has spurred theoretical and experimental investigations which have successfully explored the basic physics of open quantum systems, significantly contributing to the understanding of different basic aspects that emerge experimentally in the behavior of such systems. Moreover, thanks to its versatility and generality, the SBM is at the basis of numerous applications in several fields, ranging from quantum information, quantum computation, and quantum simulation, to quantum optics and condensed matter physics [21,22,23,24,25].
In the last years a growing attention has been focusing on the decoherent and dissipative dynamics of the two-qubit SBM [17,26,27,28,29]. These models turn out to be useful to describe physical systems consisting of a collection of bi-nuclear units [30,31]. In the two-qubit SBMs until now analysed, either decoupled qubits or the simplest spin–spin coupling have been considered [17,27,32]. The model we are going to investigate in the following sections, instead, presents a non-trivial isotropic Heisenberg interaction between the two spins. The inclusion of such an interaction term is of crucial importance for those systems where the spin–spin interaction cannot be neglected. In particular it plays a crucial role in quantum computation [33,34] and, more generally, for those scenarios where the spin–spin coupling is the key ingredient for both performing controlled gates [35,36] and generating multipartite entangled states [37,38,39].
In this work we show that, thanks to the symmetries exhibited by the Hamiltonian, we are able to find the analytical expressions of both eigenstates and (related) eigenenergies. The dynamical problem can be exactly solved and this fact enables to bring to light the occurrence of first-order QPTs, characterized by abrupt changes of three physical quantities of experimental interest: the level of entanglement between the two spins (estimated through the concurrence [40]), the net two-spin magnetization, and the mean value of the number of bosonic field excitations. In particular, the change from a vanishing to a non-vanishing value of the mean photon number allows to speak of superradiant phase transition. The paper is structured as follows. In Section 2 the general model introduced in Ref. [41] and its symmetry properties are presented. In Section 3 the physical conditions which realize the exchange symmetry between the two spins and which make the model exactly solvable are considered. The analytical expressions of the eigenstates and eigenenergies are further derived, with the occurrence of first-order QPTs with respect to the parameters characterizing the Hamiltonian model. Final comments are reported in the last section.

2. Model and Symmetries

Let us consider the following open X Y Z Heisenberg quantum model (in units of )
H = Ω 1 σ ^ 1 z + Ω 2 σ ^ 2 z + j = 1 N ω j a ^ j a ^ j + γ x σ ^ 1 x σ ^ 2 x + γ y σ ^ 1 y σ ^ 2 y + γ z σ ^ 1 z σ ^ 2 z + k = 1 2 j = 1 N c i j a ^ j + a ^ j σ ^ i z ,
describing two interacting spin-1/2’s subjected to local longitudinal (along the z direction) fields and coupled to a common reservoir. A pictorial representation of the system is shown in Figure 1.
σ ^ k x , σ ^ k y and σ ^ k z ( k = 1 , 2 ) are the Pauli matrices, and a j and a j are the annihilation and creation boson operators of the j-th field mode. Ω i and ω j are the characteristic frequencies of the two spin-qubits ( i = 1 , 2 ) and the j-th field mode, respectively, and γ k ( k = x , y , z ) are the three real parameters characterizing the spin–spin anisotropic Heisenberg interaction.
The Hamiltonian is invariant when each spin is rotated by π around the z ^ -axis. The unitary operator accomplishing such a transformation is [41,42,43,44,45]
e i π σ ^ 1 z / 2 e i π σ ^ 2 z / 2 = σ ^ 1 z σ ^ 2 z = cos π 2 ^ z ,
where ^ z σ ^ 1 z + σ ^ 2 z . Thus σ ^ 1 z σ ^ 2 z is a constant of motion, which in turn implies the existence of two dynamically invariant subspaces ( H a and H b ) related to its two eigenvalues ( ± 1 ) . It is possible to see that the two effective Hamiltonians, governing the dynamics of the two-qubit/bath system in each subspace, read [41]
H a = Ω a σ ^ a z + γ a σ ^ a x + γ z 1 ^ a + j = 1 N ω j a ^ j a ^ j + j = 1 N c j a a ^ j + a ^ j σ ^ a z ,
for σ 1 z σ 2 z = 1 , and
H b = Ω b σ ^ b z + γ b σ ^ b x + γ z 1 ^ b + j = 1 N ω j a ^ j a ^ j + j = 1 N c j b a ^ j + a ^ j σ ^ b z ,
for σ 1 z σ 2 z = 1 , with
Ω a / b = Ω 1 ± Ω 2 , γ a / b = γ x ± γ y , c j a / b = c 1 j ± c 2 j .
The original Hamiltonian can be then written as H = H a H b . Thus, the dynamics of the two-spin/bath system in each subspace can be simulated by that of an effective single-impurity spin-boson model (SISBM), that is a fictitious two-level system immersed in a fictitious magnetic field and coupled with a reservoir through effective coupling constants. In this case, the role of the effective transverse field is played by the coupling existing between the two spin-1/2’s. It is important to underline that, although the two effective Hamiltonians are qualitatively similar, they can deeply differ under specific physical conditions, leading to a remarkably different dynamics of the physical system described by Equation (1) in the two different subspaces of the two-spin/bath system.
The subspace a related to the eigenvalue + 1 of σ ^ 1 z σ ^ 2 z is spanned by
{ | + + , | } j = 1 N { | n j } n = 0 ,
with σ ^ z | ± = ± | ± , and a j ^ a j ^ | n j = n j | n j ( j = 1 , , N ). It means that the two states { | + + , | } are mapped into the states { | + a , | a } of the fictitious two-level system a. So, in this case, by studying the dynamics of the fictitious spin-1/2 a effectively coupled to a bath, we study the dynamics of the two-spin/bath system within the subspace H a ruled by the effective Hamiltonian (3). Analogously, the subspace H b related to the eigenvalue 1 of σ ^ 1 z σ ^ 2 z is spanned by the two-spin/bath states
{ | + , | + } j = 1 N { | n j } n = 0 ,
and the effective Hamiltonian ruling the dynamics is that given in Equation (4). In this case, the two states { | + b , | b } of the fictitious spin-1/2 b are the mapping images of the two two-spin states { | + , | + } , respectively. Thanks to such a dynamic separation, the time evolution from initial conditions involving the two invariant subspaces can be easily achieved.
Finally, it is worth pointing out that, being each subdynamics ruled by an effective spin-boson Hamiltonian, all the results obtained for the spin-boson model can be applied to each subdynamics and interpreted in terms of the two interacting spin-1/2’s. Then, we can disclose the dynamics of the two-spin/bath system by separately solving the two effective spin-boson dynamical problems and ‘merging’ the obtained results.

