Projection based adiabatic elimination of bipartite open quantum systems
Abstract
Adiabatic elimination methods allow the reduction of the space dimension needed to describe systems dynamics which exhibits separation of time scale. For open quantum system, it consists in eliminating the fast part assuming it has almost instantaneously reached its steady-state and obtaining an approximation of the evolution of the slow part. These methods can be applied to eliminate a linear subspace within the system Hilbert space, or alternatively to eliminate a fast subsystems in a bipartite quantum system. In this work, we extend an adiabatic elimination method used for removing fast degrees of freedom within a open quantum system (Phys. Rev. A 2020, 101, 042102) to eliminate a subsystem from an open bipartite quantum system. As an illustration, we apply our technique to a dispersively coupled two-qubit system and in the case of the open Rabi model.
I Introduction
Adiabatic elimination is a method whereby the fast degrees of freedom of a system are removed while retaining an effective description of the slow degrees of freedom. This simplification can be very useful to obtain tractable and intuitive equations when only a coarse-grained or long times description is desired Haken (1975, 1977); Lax (1967); Cohen-Tannoudji (1992); Paulisch et al. (2014); Brion et al. (2007); You et al. (2003); Nagy et al. (2010); Douglas et al. (2015), depending on if the target system has a conservative Paulisch et al. (2014); Sinatra et al. (1995); Brion et al. (2007) or a dissipative Azouit et al. (2016, 2017a); Azouit (2017); Azouit et al. (2017b); Sarlette et al. (2020); Mirrahimi and Rouchon (2009); Reiter and Sørensen (2012); Kessler (2012) evolution. There are two classes of manifolds on which adiabatic elimination has been applied, i) those that consist of levels within a subsystem, for example the excited states of an atom, and ii) those that consist of a separate subsystem, such as ancillary qubits or measuring devices. For a slow and fast parts described by Hilbert spaces and , the first case corresponds to the Hilbert space (direct sum) while the second case corresponds to the Hilbert space (tensor product). Adiabatic elimination is useful in developing protocols for dissipative state preparation in ion traps Lin et al. (2013); Albert et al. (2019), reservoir engineering Pastawski et al. (2011) and the description of measurement devices Černotík et al. (2015). The simplicity of the resulting equations can also be computationally advantageous in the study of quantum phase transitions where the size of the system is cumbersomly large Minganti et al. (2018).
There are several approaches to obtain effective operators, ranging from perturbative expansions of the Liouville operator Reiter and Sørensen (2012); Kessler (2012), the corresponding Kraus maps Mirrahimi and Rouchon (2009); Azouit et al. (2016, 2017a, 2017b); Azouit (2017), the resolvent Finkelstein-Shapiro et al. (2020), or using stochastic methods Černotík et al. (2015). Eliminating a fast subsystem (that forms a tensor product with the slow subsystem) is typically done with a partial trace over the fast subsystem. This can result in a set of hierarchical equations that allows error estimation and correcting the approximation as the slow and fast timescales get closer. Importantly, the expansion can be built to preserve the Lindblad structure and as a consequence the physicality of the map Azouit et al. (2017a); Lesanovsky and Garrahan (2013); Marcuzzi et al. (2014). The procedure for eliminating sublevels within a subsystem (direct sum with the slow subsystem) is best carried out with Feshbach projections Reiter and Sørensen (2012); Finkelstein-Shapiro et al. (2020). However, as the fast-slow separation breaks down, or when incoherent pumping channels exist, the population of the fast subsystem becomes non-negligible (i.e. there can be a finite fraction of population in the excited states). When this happens, the exact time evolution of the slow part becomes non-trace preserving. The loss of trace can be corrected using contour integral methods Finkelstein-Shapiro et al. (2020). It would be however advantageous to have a method that can handle both classes of fast manifolds. This is more important considering that systems from atomic physics are inspiring a number of chemical versions that have much more complicated Hamiltonians and it would be ideal to transform them into effective operators for a direct comparison to the atomic physics counterparts Castellini et al. (2018); Finkelstein-Shapiro et al. ; Ribeiro et al. (2018).