3. Exactly Solvable Case

In this section we specialize the model (1) making it invariant under the two spin exchange. The simplest way to reach this goal is to introduce the following physically transparent links between the parameters appearing in H:
Ω 1 = Ω 2 = Ω / 2 , γ x = γ y = γ / 2 , c 1 j = c 2 j = c j / 2 ( j ) .
For this exchange-symmetry case the Hamiltonian of the two-spin/bath system reads
H = Ω 2 ( σ ^ 1 z + σ ^ 2 z ) + j = 1 N ω j a ^ j a ^ j + γ 2 ( σ ^ 1 x σ ^ 2 x + σ ^ 1 y σ ^ 2 y ) + γ z σ ^ 1 z σ ^ 2 z + j = 1 N c j 2 a ^ j + a ^ j ( σ ^ 1 z + σ ^ 2 z ) .
With respect to Hamiltonian (1), where only the square of ^ z is a constant of motion, Hamiltonian (9) exhibits a higher symmetry, being also the z-component of the total spin a constant of motion. The two effective Hamiltonians, which now rule the dynamics of the two-spin/bath system in the two invariant subspaces (still definable), become
H a = Ω σ ^ a z + γ z 1 ^ a + j = 1 N ω j a ^ j a ^ j + j = 1 N c j a ^ j + a ^ j σ ^ a z , H b = γ σ ^ b x γ z 1 ^ b + j = 1 N ω j a ^ j a ^ j .
We underline that, under the particular physical conditions in Equation (8), H b presents an effective decoupling of the fictitious two-level system b (which simulates the behaviour of the two actual spins within the subspace b) from the relative bosonic bath. This means that the subspace b is a decoherence-free subspace, or, in other words, that any initial state of the two-spin system belonging to such a subspace evolves as if the coupling between the two spins and the bath were absent [46,47,48]. The physical reason at the basis of this occurrence relies on the equal coupling of the two spins to the bath, i.e., c 1 j = c 2 j , j . The two couplings in a certain sense compensate each other, canceling out the effective spin-bath interaction in the subspace H b .
Moreover, we note that both Hamiltonians in Equation (10) result to be exactly diagonalizable. The homogeneity of the longitudinal magnetic field applied to the spin pair, namely Ω 1 = Ω 2 , causes σ ^ a z to be a constant of motion of H a . Thanks to the isotropy of the spin–spin coupling strength, i.e., γ x = γ y , σ ^ b x is instead a constant of motion of H b . The subspace H a ( H b ) can be then separated in two invariant subspaces, H a ± ( H b ± ), labeled by the two eigenvalues ( ± 1 ) of the corresponding constant of motion σ ^ a z ( σ ^ b x ). Therefore, the total Hilbert space H is the direct sum of these four infinite-dimensional, dynamically invariant Hilbert subspaces, namely
H = H a + H a H b + H b .
We underline that these four orthogonal invariant subspaces are identified and characterized by constant values of physical variables related only to the two-spin system. A consequence of Equation (11) is that, preparing the two-spin/bath system in a factorized spin-bath state belonging to one out of the four invariant subspaces given in Equation (11), the evolved state is still in a factorized form and the two-spin factor is stationary.
The two Hamiltonians
H a ± = ± Ω + γ z 1 ^ a + j = 1 N ω j a ^ j a ^ j ± j = 1 N c j a ^ j + a ^ j ,
corresponding to the two eigenvalues of σ ^ a z , must be intended as the effective Hamiltonians governing the dynamics of the two-spin/bath system within the subspace H a + and H a , respectively. The eigenstates (written in terms of the two-qubit/bath states, in view of the mapping discussed in Section 2) and eigenenergies of these two Hamiltonians can be written, respectively, as
| ψ { n j } a ± = | ± ± j = 1 N D ( α j ) ( a ^ j ) n j n j ! | 0 j , E { n j } a ± = ± Ω + γ z + j = 1 N ( n j α j 2 ) ω j
with D ( α j ) ( α j = c j / ω j ) and | 0 j being the displacement operator and the vacuum state of the j-th quantized bosonic mode, respectively [49]. Considering an infinite number of harmonic oscillators of the bath ( N ), the last term in the expression of the eigenenergies diverges. In order to obtain a finite esteem of this energy contribution, we consider an Ohmic spectral density, that is J ( ω ) = j c j 2 δ ( ω ω j ) α ω e ω / ω c ( α and ω c are the dimensionless parameter accounting for the spin-bath coupling strength and the cut-off energy of the bath, respectively). With this choice the energy contribution under scrutiny is finite and can be exactly derived, namely j α j 2 ω j α ω c (the sum is replaced by the integral, namely j J ( ω ) d ω [50,51]). In this case the lowest-energy states and the corresponding energies of the two Hamiltonians are uniquely defined and read
| ψ { 0 j } a ± | ψ 0 a ± = | ± ± j = 1 N D ( α j ) | 0 j , E { 0 j } a ± E 0 a ± = ± Ω + γ z α ω c .
We underline that the assumption of the Ohmic spectral density has been done only to obtain a finite esteem of the energy contribution of the bath (composed of an infinite number of quantum harmonic oscillators, N ). However, this choice does not affect the exact solvability of our model. Indeed, if we had considered a finite number of harmonic oscillators the expressions of the eigenenergies would have been the same, and the assumption of the Ohmic spectral density would not has been necessary, since the energy contribution of the bath would result finite.
The eigenstates (mapped into the two-qubit/bath language) and the corresponding eigenenergies of the two Hamiltonians H b ± related to the two eigenvalues of σ ^ b x read instead
| ψ { n j } b ± = | + ± | + 2 j = 1 N ( a ^ j ) n j n j ! | 0 j , E { n j } b ± = ± γ γ z + j n j ω j ,
with lowest-energy states and corresponding energies given by
| ψ { 0 j } b ± | ψ 0 b ± = | + ± | + 2 j = 1 N | 0 j , E { 0 j } b ± E 0 b ± = ± γ γ z .
Finally, it is worth noticing that, although the two effective Hamiltonians H a and H b simplify under the considered conditions (resulting to be analytically diagonalizable), the original two-spin/bath Hamiltonian does not acquire a trivial form, from which it is easy to understand the exact solvability of the model. Rather, it is the decomposition procedure and the effective description in terms of fictitious systems which allow to easily read and recognize simple structures, leading to the solution of the dynamical problem.