In this work, we extend the methodology developed in Ref. Finkelstein-Shapiro et al. (2020) to bipartite open quantum systems whose dynamics are described by a Lindblad operator Lindblad (1976); Gorini et al. (1976). We use the projection operator method suggested by Knezevic and Berry Knezevic and Ferry (2002) in order to derive equations for a slow subsystem coupled to a fast subsystem . The paper is organized as follows. We first recall the main results of Ref Finkelstein-Shapiro et al. (2020). We then apply it to the general bipartite case to obtain a recipe for describing the slow subsystem. Finally, we illustrate the method in the case of a spin dispersively coupled to a second highly dissipative driven spin and to describe the dynamics of the open Rabi model.
II Theory
II.1 Adiabatic elimination through projectors techniques
Let be the density operator on the Hilbert space describing the quantum state of the system at time . We suppose that the evolution of is generated by a Lindblad operator : . We define the Hilbert space of operator on , equipped with the scalar product . We first recall the main results presented in Ref Finkelstein-Shapiro et al. (2020) related to projector techniques. Let the projector such that is the the slow part of the system and write , where is the identity operator on . Let be the resolvent of the Lindblad operator . Operators like or are operators on . They are sometimes called super-operators. They are here denoted with calligraphic letter, to distinguish them from operators on (belonging to ), like the density matrix .
We define the effective Lindblad operator , such that . The effective Lindblad operator can be written as :
(1) |
where
(2) |
For any , such that , the slow dynamics inside can be obtained with the inverse Laplace transform as:
(3) |
where and the integral on the complex plane is performed on a straight line . At this point no approximation has been made. Eq. (3) gives the exact dynamics inside , as long as the initial condition is also inside , that is . Because captures the effect of the dynamics in , it is a nonlinear superoperator, in the sense that is a nonlinear eigenvalue problem.
The approximation of a slow dynamics of , with respect to the fast dynamics of is equivalent to considering the dynamics inside in the vicinity of the stationary state reached in the limit . In this long time limit only the limit will contribute to the inverse Laplace transform of Eq. (3). We thus approximate to the lowest relevant order:
(4) |
where and . Using the expression of given by Eq. (1) allows to express and as:
(5) |
In this work, we consider the approximation given by Eq. (4) with only, which is a standard approximation for most of the effective operators that are calculated explicitly. The systematic study of higher order approximations () will be considered in a future work. Within the approximation given by Eq. (4), with , the inverse Laplace transform of Eq. (3) can be computed explicitly. We obtain:
(6) |
The stationary state of this dynamics, reached at , is in the kernel of . We note that although the dynamics described by Eq. (6) is an approximation, the final reached stationary state is the exact one.
To conclude, after the adiabatic elimination of the fast part, the generator of the slow dynamics is approximatively given by , , where and can in principle be computed using Eq. (5). The hard part in these equations is the evaluation of the inverse , which in the most general case, as we will see later, can only be achieved through a perturbative expansion.
Theses results are very general, and require only the definition of a projector and that the initial state fulfills the condition . We note that doesn’t have to be hermitian, that is the projection does not need to be orthogonal.
In Ref. Finkelstein-Shapiro et al. (2020), this formalism was applied to the case where the slow and fast degrees of freedom correspond to a partition of the underlying Hilbert space in two complementary sub-spaces, that is . In the next section we will adapt this general formalism to the bipartite case where .
II.2 Adiabatic elimination in a bipartite system
We suppose that the state of the bipartite system at time is described by a density operator acting on the Hilbert space . We consider that the dynamics of subsystem is very slow compared to the dynamics of subsystem . We suppose that the exact stationary state in is unique and that it is a product state , where and with , the Hilbert space of operators on . As the the dynamics of subsystem is very fast, we suppose that it is a good approximation to consider that at , . In other word, we consider that reaches its steady state instantaneously in the time scale of the subsystem . We thus define as
(7) |
where denotes the partial trace over . The reduced density operator in can then be obtained as .
For the purpose of simplifying some expressions and calculations, it will be useful to use the operator-vector isomorphism Havel (2003) which maps each element of to a vector in as follows. An operators such as is mapped to the vector in the Hilbert space, where is the complex conjugate of . Consequently, any density matrix is mapped to a column vector , with elements, by stacking the columns of the matrix. Under this isomorphism, super-operators on are mapped to operators on . In particular, the super-operator performing the operation , with and operators in , is mapped to , where denotes the complex conjugate of ; that is , where is the adjoint and is the transpose of . In this way, the scalar product between two operators and in is equal to the usual scalar product in . Some useful remarks can be made. The identity operator in is mapped to the maximally entangled state
(8) |
in , where is an orthonormal basis of . We also note that the usual density matrix normalization does not correspond to the normalization induced by the scalar product (except in the case of a pure state). Using the previous remark, we have that is mapped to . For our bipartite case, . Therefore, an operator in as , where , is mapped to . The partial trace over , is mapped to , where in , where is now an orthonormal basis of .
Consequently, the operator is mapped to the vector . The projector acting on can thus be written as:
(9) |
where is the identity operator on . The operator simply reads as:
(10) |
where is the identity operator on .
In general, the Lindblad operator of the system can be split in 3 terms as follows:
(11) |
where is a Lindblad operator acting on the part only. The decomposition of with the help of and reads:
(12) | ||||
(13) | ||||
(14) | ||||
(15) |
Where we have used the fact that , as is a trace preserving operator. In some specific cases, it can be a good approximation to take as , a stationary state of . In that case and in Eq. (14) can be simplified as
(16) |
For computing using Eq. (6), the main difficulty resides in the inversion of . In general this inversion can not be done explicitly, but a perturbative expansion can give a good approximation. In many cases can be divided into two terms, , where the computation of is easy, and is small compared to . In that case we can write
(17) |
Retaining the terms up to or may give a good approximation.
In some cases, the term in Eq. (15) is dominant over the two other terms. In that case, approximating the inverse of by
(18) |
can be sufficient as we will see in the examples in the next section. So, can be approximated by the following expression:
(19) |
This is the main result of this work.
III Examples
We apply the formalism of the preceding section to two examples. We first address the case of a strongly dissipative driven qubit dispersively coupled to a target qubit . Then, as a second example, we consider the open Rabi model in the regime where the dynamics of the spin is very fast compared to the boson frequency.
III.1 A 2-qubit system
This two qubits system has been considered previously by Azouit et in Ref. Azouit et al. (2017a) to test another method of bipartite adiabatic elimination (note that in their work, it is the spin which is the strongly dissipative spin). It consists in a strongly dissipative driven qubit dispersively coupled to a target qubit . This model is used in Ref. Sarlette et al. (2020) to describe the continuous measurement of a harmonic oscillator excitation number (corresponding to system ) by a spin (corresponding system ). The Lindblad equation for the bipartite system can be written as:
(20) |
where , and are the eignvectors of the Pauli matrix with eigenvalues respectively. Defining the new parameters , , and and using the column-vector isomorphism, we write the superoperator form of the Liouvillian as:
where we have used the relations
(21) |
From now on, we will drop the prime in all the parameters , . As the qubit only comes into play through the Hamiltonian term (see Eq. (20)), the kernel of the Lindblad operator is two dimensional, according to the two eigenvectors . Hence the kernel can be considered as the span of , where , and (see Appendix for details):
(22) | ||||
(23) |
In order to avoid any unnecessary complications, we will assume that the qubit has an extremely slow dissipation rate that we will omit from our calculations, but will ensure the uniqueness of the steady state. In other words, the steady state of the considered system will be . We thus assume that the initial state is of the form , and we define the projector (see Eq. (9)):
(24) |
where according to Eq. (8), . This ensures that verifies as it should be. In Appendix A, we calculate in detail all the quantities necessary to compute and as given by Eq. (5) :
(25) | ||||
(26) | ||||
(27) | ||||
(28) |
where , , are operators acting on and we have simplified the notation as , see Appendix. A. From the block diagonal form of Eq. (28), it is relatively easy to invert exactly. In addition, we are only interested in quantities of the form . Hence, from the form of in Eqs.(26), we only need to calculate and .