Quantum Phase Transitions

Thanks to the analytical solutions we can investigate and exactly derive the occurrence of quantum phase transitions. We consider the case of full isotropy in the spin–spin coupling, that is assuming γ z = γ x = γ y = γ / 2 in the Hamiltonian (9). In this instance the four lowest energies within each invariant subspace read
E 0 a + = Ω + γ / 2 α ω c , E 0 a = Ω + γ / 2 α ω c , E 0 b + = γ / 2 , E 0 b = 3 γ / 2 .
In Figure 2a–c the dependence of the above four eigenenergies is plotted as a function of the spin–spin coupling γ , the spin-bath coupling α , and the externally applied magnetic field Ω , respectively. It is possible to see that several level crossings occur. In particular, the one appearing between the two eigenenergies E 0 a (dotted green line) and E 0 b (solid black line) is of remarkable importance since it involves the ground state of the two-spin/bath system, giving rise to a first-order QPT.
Specifically, in Figure 2a we see that E 0 a < E 0 b for γ < γ c , while E 0 a > E 0 b for γ > γ c ( γ c being the crossing point, that is, the critical value of γ for which the ground state is degenarate and then not uniquely defined). Therefore, the ground state corresponds to | ψ 0 a for γ < γ c and to | ψ 0 b for γ > γ c . It means that, in this case, the ground state ‘moves’ from H a to H b by increasing the spin–spin coupling γ and crossing the critical point corresponding to the critical value γ c . This fact suggests that the ground state can belong to different invariant subspaces with respect to the parameter-space region we consider.
On the other side, by increasing the spin-bath coupling α , the opposite change occurs: E 0 a > E 0 b and E 0 a < E 0 b before and after, respectively, the crossing point α c (see Figure 2b). The same transition is observed by varying the parameter Ω around its critical values Ω c (see Figure 2c). The ground state is thus placed in H b (being | ψ 0 b ) for α < α c and Ω < Ω c . Then, it ‘moves’ to H a (becoming | ψ 0 a ), after crossing the critical values of the two parameters, that is for α > α c and Ω > Ω c (see Figure 2b and Figure 2c, respectively).
These crossings are of course accompanied by relevant physical changes in the ground state properties of the system. The two states | ψ 0 a and | ψ 0 b are indeed characterized by different values of three relevant physical observables, namely: (i) the level of entanglement between the two spins estimated through the concurrence C 0 x [40] which, in case of a general two-qubit state
| ψ = c + + | + + + c + | + + c + | + + c | ,
with | c + + | 2 + | c + | 2 + | c + | 2 + | c | 2 = 1 , reads
C = 2 | c + + c c + c + | ;
(ii) the net two-spin magnetization M 0 x ψ 0 x | ^ z | ψ 0 x ; (iii) the mean number of photons (or excitations) N 0 x ψ 0 x | a ^ a ^ | ψ 0 x of the boson field ( x = a , b ).
By explicitly calculating on | ψ 0 a the average values of the three physical observable listed above, we obtain:
{ C 0 a = 0 , M 0 a = 1 , N 0 a = j = 1 N α j 2 } .
The same calculation performed for the state | ψ 0 b gives
{ C 0 b = 1 , M 0 b = 0 , N 0 b = 0 } .
We see that the three physical quantities considered show different values for the two states which alternately result to be the ground state of the two-spin/bath system, depending on the parameter-space region taken into account. These observables then undergo an abrupt change in correspondence of the level-crossings occurring at the critical values of the control parameters. Therefore, we can speak of two different phases and then of first-order QPTs (at the crossing points) characterized by: (i) γ , α , and Ω as control parameters; (ii) the three physical quantities C 0 , M 0 , and N 0 as order parameters; (iii) the two states | ψ 0 a and | ψ 0 b as ground states of the two-spin/bath system in the two different phases.
We point out that the thermodynamic limit is not necessary, in general, for first-order QPTs [11]. In Refs. [52,53,54,55,56] the authors, studying the quantum Rabi model, make clear the difference between first-order QPTs occurring in the thermodynamic (classical) limit (vanishing frequencies for the bosonic field) and the first-order QPTs occurring in the fully quantum regime (that is, at finite frequencies). Precisely, they bring to light how the phase diagram (and then the physical feature) of the ground state of the system is substantially and remarkably different in the two regimes [52,53,54,55,56]. This circumstance clearly shows how few-body first-order QPTs offer a rich field of study to understand both semiclassical and fully quantum physical characteristics of microscopic light-matter systems.
In our case, however, the system is constituted by two interacting qubits and a multi-mode bosonic field. We underline that the bath is not a reservoir playing a passive role in the dynamics of the system. Its state is indeed not fixed and constantly equal to a thermal state, as in the case of a reservoir. We have indeed different states for the bath in the two ground states of the system related to the two different phases. Further, since the bath consists of an infinite number of quantum harmonic oscillators, the size of the system is then not finite. To esteem the energy contribution of the bath we have indeed taken the limit N , after assuming an Ohmic spectral density. In this case thus the QPTs brought to light can be intended occurring in the thermodynamic limit.
It is particularly worth pointing out the different level of entanglement between the spins, which characterizes the two phases: the two spins are in a disentangled state in the space H a , while exhibit a maximally entangled state in the space H b . Furthermore, the (abrupt) change from a vanishing to a non-vanishing value (or vice versa) of the mean photon number suggests the occurrence of a superradiant QPT. Precisely, the phase corresponding to the ground state | ψ 0 b can be intended as the normal phase, while that corresponding to the ground state | ψ 0 a can be thought of as the superradiant phase [57]. The peculiarity of such a superradiant QPT lies in the fact that it is a first-order QPT [58,59], differently from both the standard superradiant phase transition of the Dicke model [60,61,62] and the superradiant QPT of the quantum Rabi model [57], which are instead phase transitions of the second order.