Using Eqs. (25) to (28) in Eq. (5), we can write and as:
(29) |
(30) |
Since is a projector of rank , we can write:
(31) |
where (see Appendix. A):
(32) |
and, see Appendix A:
(33) |
With these variables and truncating (4) to the zeroth order, the effective Liouville equation can be written in operator space as:
(34) |
Note that approximating the expression of and (given in Appendix. A) by their lowest order in gives for Eq. (34) the same result as the one obtained by Azouit et al. Azouit et al. (2017a) using a completely different method. Finally, it is straightforward to calculate exactly , given that (Eq. (30)) is block diagonal:
(35) |
Let us define the modified initial state of the qubit as :
(36) |
where represent population in the state while represent initial coherences. The physical meaning of this redefined initial density matrix is the state of qubit immediately after qubit has reached its steady-state. In this case this corresponds to a rescaling of the coherences. Using , we can rewrite Eq. (6) to describe the dynamics of the slow qubit as:
(37) |
where . The evolution operator is simple to calculate:
(38) |
where we have defined
(39) |
The evolution of the approximated expectation value of the Pauli matrices for qubit and for and and with the initial state taken as
are compared in Figure 1 to the ones obtained through an exact full numerical propagation. We see that the adiabatic elimination captures the exact dynamics faithfully and that indeed qubit reaches its steady-state before any appreciable dynamics in has taken place.
It is worth mentioning that when adiabatic elimination is valid, i.e. , the exact final state of the fast qubit is very close to the steady state of :
(40) |
Defining the projector leads to considerable simplifications in Eq. (14) where the first term becomes zero. Thus, taking the interaction to be small, we only need the term in Eq. (17) and taking the zero order only in (4), one can check that it leads to the same Lindblad operator derived in Azouit et al. (2017a) which is enough to obtain a very good approximation in the adiabatic limit.


III.2 Open Rabi model
The open Rabi model has been considered recently by Garbe et al. Garbe et al. (2020) in a quantum metrology context. It consist in a spin- (with frequency ) interacting with one bosonic mode (frequency ) of a cavity described by the following Hamiltonian:
(41) |
where () is the annihilation (creation) operator of the bosonic mode. The dynamics of the open Rabi model where the relaxation of the spin (at a rate ) and the photon losses from the cavity (at a rate ) are taken into account is generated by the following Lindblad operator:
(42) |
where we have used the notation
We also assume that and . If we rescale the time by dividing the above equation by , and define:
(43) |
we rewrite the Lindbladian:
(44) |
with :
(45) |
It has been shown that, in the limit where , this model exhibits a quantum phase transition when increases Hwang et al. (2015); Puebla et al. (2017); Hwang et al. (2018); Garbe et al. (2020). The critical point corresponding to separates a normal phase () from a superadiant phase (). Here we show that our method can be used to obtain an effective Lindblad operator for the boson in the normal phase after the elimination of the fast spin. After rescaling, the adiabatic limit corresponds to .
In the normal phase (), the steady state of the system, which is the kernel of the Lindblad operator given by Eq. (42), is separable and unique Garbe et al. (2020). It is straight forward to verify that the steady state of is:
(46) |
Following the same steps of the last example, see Appendix. B, we define the projector and we only calculate the term , the inverse of which corresponds to up to zeroth order in Eq. (17). Simple and straightforward calculations lead to the following Lindbladian evolution of the boson:
(47) |
which is exactly the formula derived in Garbe et al. (2020) using a completely different method, where one should take into consideration that the parameters in (42) are double those considered in Garbe et al. (2020).
IV Conclusion
We have derived a projection based adiabatic elimination method that works for bipartite systems. This work provides a direct connection to earlier work on adiabatic elimination of a subspace of the system Hilbert space Finkelstein-Shapiro et al. (2020) so that in principle now subsystems as well as sublevels can be eliminated at the same time. We have illustrated this with two simple examples of two dispersively coupled spins and the open Rabi model. In both cases, using the lowest order approximations, we have obtain the same expressions that have been previously obtained by completely different methods. We expect that this work will find applications in the case of molecules in cavities where the cavity and part of the molecular levels could be adiabatically eliminated.
Appendix A Detailed Calculation for the 2-qubit system
Here we present in detail all the calculations involved in the example presented in section III.1: first we write in the standard basis:
(48) |
where in this appendix we use the notation to alleviate the complexity of mathematical expressions. Then, we define the necessary projectors of the partial trace, namely:
(49) |
and
(50) |
where we have defined the following vectors:
(51) |
Later on, it will be useful to define
(52) |
and the unitary matrix:
(53) |
as well. From Eqs. (A), (III.1) and (A), we find that:
(54) |
Where we have defined the following matrices on
(55) | ||||
(56) |
(57) |
(58) |
With this diagonal form of , it is straightforward to compute . It consists in computing , , which are matrices. Moreover, since we are solely interested in quantities of the form and is of the form (26), we only need to compute .
To simplify this task, we define the unitary matrix
(59) |
and we compute the pseudo-inverse of . Multiplying by boils down to replacing the ket by in (A), which implies that can be represented as a matrix in the standard basis ”simplifying” the task of finding . To find the pseudo inverse of , we simply solve the set of equations corresponding to:
(60) |
where is the hermitian projector to the range of . If we define the following quantities
(61) | ||||
(62) | ||||
(63) | ||||
(64) | ||||
(65) | ||||
(66) |
and make the identification , and , then:
(67) |
We can check that verify all the Moore-Penrose conditions Penrose (1955):
(68) |
Finally, we can easily see that the pseudo-inverse of is:
(69) |
with that, we have all the necessary ingredients to compute and . Using Eqs.(54), we find that:
(70) |
(71) |
Since is a projector of rank , we can write:
(72) |
Because is of the form , we find that:
(73) |
where we have used the fact that and . We also have:
(74) |
where we have used the fact that , , and . In a similar way we can also show that:
(75) |
Let us define where . A tedious calculation leads to:
(76) |
and
where
(77) |
Finally, it is straightforward to calculate exactly , given that (71) is block diagonal:
(78) |
Taking the fact into account and defining the modified initial state of the qubit as:
(79) |
where represent population in the state while represent initial coherences, we can simplify (6) to describe the dynamics of the slow qubit as:
(80) |
where . The evolution operator is simple to calculate:
(81) |
where we have defined
(82) |
Appendix B open Rabi model
In this section, we carry out all the calculations needed to adiabatically eliminate a fast qubit interacting with a slow cavity mode according to the open Rabi model:
(83) |
The first step is to write in the super-operator representation:
If we define:
(84) |
and
(85) |
then we can compute the needed quantities for
(86) |
represents the dominant term of which can be inverted quite easily:
(87) |
Hence we can easily calculate to be
(88) | |||||
From which we can deduce the reduced dynamics governing the evolution of the slow system to be:
(89) |
where we have defined:
(90) |
References
- Haken (1975) H. Haken, Z Physik B 20, 413 (1975).
- Haken (1977) Haken, Synergetics–An introduction (Springer Berlin, 1977).
- Lax (1967) M. Lax, Phys. Rev. 157, 213 (1967).
- Cohen-Tannoudji (1992) C. Cohen-Tannoudji, Physics Reports 219, 153 (1992).
- Paulisch et al. (2014) V. Paulisch, H. Rui, H. K. Ng, and B.-G. Englert, Eur. Phys. J. Plus 129, 12 (2014).
- Brion et al. (2007) E. Brion, L. H. Pedersen, and K. Mølmer, J. Phys. A: Math. Theor. 40, 1033 (2007).
- You et al. (2003) L. You, X. X. Yi, and X. H. Su, Phys. Rev. A 67, 032308 (2003).
- Nagy et al. (2010) D. Nagy, G. Kónya, G. Szirmai, and P. Domokos, Phys. Rev. Lett. 104, 130401 (2010).
- Douglas et al. (2015) J. S. Douglas, H. Habibian, C.-L. Hung, A. V. Gorshkov, H. J. Kimble, and D. E. Chang, Nature Photonics 9, 326 (2015).
- Sinatra et al. (1995) A. Sinatra, F. Castelli, L. A. Lugiato, P. Grangier, and J. P. Poizat, Quantum Semiclass. Opt. 7, 405 (1995).
- Azouit et al. (2016) R. Azouit, A. Sarlette, and P. Rouchon, arXiv:1603.04630 [quant-ph] (2016), arXiv: 1603.04630.
- Azouit et al. (2017a) R. Azouit, F. Chittaro, A. Sarlette, and P. Rouchon, Quantum Sci. Technol. 2, 044011 (2017a).
- Azouit (2017) R. Azouit, Adiabatic elimination for open quantum systems, Ph.D. thesis, PSL Research University (2017).