4. Conclusions

The present work studies a model which describes an ubiquitous physical situation that can be formalized in terms of two interacting qubits coupled to a common bosonic field. We have shown that, despite its non-triviality, the model turns out to be exactly solvable when a full symmetry between the two spins is considered. Precisely, we have focused on an isotropic Heisenberg spin–spin interaction, an homogeneous magnetic field applied to the spin pair, and an equal coupling of the spin–boson type between each spin and each mode of the field. In this instance it is then possible to analytically derive eigenenergies and eigenstates of the system. Thanks to our exact approach it has been possible to unveil the occurrence of QPTs with respect to the three parameters characterizing the model (the spin–spin coupling, the spin-bath coupling, and the strength of the classical magnetic field). These QPTs are characterized by abrupt changes in relevant physical observables of the system: the two-spin concurrence, the net spin magnetization, and the mean photon number. Such sudden changes are due to the fact that these QPTs are of the first order (presence of a level crossing). We have in particular emphasized that the change of the mean photon number from a vanishing to a non-vanishing value between the two phases allows to speak of superradiant phase transition. The peculiarity of such a superradiant phenomenon is related to the fact that it corresponds to a first-order QPT, differently from what happens for both the standard superradiant phase transition of the Dicke model [60,61,62] and the superradiant QPT of the quantum Rabi model [57], which are instead second-order phase transitions.
We underline that the analytical approach [63,64,65,66,67] used here allowed to exactly solve the time-independent Shrödinger equation. The method is further valid in every region of the parameter space: no constraints about the strength of the Hamiltonian parameters have been indeed introduced. The only requirement is the totally symmetric role of the two spins.
Since not so many exactly solvable (non-trivial) models, treating the coupling with a bath, are present in literature, our study can be then at the basis of an interesting insight on physics of open quantum systems. Moreover, thanks to its generality, our model can be applied to a plethora of physical systems useful for future quantum technologies. Until now, the two-spin-boson models considered in literature [17,27,32] have taken into account either decoupled qubits or the simplest spin–spin coupling. In our case, instead, the physical effects stemming from the presence of a (non-trivial) isotropic dipolar spin–spin coupling has been investigated for the first time. This aspect is of crucial importance since in some contexts the spin–spin interaction cannot be neglected, and in other scenarios, such as in quantum computation, it is fundamental to perform two-qubit quantum logic gates [35,36] and to generate entangled states of the system [37,38,39].
Finally, our exactly solvable model can be also exploited to test the accuracy of the standard techniques employed for mathematically treating open quantum systems, such as: (i) the standard Lindblad theory [68,69]; (ii) the partial Wigner transform [70,71,72]; (iii) the non-Hermitian formalism [73,74,75,76,77,78,79,80,81]; (iv) the stochastic approach [82,83,84,85,86,87,88].

Author Contributions

Conceptualization, R.G. and A.M.; Methodology, R.G. and A.M.; Validation, R.G., D.V. and A.M.; Formal analysis, R.G.; Investigation, R.G., D.V., A.S. and A.M.; Writing—original draft, R.G.; Writing—review editing, D.V., A.S. and A.M.; Supervision, D.V. and A.M.; Project administration, D.V.; Funding acquisition, D.V. All authors have read and agreed to the published version of the manuscript.

Funding

R.G. and D.V. acknowledge the financial support from PRIN Project PRJ-0232—Impact of Climate Change on the Biogeochemistry of Contaminants in the Mediterranean sea (ICCC).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Santi Prestipino for useful discussions about the subject of the work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Breuer, H.P.; Petruccione, F. The Theory of Open Quantum Systems; Oxford University Press on Demand: Oxford, UK, 2002. [Google Scholar]
  2. Leggett, A.J.; Chakravarty, S.; Dorsey, A.T.; Fisher, M.P.A.; Garg, A.; Zwerger, W. Dynamics of the dissipative two-state system. Rev. Mod. Phys. 1987, 59, 1–85. [Google Scholar] [CrossRef]
  3. Le Hur, K.; Doucet-Beaupré, P.; Hofstetter, W. Entanglement and Criticality in Quantum Impurity Systems. Phys. Rev. Lett. 2007, 99, 126801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Vojta, M.; Tong, N.H.; Bulla, R. Quantum Phase Transitions in the Sub-Ohmic Spin-Boson Model: Failure of the Quantum-Classical Mapping. Phys. Rev. Lett. 2005, 94, 070604. [Google Scholar] [CrossRef] [Green Version]
  5. Bulla, R.; Tong, N.H.; Vojta, M. Numerical Renormalization Group for Bosonic Systems and Application to the Sub-Ohmic Spin-Boson Model. Phys. Rev. Lett. 2003, 91, 170601. [Google Scholar] [CrossRef] [Green Version]
  6. Hur, K.L. Entanglement entropy, decoherence, and quantum phase transitions of a dissipative two-level system. Ann. Phys. 2008, 323, 2208–2240. [Google Scholar] [CrossRef] [Green Version]
  7. Nazir, A.; McCutcheon, D.P.S.; Chin, A.W. Ground state and dynamics of the biased dissipative two-state system: Beyond variational polaron theory. Phys. Rev. B 2012, 85, 224301. [Google Scholar] [CrossRef] [Green Version]
  8. Winter, A.; Rieger, H.; Vojta, M.; Bulla, R. Quantum Phase Transition in the Sub-Ohmic Spin-Boson Model: Quantum Monte Carlo Study with a Continuous Imaginary Time Cluster Algorithm. Phys. Rev. Lett. 2009, 102, 030601. [Google Scholar] [CrossRef] [Green Version]
  9. Deng, T.; Yan, Y.; Chen, L.; Zhao, Y. Dynamics of the two-spin spin-boson model with a common bath. J. Chem. Phys. 2016, 144, 144102. [Google Scholar] [CrossRef]
  10. Wang, L.; Fujihashi, Y.; Chen, L.; Zhao, Y. Finite-temperature time-dependent variation with multiple Davydov states. J. Chem. Phys. 2017, 146, 124127. [Google Scholar] [CrossRef] [Green Version]
  11. Vojta, M. Quantum phase transitions. Rep. Prog. Phys. 2003, 66, 2069–2110. [Google Scholar] [CrossRef]
  12. Rossini, D.; Vicari, E. Coherent and dissipative dynamics at quantum phase transitions. Phys. Rep. 2021, 936, 1–110. [Google Scholar] [CrossRef]
  13. Carollo, A.; Valenti, D.; Spagnolo, B. Geometry of quantum phase transitions. Phys. Rep. 2020, 838, 1–72. [Google Scholar] [CrossRef] [Green Version]
  14. Tong, Q.J.; An, J.H.; Luo, H.G.; Oh, C.H. Quantum phase transition in the delocalized regime of the spin-boson model. Phys. Rev. B 2011, 84, 174301. [Google Scholar] [CrossRef] [Green Version]
  15. De Filippis, G.; de Candia, A.; Cangemi, L.M.; Sassetti, M.; Fazio, R.; Cataudella, V. Quantum phase transitions in the spin-boson model: Monte Carlo method versus variational approach à la Feynman. Phys. Rev. B 2020, 101, 180408. [Google Scholar] [CrossRef]
  16. Zhang, Y.Y.; Chen, Q.H.; Wang, K.L. Quantum phase transition in the sub-Ohmic spin-boson model: An extended coherent-state approach. Phys. Rev. B 2010, 81, 121105. [Google Scholar] [CrossRef] [Green Version]
  17. Dolgitzer, D.; Zeng, D.; Chen, Y. Dynamical quantum phase transitions in the spin-boson model. Opt. Express 2021, 29, 23988–23996. [Google Scholar] [CrossRef]
  18. Wang, Y.Z.; He, S.; Duan, L.; Chen, Q.H. Rich phase diagram of quantum phases in the anisotropic subohmic spin-boson model. Phys. Rev. B 2020, 101, 155147. [Google Scholar] [CrossRef]
  19. Winter, A.; Rieger, H. Quantum phase transition and correlations in the multi-spin-boson model. Phys. Rev. B 2014, 90, 224401. [Google Scholar] [CrossRef] [Green Version]
  20. Kaur, K.; Sépulcre, T.; Roch, N.; Snyman, I.; Florens, S.; Bera, S. Spin-Boson Quantum Phase Transition in Multilevel Superconducting Qubits. Phys. Rev. Lett. 2021, 127, 237702. [Google Scholar] [CrossRef]
  21. Dunnett, A.J.; Chin, A.W. Matrix Product State Simulations of Non-Equilibrium Steady States and Transient Heat Flows in the Two-Bath Spin-Boson Model at Finite Temperatures. Entropy 2021, 23, 77. [Google Scholar] [CrossRef]
  22. Puebla, R.; Casanova, J.; Houhou, O.; Solano, E.; Paternostro, M. Quantum simulation of multiphoton and nonlinear dissipative spin-boson models. Phys. Rev. A 2019, 99, 032303. [Google Scholar] [CrossRef] [Green Version]