- Azouit et al. (2017b) R. Azouit, F. Chittaro, A. Sarlette, and P. Rouchon, IFAC-PapersOnLine 20th IFAC World Congress, 50, 13026 (2017b).
- Sarlette et al. (2020) A. Sarlette, P. Rouchon, A. Essig, Q. Ficheux, and B. Huard, (2020), arXiv:2001.02550 .
- Mirrahimi and Rouchon (2009) M. Mirrahimi and P. Rouchon, IEEE Transactions on Automatic Control 54, 1325 (2009).
- Reiter and Sørensen (2012) F. Reiter and A. S. Sørensen, Phys. Rev. A 85, 032111 (2012).
- Kessler (2012) E. M. Kessler, Phys. Rev. A 86, 012126 (2012).
- Lin et al. (2013) Y. Lin, J. P. Gaebler, F. Reiter, T. R. Tan, R. Bowler, A. S. Sørensen, D. Leibfried, and D. J. Wineland, Nature 504, 415 (2013).
- Albert et al. (2019) V. Albert, K. Noh, and F. Reiterr, arXiv:1809.07324 (2019).
- Pastawski et al. (2011) F. Pastawski, L. Clemente, and J. I. Cirac, Phys. Rev. A 83, 012304 (2011).
- Černotík et al. (2015) O. c. v. Černotík, D. V. Vasilyev, and K. Hammerer, Phys. Rev. A 92, 012124 (2015).
- Minganti et al. (2018) F. Minganti, A. Biella, N. Bartolo, and C. Ciuti, Physical Review A 98, 042118 (2018).
- Finkelstein-Shapiro et al. (2020) D. Finkelstein-Shapiro, D. Viennot, I. Saideh, T. Hansen, T. o. Pullerits, and A. Keller, Phys. Rev. A 101, 042102 (2020).
- Lesanovsky and Garrahan (2013) I. Lesanovsky and J. P. Garrahan, Phys. Rev. Lett. 111, 215305 (2013).
- Marcuzzi et al. (2014) M. Marcuzzi, J. Schick, B. Olmos, and I. Lesanovsky, Journal of Physics A: Mathematical and Theoretical 47, 482001 (2014).
- Castellini et al. (2018) A. Castellini, H. R. Jauslin, B. Rousseaux, D. Dzsotjan, G. Colas des Francs, A. Messina, and S. Guérin, The European Physical Journal D 72, 223 (2018).
- (28) D. Finkelstein-Shapiro, P.-A. Mante, S. Sarizosen, L. Wittenbecher, I. Minda, S. Balci, T. Pullerits, and D. Zigmantas, arXiv:2002.05642 .
- Ribeiro et al. (2018) R. F. Ribeiro, L. A. Martínez-Martínez, M. Du, J. Campos-Gonzalez-Angulo, and J. Yuen-Zhou, Chem. Sci. 9, 6325 (2018).
- Lindblad (1976) G. Lindblad, Communications in Mathematical Physics 48, 119 (1976).
- Gorini et al. (1976) V. Gorini, A. Kossakowski, and E. C. G. Sudarshan, Journal of Mathematical Physics 17, 821 (1976).
- Knezevic and Ferry (2002) I. Knezevic and D. K. Ferry, Phys. Rev. E 66, 016131 (2002).
- Havel (2003) T. F. Havel, Journal of Mathematical Physics 44, 534 (2003).
- Garbe et al. (2020) L. Garbe, M. Bina, A. Keller, M. G. A. Paris, and S. Felicetti, Phys. Rev. Lett. 124, 120504 (2020).
- Hwang et al. (2015) M.-J. Hwang, R. Puebla, and M. B. Plenio, Physical Review Letters 115, 180404 (2015).
- Puebla et al. (2017) R. Puebla, M.-J. Hwang, J. Casanova, and M. B. Plenio, Physical Review Letters 118, 073001 (2017), publisher: American Physical Society.
- Hwang et al. (2018) M.-J. Hwang, P. Rabl, and M. B. Plenio, Physical Review A 97, 013825 (2018), publisher: American Physical Society.
- Penrose (1955) R. Penrose, Mathematical Proceedings of the Cambridge Philosophical Society 51, 406 (1955).