  23. Magazzù, L.; Denisov, S.; Hänggi, P. Asymptotic Floquet states of a periodically driven spin-boson system in the nonperturbative coupling regime. Phys. Rev. E 2018, 98, 022111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Magazzù, L.; Forn-Díaz, P.; Belyansky, R.; Orgiazzi, J.L.; Yurtalan, M.; Otto, M.R.; Lupascu, A.; Wilson, C.; Grifoni, M. Probing the strongly driven spin-boson model in a superconducting quantum circuit. Nat. Commun. 2018, 9, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Casanova, J.; Puebla, R.; Moya-Cessa, H.; Plenio, M.B. Connecting nth order generalised quantum Rabi models: Emergence of nonlinear spin-boson coupling via spin rotations. npj Quantum Inf. 2018, 4, 47. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, Y.Z.; He, S.; Duan, L.; Chen, Q.H. Quantum tricritical point emerging in the spin-boson model with two dissipative spins in staggered biases. Phys. Rev. B 2021, 103, 205106. [Google Scholar] [CrossRef]
  27. Zhou, N.; Zhang, Y.; Lü, Z.; Zhao, Y. Variational Study of the Two-Impurity Spin–Boson Model with a Common Ohmic Bath: Ground-State Phase Transitions. Annalen der Physik 2018, 530, 1800120. [Google Scholar] [CrossRef] [Green Version]
  28. Bonart, J. Dissipative phase transition in a pair of coupled noisy two-level systems. Phys. Rev. B 2013, 88, 125139. [Google Scholar] [CrossRef] [Green Version]
  29. McCutcheon, D.P.S.; Nazir, A.; Bose, S.; Fisher, A.J. Separation-dependent localization in a two-impurity spin-boson model. Phys. Rev. B 2010, 81, 235321. [Google Scholar] [CrossRef] [Green Version]
  30. Calvo, R.; Abud, J.E.; Sartoris, R.P.; Santana, R.C. Collapse of the EPR fine structure of a one-dimensional array of weakly interacting binuclear units: A dimensional quantum phase transition. Phys. Rev. B 2011, 84, 104433. [Google Scholar] [CrossRef]
  31. Napolitano, L.M.B.; Nascimento, O.R.; Cabaleiro, S.; Castro, J.; Calvo, R. Isotropic and anisotropic spin-spin interactions and a quantum phase transition in a dinuclear Cu(II) compound. Phys. Rev. B 2008, 77, 214423. [Google Scholar] [CrossRef]
  32. Orth, P.P.; Roosen, D.; Hofstetter, W.; Le Hur, K. Dynamics, synchronization, and quantum phase transitions of two dissipative spins. Phys. Rev. B 2010, 82, 144423. [Google Scholar] [CrossRef] [Green Version]
  33. Barenco, A.; Bennett, C.H.; Cleve, R.; DiVincenzo, D.P.; Margolus, N.; Shor, P.; Sleator, T.; Smolin, J.A.; Weinfurter, H. Elementary gates for quantum computation. Phys. Rev. A 1995, 52, 3457–3467. [Google Scholar] [CrossRef] [Green Version]
  34. Hua, M.; Tao, M.J.; Deng, F.G. Universal quantum gates on microwave photons assisted by circuit quantum electrodynamics. Phys. Rev. A 2014, 90, 012328. [Google Scholar] [CrossRef] [Green Version]
  35. Romero, G.; Ballester, D.; Wang, Y.M.; Scarani, V.; Solano, E. Ultrafast Quantum Gates in Circuit QED. Phys. Rev. Lett. 2012, 108, 120501. [Google Scholar] [CrossRef]
  36. Barends, R.; Quintana, C.M.; Petukhov, A.G.; Chen, Y.; Kafri, D.; Kechedzhi, K.; Collins, R.; Naaman, O.; Boixo, S.; Arute, F.; et al. Diabatic Gates Freq.-Tunable Supercond. Qubits. Phys. Rev. Lett. 2019, 123, 210501. [Google Scholar] [CrossRef] [Green Version]
  37. Kang, Y.H.; Chen, Y.H.; Wu, Q.C.; Huang, B.H.; Song, J.; Xia, Y. Fast generation of W states of superconducting qubits with multiple Schrödinger dynamics. Sci. Rep. 2016, 6, 1–13. [Google Scholar] [CrossRef] [Green Version]
  38. Lu, M.; Xia, Y.; Song, J.; An, N.B. Generation of N-atom W-class states in spatially separated cavities. J. Opt. Soc. Am. B 2013, 30, 2142–2147. [Google Scholar] [CrossRef] [Green Version]
  39. Li, J.; Paraoanu, G.S. Generation and propagation of entanglement in driven coupled-qubit systems. New J. Phys. 2009, 11, 113020. [Google Scholar] [CrossRef]
  40. Wootters, W.K. Entanglement of Formation of an Arbitrary State of Two Qubits. Phys. Rev. Lett. 1998, 80, 2245–2248. [Google Scholar] [CrossRef] [Green Version]
  41. Grimaudo, R.; Messina, A.; Nakazato, H.; Sergi, A.; Valenti, D. Spin-spin coupling-based quantum and classical phase transitions in two-impurity spin-boson models. arXiv 2022, arXiv:2205.09367. [Google Scholar]
  42. Grimaudo, R.; Messina, A.; Nakazato, H. Exactly solvable time-dependent models of two interacting two-level systems. Phys. Rev. A 2016, 94, 022108. [Google Scholar] [CrossRef]
  43. Grimaudo, R.; Vitanov, N.V.; Messina, A. Coupling-assisted Landau-Majorana-Stückelberg-Zener transition in a system of two interacting spin qubits. Phys. Rev. B 2019, 99, 174416. [Google Scholar] [CrossRef] [Green Version]
  44. Grimaudo, R.; Vitanov, N.V.; Messina, A. Landau-Majorana-Stückelberg-Zener dynamics driven by coupling for two interacting qutrit systems. Phys. Rev. B 2019, 99, 214406. [Google Scholar] [CrossRef] [Green Version]
  45. Grimaudo, R.; Nakazato, H.; Messina, A.; Vitanov, N.V. Dzyaloshinskii-Moriya and dipole-dipole interactions affect coupling-based Landau-Majorana-Stückelberg-Zener transitions. Phys. Rev. Res. 2020, 2, 033092. [Google Scholar] [CrossRef]
  46. Napoli, A.; Guccione, M.; Messina, A.; Chruściński, D. Interaction-free evolving states of a bipartite system. Phys. Rev. A 2014, 89, 062104. [Google Scholar] [CrossRef] [Green Version]
  47. Chruściński, D.; Messina, A.; Militello, B.; Napoli, A. Interaction-free evolution in the presence of time-dependent Hamiltonians. Phys. Rev. A 2015, 91, 042123. [Google Scholar] [CrossRef] [Green Version]
  48. Militello, B.; Chruściński, D.; Messina, A.; Należyty, P.; Napoli, A. Generalized interaction-free evolutions. Phys. Rev. A 2016, 93, 022113. [Google Scholar] [CrossRef] [Green Version]
  49. Cahill, K.E.; Glauber, R.J. Ordered Expansions in Boson Amplitude Operators. Phys. Rev. 1969, 177, 1857–1881. [Google Scholar] [CrossRef] [Green Version]
  50. Weiss, U. Quantum Dissipative Systems; World Scientific: Singapore, 2012. [Google Scholar] [CrossRef]
  51. Albash, T.; Boixo, S.; Lidar, D.A.; Zanardi, P. Quantum adiabatic Markovian master equations. New J. Phys. 2012, 14, 123016. [Google Scholar] [CrossRef]
  52. Ying, Z.J.; Cong, L.; Sun, X.M. Quantum phase transition and spontaneous symmetry breaking in a nonlinear quantum Rabi model. J. Phys. A: Math. Theor. 2020, 53, 345301. [Google Scholar] [CrossRef]
  53. Ying, Z.J. Symmetry-breaking patterns, tricriticalities, and quadruple points in the quantum Rabi model with bias and nonlinear interaction. Phys. Rev. A 2021, 103, 063701. [Google Scholar] [CrossRef]
  54. Liu, J.; Liu, M.; Ying, Z.J.; Luo, H.G. Fundamental Models in the Light–Matter Interaction: Quantum Phase Transitions and the Polaron Picture. Adv. Quantum Technol. 2021, 4, 2000139. [Google Scholar] [CrossRef]
  55. Ying, Z.J. From Quantum Rabi Model To Jaynes–Cummings Model: Symmetry-Breaking Quantum Phase Transitions, Symmetry-Protected Topological Transitions and Multicriticality. Adv. Quantum Technol. 2022, 5, 2100088. [Google Scholar] [CrossRef]
  56. Ying, Z.J. Hidden Single-Qubit Topological Phase Transition without Gap Closing in Anisotropic Light-Matter Interactions. Adv. Quantum Technol. 2022, 5, 2100165. [Google Scholar] [CrossRef]
  57. Hwang, M.J.; Puebla, R.; Plenio, M.B. Quantum Phase Transition and Universal Dynamics in the Rabi Model. Phys. Rev. Lett. 2015, 115, 180404. [Google Scholar] [CrossRef] [Green Version]
  58. Lee, C.F.; Johnson, N.F. First-Order Superradiant Phase Transitions in a Multiqubit Cavity System. Phys. Rev. Lett. 2004, 93, 083001. [Google Scholar] [CrossRef] [Green Version]
  59. Li, X.; Dreon, D.; Zupancic, P.; Baumgärtner, A.; Morales, A.; Zheng, W.; Cooper, N.R.; Donner, T.; Esslinger, T. First order phase transition between two centro-symmetric superradiant crystals. Phys. Rev. Res. 2021, 3, L012024. [Google Scholar] [CrossRef]
  60. Hepp, K.; Lieb, E.H. On the superradiant phase transition for molecules in a quantized radiation field: The dicke maser model. Ann. Phys. 1973, 76, 360–404. [Google Scholar] [CrossRef]
  61. Wang, Y.K.; Hioe, F.T. Phase Transition in the Dicke Model of Superradiance. Phys. Rev. A 1973, 7, 831–836. [Google Scholar] [CrossRef]
  62. Leonardi, C.; Persico, F.; Vetri, G. Dicke model and the theory of driven and spontaneous emission. Riv. Nuovo Cimento (1978–1999) 1986, 9, 1–85. [Google Scholar] [CrossRef]
  63. Grimaudo, R.; Messina, A.; Ivanov, P.A.; Vitanov, N.V. Spin-1/2 sub-dynamics nested in the quantum dynamics of two coupled qutrits. J. Phys. A Math. Theor. 2017, 50, 175301. [Google Scholar] [CrossRef] [Green Version]
  64. Grimaudo, R.; Lamata, L.; Solano, E.; Messina, A. Cooling of many-body systems via selective interactions. Phys. Rev. A 2018, 98, 042330. [Google Scholar] [CrossRef] [Green Version]
  65. Grimaudo, R.; Isar, A.; Mihaescu, T.; Ghiu, I.; Messina, A. Dynamics of quantum discord of two coupled spin-1/2’s subjected to time-dependent magnetic fields. Results Phys. 2019, 13, 102147. [Google Scholar] [CrossRef]
  66. Grimaudo, R.; Belousov, Y.; Nakazato, H.; Messina, A. Time evolution of a pair of distinguishable interacting spins subjected to controllable and noisy magnetic fields. Ann. Phys. 2018, 392, 242–259. [Google Scholar] [CrossRef] [Green Version]
  67. Grimaudo, R.; Man’ko, V.I.; Man’ko, M.A.; Messina, A. Dynamics of a harmonic oscillator coupled with a Glauber amplifier. Phys. Scr. 2019, 95, 024004. [Google Scholar] [CrossRef] [Green Version]
  68. Gorini, V.; Kossakowski, A.; Sudarshan, E.C.G. Completely positive dynamical semigroups of N-level systems. J. Math. Phys. 1976, 17, 821–825. [Google Scholar] [CrossRef]
  69. Lindblad, G. On the generators of quantum dynamical semigroups. Comm. Math. Phys. 1976, 48, 119–130. [Google Scholar] [CrossRef]
  70. Kapral, R.; Ciccotti, G. Mixed quantum-classical dynamics. J. Chem. Phys. 1999, 110, 8919–8929. [Google Scholar] [CrossRef]
  71. Kapral, R. Quantum-classical dynamics in a classical bath. J. Phys. Chem. A 2001, 105, 2885–2889. [Google Scholar] [CrossRef]
  72. Sergi, A. Non-Hamiltonian commutators in quantum mechanics. Phys. Rev. E 2005, 72, 066125. [Google Scholar] [CrossRef]
  73. Feshbach, H. Unified theory of nuclear reactions. Ann. Phys. 1958, 5, 357–390. [Google Scholar] [CrossRef]
  74. Bender, C.M.; Boettcher, S. Real Spectra in Non-Hermitian Hamiltonians Having PT Symmetry. Phys. Rev. Lett. 1998, 80, 5243–5246. [Google Scholar] [CrossRef] [Green Version]
  75. Mostafazadeh, A. Conceptual aspects of PT-symmetry and pseudo-Hermiticity: A status report. Phys. Scr. 2010, 82, 038110. [Google Scholar] [CrossRef]
  76. Rotter, I.; Bird, J.P. A review of progress in the physics of open quantum systems: Theory and experiment. Rep. Prog. Phys. 2015, 78, 114001. [Google Scholar] [CrossRef] [Green Version]
  77. Sergi, A.; Zloshchastiev, K.G. Time correlation functions for non-Hermitian quantum systems. Phys. Rev. A 2015, 91, 062108. [Google Scholar] [CrossRef] [Green Version]
  78. Brody, D.C.; Graefe, E.M. Mixed-State Evolution in the Presence of Gain and Loss. Phys. Rev. Lett. 2012, 109, 230405. [Google Scholar] [CrossRef] [Green Version]
  79. Grimaudo, R.; de Castro, A.S.M.; Kuś, M.; Messina, A. Exactly solvable time-dependent pseudo-Hermitian su(1,1) Hamiltonian models. Phys. Rev. A 2018, 98, 033835. [Google Scholar] [CrossRef] [Green Version]
  80. Grimaudo, R.; de Castro, A.S.M.; Nakazato, H.; Messina, A. Analytically solvable 2 × 2 PT-symmetry dynamics from su(1,1)-symmetry problems. Phys. Rev. A 2019, 99, 052103. [Google Scholar] [CrossRef] [Green Version]
  81. Grimaudo, R.; Messina, A.; Sergi, A.; Vitanov, N.V.; Filippov, S.N. Two-Qubit Entanglement Generation through Non-Hermitian Hamiltonians Induced by Repeated Measurements on an Ancilla. Entropy 2020, 22, 1184. [Google Scholar] [CrossRef]
  82. Benguria, R.; Kac, M. Quantum Langevin Equation. Phys. Rev. Lett. 1981, 46, 1–4. [Google Scholar] [CrossRef]
  83. Cortés, E.; West, B.J.; Lindenberg, K. On the generalized Langevin equation: Classical and quantum mechanical. J. Chem. Phys. 1985, 82, 2708–2717. [Google Scholar] [CrossRef]
  84. Ford, G.; Kac, M. On the quantum Langevin equation. J. Stat. Phys. 1987, 46, 803–810. [Google Scholar] [CrossRef]
  85. Ford, G.W.; Lewis, J.T.; O’Connell, R.F. Quantum Langevin equation. Phys. Rev. A 1988, 37, 4419–4428. [Google Scholar] [CrossRef] [Green Version]
  86. Gardiner, C. Quantum noise and quantum Langevin equations. IBM J. Res. Dev. 1988, 32, 127–136. [Google Scholar] [CrossRef]
  87. Reitz, M.; Sommer, C.; Genes, C. Langevin Approach to Quantum Optics with Molecules. Phys. Rev. Lett. 2019, 122, 203602. [Google Scholar] [CrossRef] [Green Version]
  88. Surazhevsky, I.; Demin, V.; Ilyasov, A.; Emelyanov, A.; Nikiruy, K.; Rylkov, V.; Shchanikov, S.; Bordanov, I.; Gerasimova, S.; Guseinov, D.; et al. Noise-assisted persistence and recovery of memory state in a memristive spiking neuromorphic network. Chaos Solitons Fractals 2021, 146, 110890. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the two-qubit spin-boson model. The two central arrows represent the two interacting qubits. The quantum harmonic oscillators constituting the bath are represented by the blue circles.
Figure 1. Schematic representation of the two-qubit spin-boson model. The two central arrows represent the two interacting qubits. The quantum harmonic oscillators constituting the bath are represented by the blue circles.
Entropy 25 00187 g001
Figure 2. Dependence of the eigenenergies E 0 a + (dashed red line), E 0 a (dotted green line), E 0 b + (dot-dashed blue line), and E 0 b (solid black line) on the dimensionless (a) spin–spin coupling γ / ω c , (b) spin–bath coupling α , and (c) the (classical) magnetic field strength Ω / ω c .
Figure 2. Dependence of the eigenenergies E 0 a + (dashed red line), E 0 a (dotted green line), E 0 b + (dot-dashed blue line), and E 0 b (solid black line) on the dimensionless (a) spin–spin coupling γ / ω c , (b) spin–bath coupling α , and (c) the (classical) magnetic field strength Ω / ω c .
Entropy 25 00187 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Grimaudo, R.; Valenti, D.; Sergi, A.; Messina, A. Superradiant Quantum Phase Transition for an Exactly Solvable Two-Qubit Spin-Boson Model. Entropy 2023, 25, 187. https://doi.org/10.3390/e25020187

AMA Style

Grimaudo R, Valenti D, Sergi A, Messina A. Superradiant Quantum Phase Transition for an Exactly Solvable Two-Qubit Spin-Boson Model. Entropy. 2023; 25(2):187. https://doi.org/10.3390/e25020187

Chicago/Turabian Style

Grimaudo, Roberto, Davide Valenti, Alessandro Sergi, and Antonino Messina. 2023. "Superradiant Quantum Phase Transition for an Exactly Solvable Two-Qubit Spin-Boson Model" Entropy 25, no. 2: 187. https://doi.org/10.3390/e25020187

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop