Abstract
We consider the solution of the equation of motion of a classical/quantum spin subject to a monochromatical, elliptically polarized external field. The classical Rabi problem can be reduced to third-order differential equations with polynomial coefficients and hence solved in terms of power series in close analogy to the confluent Heun equation occurring for linear polarization. Application of Floquet theory yields physically interesting quantities like the quasienergy as a function of the problem’s parameters and expressions for the Bloch–Siegert shift of resonance frequencies. Various limit cases are thoroughly investigated.
1 Introduction
In recent years, theoretical and experimental evidence has shown that periodic driving can be a key element for engineering exotic quantum mechanical states of matter, such as time crystals and superconductors at room temperature [1], [2], [3]. The renewed interest in Floquet engineering, i.e., the control of quantum systems by periodic driving, is due to (a) the rapid development of laser and ultrashort spectroscopy techniques [4], (b) the discovery and understanding of various “quantum materials” that exhibit interesting exotic properties [5], [6], and (c) the interaction with other emerging fields of physics such as programmable matter [7] and periodic thermodynamics [8], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18], [19], [20], [21], [22].
One of the simplest system to study periodic driving is a two level system (TLS) interacting with a classical periodic radiation field. The special case of a constant magnetic field in, say, x-direction plus a circularly polarized field in the y−z-plane was already solved more than eight decades ago by Rabi [23] and can be found in many textbooks. This case is referred to as Rabiproblem withcircular polarization (RPC) in the following. Shortly thereafter, Bloch and Siegert [24] considered the analogous problem of a linearly polarized magnetic field orthogonal to the direction of the constant field henceforth called, Rabi problem with linear polarization (RPL) and proposed the so-called rotating wave approximation. They also investigated the shift of the resonant frequencies due to the approximation error of the rotating wave approximation, since then called the “Bloch–Siegert shift.”
In the following decades, one noticed [25], [26] that the underlying mathematical problem leads to the Floquet Theory [27], which deals with linear differential matrix equations with periodic coefficients [28], [29]. Accordingly, analytical approximations for solutions were worked out, which formed the basis for subsequent research. In particular, the groundbreaking work of Shirley [26] has received widespread attention and many citations. Among the numerous applications of the theory of periodically driven TLS are nuclear magnetic resonance [30], ac-driven quantum dots [31], Josephson qubit circuits [32], and coherent destruction of tunneling [33]. On a theoretical level, the methods for solving the RPL and related problems have been gradually refined and include power series approximations for Bloch–Siegert shifts [34], [35], perturbation theory and/or various boundary cases [36], [37], [38], [39], [40], [41] and the hybridized rotating wave approximation [42]. Also, the inverse method yields analytical solutions for certain periodically driven TLS [43], [44], [45], [46].
In the meantime also the RPL has been analytically solved [47], [48]. This solution is based on a transformation of the Schrödinger equation into a confluent Heun differential equation. A similar approach was previously applied to the TLS subject to a magnetic pulse [49], [50] and has been extended to other cases of physical interest [51], [52]. In the special case of the RPL, the analytical solution has been further elaborated to include time evolution over a full period and explicit expressions for the quasienergy [53].
In this paper, we will extend these results to the Rabi problem with elliptic polarization (RPE) that is also of experimental interest, see [54], [55]. Here, we will approach the Floquet problem of the TLS via its well-known classical limit, see, e.g., [56]. It has been shown that, for the particular problem of a TLS with periodic driving, the classical limit is already equivalent to the quantum problem [57], [58]. More precisely, to each periodic solution of the classical equation of motion, there exists a Floquet solution of the original Schrödinger equation that can be explicitly calculated via integrations. Especially, the quasienergy is essentially given by the action integral over one period of the classical solution. This is reminiscent of the semiclassical Floquet theory developed in a study by Breuer et al. [59].
The motion of a classical spin vector S(τ) in a monochromatical magnetic field with elliptic polarization and an orthogonal constant component can be analyzed by following an approach analogous to that leading to the confluent Heun equation in a study by Ma and Li [47] and Xie and Hai [48]. We differentiate the first-order equation of motion twice and eliminate two components of the spin vector. The resulting third-order differential equation for the remaining component x(τ) can be transformed into a differential equation with polynomial coefficients by the change from the dimensionless time variable τ to
The structure of the paper is the following. In Section 2, we present the scenario of the classical Rabi problem with elliptic polarization and its connection to the underlying Schrödinger equation. The abovementioned reduction of the time evolution to the first quarter period is made in Section 3. Already in the following Section 4, before solving the equation of motion, it can be shown that the fully periodic monodromic matrix depends only on two parameters r and α, which determine the quasienergy and the initial value of the periodic solution S(τ), respectively. The Fourier series of this solution necessarily have the structure of an even/odd cos-series for
The quasienergy is discussed in more details in Section 8 with the emphasis on curves in parameter space where it vanishes. The resonance frequencies
2 The classical Rabi problem: general definitions and results
We consider the Schrödinger equation
of a spin with quantum number s = 1/2,
where the σi, i = 1, 2, 3, are the Pauli matrices
Hence H(t) can be understood as a Zeeman term w. r. t. a (dimensionless) magnetic field
Alternatively,
Setting ℏ = 1 and passing to a dimensionless time variable τ = ωt we may rewrite (1) in the form
where G = gω, F = fω and ω0 = νω. The dimensionless period is always Tω = 2π. Sometimes, we will denote the derivative w. r. t. τ by an overdot
Let
denote the one-dimensional time-dependent projector onto a solution of (5) and
its expansion w. r. t. the basis
and hence S(τ) can be viewed as a classical spin vector (not necessarily normalized). Moreover,
denotes the dimensionless magnetic field vector (4) written as a function of τ.
Conversely, to each solution of (8) one obtains the corresponding solution of (5) up to a time-dependent phase that can be obtained by an integration, see [57] for the details.
The coefficients of the Taylor series w. r. t. τ of x(τ), y(τ) and z(τ) can be recursively determined by using (8) and the initial values x(0), y(0) and z(0). Note that h1 and h2 are even functions of τ and that h3 is an odd one. Hence, there exist special solutions of (8) such that x(τ) and y(τ) are even functions of τ and z(τ) is an odd one, symbolically:
In fact, this is consistent with (8) and (9) since
and
and can be proven by induction over the degree of the Taylor series coefficients of S(τ) using the necessary initial condition z(0) = 0.
Analogously, there exist solutions S(τ) of type
satisfying x(0) = y(0) = 0. We will state these results in the following form:
Proposition 1:
1. The solutionS(τ) of(8)is of type(10)iff z(0) = 0.
For general initial conditions, the solution S(τ) of (8) will be of mixed type.
Next, let
with initial condition
Here,
The differential Equation (14) with initial condition (15) has a unique solution R(τ, τ0) for all τ, τ0 ∈ ℝ, see, e.g., theorem 3.9 in [29]. Obviously, this implies the composition law
and hence
for all
Usually, we will set τ0 = 0. The matrix H(τ) is obviously 2π-periodic. Hence, we may apply Floquet theory to the classical equation of motion (8). The monodromy matrix R(2π, 0) has the eigenvalues
uniquely defined up to integer multiples (note that effectively ω = 1 in our approach).
The connection to the quasienergy
Taking into account the mentioned ambiguity of
Another way to obtain
where the overline indicates the time average over one period of a 2π-periodic solution S(τ) of (8). An equivalent expression, that is manifestly invariant under rotations, is given by
see Eq. (46) in [58]. Periodic solutions of (8) can be found by using the initial value S(0) = r, where r is the normalized eigenvector of R(2π, 0) corresponding to the eigenvalue 1, see also in the study by Schmidt et al. [58].
Of course, both ways, (20) and (21), to obtain
3 Reduction to the first quarter period
Due to the discrete symmetries of the polarization ellipse, it is possible to reduce the time evolution of the classical spin to the first quarter period
and T(ij) ≡ T(i)T(j), for example,
First, we will formulate a proposition that allows us to reduce the time evolution for the classical spin from the full period to the first half period
Proposition 2:
for all
Proof: Let
In (29) we have used that sin (π+τ) = −sin τ, cos (π+τ) = −cos τ and hence
It follows that
which completes the proof of the proposition.
Setting τ = π in (25) gives
Next, we show how to further reduce the time evolution to the first quarter period
Proposition 3:
Proof: The proof is similar to that of proposition 2 except that an additional time reflection is involved. Let
In (39), we have used that sin (π−τ) = sin τ, cos (π−τ) = −cos τ and hence
It follows that
Consequently,
which completes the proof of the proposition.
Setting τ = π in (43) implies
and hence
Moreover, if we set
and hence, solving for R(π, 0),
Thus (35) can be re-written as
and hence the evolution data for
4 Fourier series and quasienergy: preliminary results
First we will re-derive (46) under more general assumptions.
Proposition 4:
Let R ∈ SO(3) and T ∈ O(3) be such that T2 = 1 and hence T⊤ = T. Define
Proof:
Let us specialize to the case T = T(13), then (51) is equivalent to the following three equations:
A general rotational matrix
Proposition 5:
Every rotational matrix
Proof: Obviously, the third column
We have still to consider the case r = 0 such that
which is obviously the most general case satisfying (52) and
Recall that the “half period monodromy matrix” R(π, 0) satisfies (46), hence, according to Prop. 4, also (52) and, by virtue of Prop. 5, must be of the form (53). In the case of linear polarization (
It will be instructive to sketch another derivation of (55). To this end, we state without proof that the monodromy matrix of the Schrödinger Equation (5) will assume the form
completely analogous to Eq. (33) of [53]. U has the eigenvalues exp(±2i arcsin r) with respective eigenvectors
where the
Like R(π, 0) also R(2π, 0) depends only on two parameters α and r and satisfies a similar equation that characterizes the corresponding two-dimensional submanifold of SO(3), to wit,
This equation can be proven either directly by checking (55) or by applying (34) and (46).
According to the general theory [57], the eigenvalues of R(2π, 0) that are generally of the form
As in [53] it follows that
The eigenvector
Choosing r as the initial value r = S(0) for the time evolution (8) yields a 2π-periodic solution. Any other unit vector in the plane P orthogonal to r will, in general, not return to its initial value after the time τ = 2π but will be rotated in the plane P by the angle
Another remarkable result follows from R(π, 0) being of the form (53):
which means that for the initial value S(0) = r the half period time evolution is equivalent to a reflection at the x−z-plane. This has further consequences for the Fourier series of the 2π-periodic functions x(τ), y(τ), and z(τ) with initial values
Now consider the sequence of linear mappings
From this we conclude
Hence, the odd terms of the cos-series must vanish and x(τ) is an even cos-series. Similarly, we conclude from (66) that y(τ) is an odd cos-series and z(τ) an odd sin-series. Summarizing, we have proven the following
5 Third order differential equations for single spin components
We consider again (8) and its higher derivatives that read
with
and
It is obvious that
Similarly, we can obtain third-order differential equations for y(τ) and z(τ). For the preparation of the next step, we make the restriction to solutions of (71) such that x(τ) and y(τ) are even functions of τ, whereas z(τ) is an odd one, according to Prop. 1. In this way, we could obtain two solutions S(1) and S(2) with different initial conditions for x(τ) and y(τ) and the initial condition z(0) = 0, the latter being a consequence of the restriction to odd functions z(τ). The third solution S(3) with x(τ) and y(τ) odd and z(τ) even is then uniquely determined by S(1) and S(2). For example, if S(1) and S(2) are chosen to be orthogonal for τ = 0 then they will be orthogonal for all τ and S(3) is just the vector product of S(1) and S(2).
Following the study by Xie and Hai [48], we will consider a transformation
the same as in the study by Xie and Hai [48], and maps the half period
The remaining function z(τ) has to be calculated differently, e.g., by using that the length of S(τ) is conserved under time evolution according to (8). This gives the result
where z(0) = 0 has been used, and the sign has to be chosen in such a way that z(τ) remains a smooth function in the neighborhood of its zeros. An alternative procedure would be possible if x(τ) and y(τ) can be written as Fourier series (maybe only locally valid for
We come back to the differential equation for X(u) and write it with polynomial coefficients pn(u) in the form
The coefficients are the following ones:
The singular points of the differential equation are the zeros of p3(u). Except the points u = 0 and u = 1 that occur also for the confluent Heun equation, see [48] and [53], we have an additional pair of singular points, real or complex ones, depending on the parameters f, g and ν. The obvious ansatz to obtain a physically relevant solution of (80) is a power series
at the singular point given by u = 0. We have not investigated its radius of convergence, but it is clear that the series diverges at least for the second singular point u = 1, which has been our motivation to restrict the application of (85) to
We will not give the details of the recurrence relation but rather sketch how to obtain it by means of computer-algebraic aids. We take a finite part
where the ai have been determined as rational functions of f, g, ν, but they are too complicated to be presented here.
The next problem is that we need the first five coefficients of
Obviously,
After these preparations it is, in principle, possible to calculate any finite number of power series coefficients ξn as a function of the physical parameters f, g and ν and the initial values x(0) and y(0), although the expressions become more and more intricate, and finally to obtain a truncated approximation of X(u(τ)). For a comparison to a numerical solution of (8) see Section 8.
Analogous considerations apply for the case of the solution y(τ) = Y(u(τ)). This time we obtain a differential equation of the form
where
The zeros of q3(u) yield five singular points. The power series solution ansatz
leads to a 9-term recursion relation and the first eight coefficients are again determined by calculating the corresponding t-power series coefficients. We show the first three ones.
Analogously to (90), ηn is a linear function of x(0) and y(0) that can be written as
The further details are too intricate to be displayed here, but, in principle, the procedure is completely analogous to the power series solution of the confluent Heun equation investigated in [53].
6 Fourier series and quasienergy: results based on the power series solutions
It is clear that
where
to wit,
Recall that the
The case of z(τ) is a bit more complicated. Using the above local Fourier series representation of x(τ) and y(τ), we may directly solve the differential equation
since the r. h. s. of (106) is again a cos-series. In general, there will be a nonvanishing constant term z0 at the r. h. s. of (106) that generates a corresponding part
The complete result is the following:
The expressions (104) and (105) for the Fourier coefficients still depend, via
This equation can be solved for the auxiliary parameter α:
if the numerator and denominator of this fraction do not vanish simultaneously. This solution is only determined modulo π in accordance with the fact that x(0) = −cosα, y(0) = −sinα also gives a periodic solution.
The determination of the second auxiliary parameter r is more involved. We consider the following procedure that does not presuppose the determination of α. First, we calculate the quarter period monodromy matrix
The latter holds since R(π, 0) is of the form (53). It will be instructive to give some more details.
First consider
because the only nonvanishing terms are
The procedure for the calculation of
For the second column of
and hence
where the entries from the first and second column of
We have checked the results (110) and (115) by comparison with a numerical solution of the
7 Time evolution: an example
As an example, we consider the time evolution over one period

The three components of the classical spin vector as functions of dimensionless time τ over one period according to the equation of motion (8). We have chosen the parameters
The alternative choice of the initial conditions as x(0) = cos α and y(0) = sin α, whereas z(0) = 0 remains unchanged, leads to 2π-periodic solutions, see Figure 2. This calculation uses the value of the auxiliary parameter α that has been determined according to (110).

The three components of the classical spin vector as periodic functions of dimensionless time τ over one period according to the equation of motion (8). We have chosen the parameters
The first few terms of the corresponding Fourier series read as follows:
8 Vanishing of the quasienergy
We will discuss the quasienergy in physical units
where usually ℏ is set to 1. Analogously to the ambiguity of
We notice that these curves qualitatively all look the same. First we note that the family of curves shows the same asymptotic behavior of ℰ(ω0, F, G, ω) for ω → ∞. In the case of circular polarization, we have

The branch of the quasienergy
To better understand this behavior in detail, we revisit the RPC. It is well known that in the special case of circular polarization the quasienergy can be analytically determined in a rather simple form. In the context of the present discussion, we note that the fundamental matrix solution of (14) with initial condition (15) assumes the form
where we have used the abbreviation
known as the “Rabi frequency”. The corresponding monodromy matrix R(2π, 0) reads:
Its trace is evaluated as
On the other hand, we may apply (21) to the periodic solution
with the well-known result
that is compatible with (127) and (20).
For

The branch of the quasienergy
The vanishing of the quasienergy is in so far interesting as it means that all solutions of (8) will be 2π-periodic, not only the special one with the initial condition S(0)=(cos α, sin α, 0)⊤ according to (61). Moreover,
The vanishing of the quasienergy implies that the linear term z0τ in (107) must vanish and hence
In order to check the consistency, we will evaluate the condition (130) by using a truncation of the power series solutions (85) and (96) to the first 10 terms. This yields the exact first five terms of G = ℰ0 (ω; 1, 1) expanded into a power series in terms of
The result is shown in Figure 3 as a black dashed curve and fits to the numerically determined red dashed curve of vanishing quasienergy in the domain
Further we note that according to [57] the quasienergy ℰ can be split into a geometrical part ℰg and a dynamical part ℰd such that ℰ = ℰg + ℰd and the slope relation
holds, see Eq. (151) in a study by Schmidt [57]. Recall that
In general the quasienergy (129) of the RPC has its first zero at
but the next terms are too intricate to be shown here.
We will consider the case of vanishing quasienergy along the curve
Recall that for a general periodic, not necessarily normalized solution
Analogous decompositions for
It is clear that
For the sake of completeness, we note that the third solution S(3) = S(1)×S(2) will be of the following type: x(3) (τ) is an odd sin-series, y(3) (τ) is an even sin-series, and

Three periodic solutions S(i)(τ), i = 1, 2, 3 (blue, red and green curves) of (8) for the parameters
In the case of the periodic solutions
where we have again passed to the dimensionless quasienergy and the Fourier coefficients are given in (104), (105) and (107). The suitable initial conditions
Here, the superscript (x) or (y) refers to the dependence of the Fourier coefficients, via
using the orthogonality of S(1)(τ) and S(2)(τ). After some calculations, it follows that the dynamical part of the quasienergy of the first solution S(1)(τ) assumes the value
As an example we consider the parameters ω0 = F = 1 and
9 Resonances
The function
defining an infinite number of hypersurfaces in the parameter space
In the circular case, a smooth representative of the quasienergy
and has a unique maximum at ω0 = ω, see Figure 7. This conforms with the intuitive picture that a resonance occurs if the driving frequency ω equals the Larmor frequency ω0 of the energy level splitting. The other maxima of the quasienergy, restricted to the domain (121), are represented by intersections of suitable branches of the quasienergy of the form ±ℰc + nω, n ∈ ℤ. For example, the next maximimum at
Note that an arbitrarily small admixture of eccentricity to the polarization leads to an avoided level crossing and a smooth maximum close to the value ω0 of the intersection, see Figure 7.

The quasienergy ℰ(ω0, F, G, ω) as a function of ω0 for fixed values of ω = 1, F = 0.5 and G = 0.1 (solid curves) calculated by numerical solutions of the Schrödinger Equation (5). One observes maxima of the quasienergy at
According to [26], the time average of the transition probability between different Floquet states assumes its maximum value
where α is the auxiliary parameter leading to a periodic solution given by (110). Together with
see (109), this implies that the matrix
has a nonvanishing null-vector and hence
We use truncated versions of (85) and (96) in order to derive the first terms of the power series representations
analogously to [57]. We will show a few results. The first resonance
We note that
The second resonance is described by the matrix
Here, analogous remarks apply as in the case of
where
It can be easily checked that the coefficients of the power series (155) coincide with the antidiagonal sums, i.e.,
Finally, we consider the third resonance described by
the antidiagonal sums of which are obtained via
The first nontrivial antidiagonal of
for n > 1. Employing the circular limit
we obtain
for n > 1. These results can be checked for n = 2, 3 by inspection of (151) and (156).
10 Special limit cases
10.1 Limit case ω → 0
This limit case (“adiabatic limit”) has been already treated in [57] in sufficient generality, such that we only need to recall the essential issues. We adopt a series representation
of the periodic solution of (8) and obtain a recursive system of inhomogeneous linear differential equations for the Sn. The starting point is
that is, the spin vector follows the direction of the slowly varying magnetic field. The corresponding zeroth term of the series for the quasienergy
can be obtained as
where E(…) denoted the complete elliptic integral of the second kind. Note that in the adiabatic limit the quasienergy
yields the zeroth order contribution
The next term S(1) of the series (163) is obtained as the solution of
such that S(0)⋅S(1) = 0 in order to guarantee normalization in linear ω-order. The result is
It leads to a linear contribution to the quasienergy of
where
According to [57], the next, quadratic term of (163) is given by
where
The corresponding quadratic correction to the quasienergy is too complicated to be calculated here. We confine ourselves to determine
The corresponding adiabatic approximation of the quasienery has been shown in Figure 8 together with the various branches of the form

Various branches of the quasienergy
10.2 Limit case F , G → 0
For sake of comparison with the analogous results in [57], we rewrite the equation of motion (8) in the form
where λ is a formal expansion parameter that is ultimately set to
In the case
Inserting these series into the differential Equations (183)–(185) and collecting powers of λ yields recurrence relations for the functions
For n > 0, the FT coefficients Rn,m, Sn,m and Tn,m can be recursively determined by means of the following relations:
where, of course, we have to set
We will show the first few terms of the FT series for X(t), Y(t) and Z(t):
where λ stands for any linear combination of F and G. We note that the coefficients contain denominators of the form
Using the FT series solution (186)–(188), it is a straightforward task to calculate the quasienergy
according to (21). The first few terms of the result are given by
This is in agreement with the result for linear polarization, see [57], Equation (198), if we set G = 0.
It will be instructive to check the first two terms of (200) by using the decomposition of the quasienergy into a dynamical and a geometrical part. In lowest order in λ, the classical RPE solution is a motion on an ellipse with semi axes
Hence, the geometrical part of the quasienergy reads
The dynamical part is obtained as
The sum of both parts together correctly yields
Moreover, the slope relation (132) is satisfied in the considered order,
in accordance with [57], Equation (202).
However, as mentioned above, the FT series for the quasienergy has poles at the values
10.3 Limit case ω 0 → 0
It is well known, see, e.g., [57] or [53], that for
where the
where
Unfortunately, it does not seem possible to generalize the above
10.3.1 Limit case ω 0 = 0 and F → 0
We can only get a result for “almost linear” polarization, i.e., in the lowest linear order of F. To achieve this result, we first note that for
Now we can perform the sum over n = 0, …, ∞ without any problems:
which, finally, yields
where we have multiplied the result by

The periodic solution of the equation of motion (183)–(185) with the values of the parameters
10.3.2 Limit case ω 0 = 0 and F → G
The Rabi problem with circular polarization (F = G) has two simple periodic (not yet normalized) solutions of (183)–(185), namely
see, e.g., [57], Equation (69). Let us consider the special solution for
and look for corrections in linear order of the parameter δ describing eccentricity, namely
To this end, we insert ω0 = 0 and G = F−δ into the FT series solution (186)–(188) and expand the FT series coefficients up to terms linear in δ. The δ = 0 parts of the coefficients satisfy
in accordance with (216). The δ-linear parts are given by
It is straightforward to perform the summations over n and to insert the results into (186)–(188) thus obtaining the analytical approximations
The quality of these approximations is surprisingly high, see Figure 10, where a deviation between analytical approximation and numerical integration is only visible for

The periodic solution of the equation of motion (183) – (185) with the values of the parameters
11 Application: work performed on a two level system
As an application of the results obtained in the preceding sections, we consider the work performed on a TLS by an elliptically polarized magnetic field during one period. For a related experiment, see [63]. In contrast to classical physics, this work is not just a number but, following [64], has to be understood in terms of two subsequent energy measurements. At the time τ = 0, the TLS is assumed to be in a mixed state according to the canonical ensemble
with dimensionless inverse temperature
Then at the time τ = 0, one performs a Lüders measurement of the instantaneous energy H(0) with the two possible outcomes
and
After this measurement, the system evolves according to the Schrödinger Equation (1) with Hamiltonian H(τ). At the time τ = 2π, the system hence is in the pure state
such that
see [53], Equation (55). A detailed investigation of the work statistics is beyond the scope of the present article. We will only give an example of the frequency dependence of

The mean value
However, a clear difference to the situation dealt with in Section 9 is that for small amplitudes the frequency
depending on the eccentricity of the elliptic polarization. The small amplitude limit
see Figure 11 for an example.
12 Summary and outlook
The time evolution of the TLS subject to a monochromatic, circularly polarized external field (RPC) can be solved in terms of elementary functions, and the analogous problem with linear polarization (RPL) leads to the confluent Heun functions. However, these two problems are only limit cases of the general Rabi problem with elliptical polarization (RPE), and it is a natural question to look for a solution of the latter valid in the realm where the rotating wave approximation breaks down. This is done in the present paper by performing the following steps:
Reduction to the classical RPE,
reduction of the classical time evolution to the first quarter period,
transformation of the classical equation of motion to two third order differential equations, and
solution of the latter by power series.
This strategy has been checked by comparison with the numerical integration of the equations of motion for an example. Moreover, we have calculated the various Fourier series of the components of the periodic solution and the corresponding quantum or classical Floquet exponent (or quasienergy). Further, we have obtained the first terms of the power series for the resonance frequencies w. r. t. the semi-axes F and G of the polarization ellipse. The latter were checked by comparison with the partially known results in the circular (F = G) and in the linear polarization limit (G = 0). This kind of result could not be obtained by a pure numerical treatment of RPE and thus justifies our analytical approach. Analogous remarks apply to the problem of how much work is performed on a TLS by the driving field. For a first overview numerical methods are sufficient, see Figure 11, but analytical methods yield more detailed results, e.g., for the small amplitude limit, see Section 11.
Other limit cases that can be discussed without recourse to the third-order differential equation are the adiabatic limit (ω → 0), the small amplitude limit (F, G → 0) and the limit of vanishing energy splitting of the TLS (ω0 → 0). In the latter case, it turns out that the exact solution of the special case
It appears that this completes the set of problems related to the RPE that can be addressed with the present methods, with one exception: In principle, it would also be possible to solve the underlying
Acknowledgments
The author is indebted to the members of the DFG research group FOR 2692 for continuous support and encouragement, especially to Martin Holthaus and Jürgen Schnack. Moreover, the author gratefully acknowledge discussions with Thomas Bröcker on the subject of this paper.
Author contribution: The author has accepted responsibility for the entire content of this submitted manuscript and approved submission.
Research funding: None declared.
Conflict of interest statement: The author declares no conflicts of interest regarding this article.
References
[1] T. Oka and S. Kitamura, “Floquet engineering of quantum materials,” Annu. Rev. Condens. Matter Phys., vol. 10, pp. 387–408, 2019, https://doi.org/10.1146/annurev-conmatphys-031218-013423.Search in Google Scholar
[2] M. Holthaus, “Floquet engineering with quasienergy bands of periodically driven optical lattices,” J. Phys. B Atom. Mol. Opt. Phys., vol. 49, 2016, 013001, https://doi.org/10.1088/0953-4075/49/1/013001.Search in Google Scholar
[3] M. S. Rudner and N. H. Lindner, The Floquet Engineer’s Handbook, 2020, arXiv:2003.08252v1 [cond-mat.mes-hall].Search in Google Scholar
[4] M. Bukov, L. D’Alesion, and A. Polkovnikov, “Universal high-frequency behavior of periodically driven systems: from dynamical stabilization to Floquet engineering,” Adv. Phys., vol. 64, no. 2, pp. 139–226, 2015, https://doi.org/10.1080/00018732.2015.1055918.Search in Google Scholar
[5] M. S. Rudner and N. H. Lindner, Floquet Topological Insulators: From Band Structure Engineering to Novel Non-equilibrium Quantum Phenomena, 2019, arXiv:1909.02008v1 [cond-mat.mes-hall].Search in Google Scholar
[6] C. J. Fujiwara, K. Singh, Z. A. Geiger, et al., “Transport in Floquet-Bloch bands,” Phys. Rev. Lett., vol. 122, 2019, 010402, https://doi.org/10.1103/physrevlett.122.010402.Search in Google Scholar
[7] T. Toffoli and N. Margolus, “Programmable matter: concepts and realization,” Physica D, vol. 47, nos 1–2, pp. 263–272, 1991, https://doi.org/10.1016/0167-2789(91)90296-l.Search in Google Scholar
[8] W. Kohn, “Periodic thermodynamics,” J. Stat. Phys., vol. 103, p. 417, 2001, https://doi.org/10.1023/a:1010327828445.10.1023/A:1010327828445Search in Google Scholar
[9] H.-P. Breuer, W. Huber, and F. Petruccione, “Quasistationary distributions of dissipative nonlinear quantum oscillators in strong periodic driving fields,” Phys. Rev. E, vol. 61, p. 4883, 2000, https://doi.org/10.1103/physreve.61.4883.Search in Google Scholar
[10] R. Ketzmerick and W. Wustmann, “Statistical mechanics of Floquet systems with regular and chaotic states,” Phys. Rev. E, vol. 82, 2010, 021114, https://doi.org/10.1103/physreve.82.021114.Search in Google Scholar
[11] D. W. Hone, R. Ketzmerick, and W. Kohn, “Statistical mechanics of Floquet systems: the pervasive problem of near-degeneracies,” Phys. Rev. E, vol. 79, 2009, 051129, https://doi.org/10.1103/physreve.79.051129.Search in Google Scholar
[12] M. Langemeyer and M. Holthaus, “Energy flow in periodic thermodynamics,” Phys. Rev. E, vol. 89, 2014, 012101, https://doi.org/10.1103/physreve.89.012101.Search in Google Scholar
[13] G. Bulnes Cuetara, A. Engel, and M. Esposito, “Stochastic thermodynamics of rapidly driven systems,” New J. Phys., vol. 17, 2015, 055002, https://doi.org/10.1088/1367-2630/17/5/055002.Search in Google Scholar
[14] T. Shirai, T. Mori, and S. Miyashita, “Condition for emergence of the Floquet-Gibbs state in periodically driven open systems,” Phys. Rev. E, vol. 91, 2015, 030101(R), https://doi.org/10.1103/physreve.91.030101.Search in Google Scholar
[15] D. E. Liu, “Classification of the Floquet statistical distribution for time-periodic open systems,” Phys. Rev. B, vol. 91, 2015, 144301, https://doi.org/10.1103/physrevb.91.144301.Search in Google Scholar
[16] T. Iadecola, T. Neupert, and C. Chamon, “Occupation of topological Floquet bands in open systems,” Phys. Rev. B, vol. 91, 2015, 235133, https://doi.org/10.1103/physrevb.91.235133.Search in Google Scholar
[17] K. I. Seetharam, C.-E. Bardyn, N. H. Lindner, M. S. Rudner, and G. Refael, “Controlled population of Floquet-Bloch states via coupling to Bose and Fermi baths,” Phys. Rev. X, vol. 5, 2015, 041050, https://doi.org/10.1103/physrevx.5.041050.Search in Google Scholar
[18] D. Vorberg, W. Wustmann, H. Schomerus, R. Ketzmerick, and A. Eckardt, “Nonequilibrium steady states of ideal bosonic and fermionic quantum gases,” Phys. Rev. E, vol. 92, 2015, 062119, https://doi.org/10.1103/physreve.92.062119.Search in Google Scholar
[19] S. Vajna, B. Horovitz, B. Dóra, and G. Zaránd, “Floquet topological phases coupled to environments and the induced photocurrent,” Phys. Rev. B, vol. 94, 2016, 115145, https://doi.org/10.1103/physrevb.94.115145.Search in Google Scholar
[20] H.-J. Schmidt, J. Schnack, and M. Holthaus, “Periodic thermodynamics of the Rabi model with circular polarization for arbitrary spin quantum numbers,” Phys. Rev. E, vol. 100, 2019, 042141, https://doi.org/10.1103/physreve.100.042141.Search in Google Scholar
[21] H.-J. Schmidt, “Periodic thermodynamics of a two spin Rabi model,” J. Stat. Math., vol. 2020, 2020, 043204, https://doi.org/10.1088/1742-5468/ab7812.Search in Google Scholar
[22] O. R. Diermann, H.-J. Schmidt, J. Schnack, and M. Holthaus, “Environment-controlled Floquet-state paramagnetism,” Phys. Rev. Research, vol. 2, 2020, 023293, https://doi.org/10.1103/physrevresearch.2.023293.Search in Google Scholar
[23] I. I. Rabi, “Space quantization in a gyrating magnetic field,” Phys. Rev., vol. 51, p. 652, 1937, https://doi.org/10.1103/physrev.51.652.Search in Google Scholar
[24] F. Bloch and A. Siegert, “Magnetic resonance for nonrotating fields,” Phys. Rev., vol. 57, p. 522, 1940, https://doi.org/10.1103/physrev.57.522.Search in Google Scholar
[25] S. H. Autler and C. H. Townes, “Stark effect in rapidly varying fields,” Phys. Rev. E, vol. 100, p. 703, 1955, https://doi.org/10.1103/physrev.100.703.Search in Google Scholar
[26] J. H. Shirley, “Solution of the Schrödinger equation with a Hamiltonian periodic in time,” Phys. Rev., vol. 138, 1965, B 979, https://doi.org/10.1103/physrev.138.b979.Search in Google Scholar
[27] G. Floquet, “Sur les équations différentielles linéaires à coefficients périodiques,” Ann. Sci. Éc. Norm. Supér, vol. 12, p. 47, 1883, https://doi.org/10.24033/asens.220.Search in Google Scholar
[28] V. A. Yakubovich and V. M. Starzhinskii, Linear Differential Equations with Periodic Coefficients, vol. 2, New York, Wiley, 1975.Search in Google Scholar
[29] G. Teschl, Ordinary Differential Equations and Dynamical SystemsGraduate Studies in Mathematics, vol. 140, Providence, American Mathematical Society, 2012.10.1090/gsm/140Search in Google Scholar
[30] I. I. Rabi, J. R. Zacharias, S. Millman, and P. Kusch, “A new method of measuring nuclear magnetic moment,” Phys. Rev., vol. 53, p. 318, 1938, https://doi.org/10.1103/physrev.53.318.Search in Google Scholar
[31] B. H. Wu and C. Timm, “Noise spectra of ac-driven quantum dots: Floquet master-equation approach,” Phys. Rev. B, vol. 81, 2010, 075309, https://doi.org/10.1103/physrevb.81.075309.Search in Google Scholar
[32] J. Q. You and F. Nori, “Atomic physics and quantum optics using superconducting circuits,” Nature, vol. 474, p. 589, 2011, https://doi.org/10.1038/nature10122.Search in Google Scholar
[33] Q. Miao and Y. Zheng, “Coherent destruction of tunneling in two-level system driven across avoided crossing via photon statistics,” Sci. Rep., vol. 6, 28959, 2016, https://doi.org/10.1038/srep28959.Search in Google Scholar
[34] P. Hannaford, D. T. Pegg, and G. W. Series, “Analytical expressions for the Bloch-Siegert shift,” J. Phys. B Atom. Mol. Phys., vol. 6, p. L222, 1973, https://doi.org/10.1088/0022-3700/6/8/009.Search in Google Scholar
[35] F. Ahmad and R. K. Bullough, “Theory of the Bloch-Siegert shift,” J. Phys. B Atom. Mol. Phys., vol. 7, p. L275, 1974, https://doi.org/10.1088/0022-3700/7/5/001.Search in Google Scholar
[36] J. M. Gomez Llorente and J. Plata, “Tunneling control in a two-level system,” Phys. Rev. A, vol. 45, 1992, R6958, https://doi.org/10.1103/physreva.45.r6958.Search in Google Scholar
[37] Y. Kayanuma, “Role of phase coherence in the transition dynamics of a periodically driven two-level system,” Phys. Rev. A, vol. 50, p. 843, 1994, https://doi.org/10.1103/physreva.50.843.Search in Google Scholar
[38] J. C. A. Barata and W. F. Wreszinski, “Strong-coupling theory of two-level atoms in periodic fields,” Phys. Rev. Lett., vol. 84, p. 2112, 2000, https://doi.org/10.1103/physrevlett.84.2112.Search in Google Scholar
[39] C. E. Creffield, “Location of crossings in the Floquet spectrum of a driven two-level system,” Phys. Rev. B, vol. 67, 2003, 165301, https://doi.org/10.1103/physrevb.67.165301.Search in Google Scholar
[40] M. Frasca, “Third-order correction to localization in a two-level driven system,” Phys. Rev. B, vol. 71, 2005, 073301, https://doi.org/10.1103/physrevb.71.073301.Search in Google Scholar
[41] Y. Wu and X. Yang, “Strong-coupling theory of periodically driven two-level systems,” Phys. Rev. Lett., vol. 98, 2007, 013601, https://doi.org/10.1103/physrevlett.98.013601.Search in Google Scholar
[42] Y. Yan, Z. Lü and H. Zheng, “Bloch-Siegert shift of the Rabi model,” Phys. Rev. A, vol. 91, 2015, 053834, https://doi.org/10.1103/physreva.91.053834.Search in Google Scholar
[43] A. Gangopadhyay, M. Dzero, and V. Galitski, “Analytically solvable driven time-dependent two-level quantum systems,” Phys. Rev. B, vol. 82, 2010, 024303, https://doi.org/10.1103/physrevb.82.024303.Search in Google Scholar
[44] E. Barnes and S. Das Sarma, “Analytically solvable driven time-dependent two-level quantum systems,” Phys. Rev. Lett., vol. 109, 2012, 060401, https://doi.org/10.1103/physrevlett.109.060401.Search in Google Scholar
[45] A. Messina and H. Nakazato, “Analytically solvable Hamiltonians for quantum two-level systems and their dynamics,” J. Phys. A: Math. Theor., vol. 47, 2014, 445302, https://doi.org/10.1088/1751-8113/47/44/445302.Search in Google Scholar
[46] T. Suzuki, H. Nakazato, R. Grimaudo, and A. Messina, “Analytic estimation of transition between instantaneous eigenstates of quantum two-level system,” Sci. Rep., vol. 8, 2018, 17433, https://doi.org/10.1038/s41598-018-35741-5.Search in Google Scholar
[47] T. Ma and S.-M. Li, Floquet System, Bloch Oscillation, and Stark Ladder, 2007, arXiv:0711.1458v2 [cond-mat.other].Search in Google Scholar
[48] Q. Xie and W. Hai, “Analytical results for a monochromatically driven two-level system,” Phys. Rev. A, vol. 82, 2010, 032117, https://doi.org/10.1103/physreva.82.032117.Search in Google Scholar
[49] P. K. Jha and Y. V. Rostovtsev, “Coherent excitation of a two-level atom driven by a far-off-resonant classical field: analytical solutions,” Phys. Rev. A, vol. 81, 2010, 033827, https://doi.org/10.1103/physreva.81.033827.Search in Google Scholar
[50] P. K. Jha and Y. V. Rostovtsev, “Analytical solutions for a two-level system driven by a class of chirped pulses,” Phys. Rev. A, vol. 82, 2010, 015801, https://doi.org/10.1103/physreva.82.015801.Search in Google Scholar
[51] A. M. Ishkhanyan and A. E. Grigoryan, “Fifteen classes of solutions of the quantum two-state problem in terms of the confluent Heun function,” Phys. Rev. A, vol. 47, 2014, 465205, https://doi.org/10.1088/1751-8113/47/46/465205.Search in Google Scholar
[52] A. M. Ishkhanyan, T. A. Shahverdyan, and T. A. Ishkhanyan, “Thirty five classes of solutions of the quantum time-dependent two-state problem in terms of the general Heun functions,” Eur. Phys. J. D, vol. 69, p. 10, 2015, https://doi.org/10.1140/epjd/e2014-50386-9.Search in Google Scholar
[53] H.-J. Schmidt, J. Schnack, and M. Holthaus, “Floquet theory of the analytical solution of a periodically driven two-level system,” Appl. Anal., 2020. Published online: 1 Jul 2019.10.1080/00036811.2019.1632439Search in Google Scholar
[54] P. London, P. Balasubramanian, B. Naydenov, L. P. McGuinness, and F. Jelezko, “Strong driving of a single spin using arbitrarily polarized fields,” Phys. Rev. A, vol. 90, 2014, 012302, https://doi.org/10.1103/physreva.90.012302.Search in Google Scholar
[55] H. Kim, Y. Song, H. Lee, and J. Ahn, “Rabi oscillations of Morris-Shore-tranformed N-state systems by elliptically polarized ultrafast laser pulses,” Phys. Rev. A, vol. 91, 2015, 053421, https://doi.org/10.1103/physreva.91.053421.Search in Google Scholar
[56] R. M. Angelo and W. F. Wreszinski, “Two-level quantum dynamics, integrability, and unitary NOT gates,” Phys. Rev. A, vol. 72, 2005, 034105, https://doi.org/10.1103/physreva.72.034105.Search in Google Scholar
[57] H.-J. Schmidt, “The Floquet theory of the two level system revisited,” Z. Naturforsch. A, vol. 73, no. 8, pp. 705–731, 2018, https://doi.org/10.1515/zna-2018-0211.Search in Google Scholar
[58] H.-J. Schmidt, “Geometry of the Rabi problem and duality of loops,” Z. Naturforsch. A, vol. 75, no. 5, pp. 381–391, 2020, https://doi.org/10.1515/zna-2019-0352.Search in Google Scholar
[59] H. P. Breuer and M. Holthaus, “A semiclassical theory of quasienergies and Floquet wave functions,” Ann. Phys., vol. 211, 1991, 2499291, https://doi.org/10.1016/0003-4916(91)90206-n.Search in Google Scholar
[60] M. V. Berry, “Quantal phase factors accompanying adiabatic changes,” Proc. Roy. Soc. Lond. A, vol. 329, pp. 45–57, 1984.10.1142/9789813221215_0006Search in Google Scholar
[61] Y. Aharonov and J. Anandan, “Phase change during a cyclic quantum evolution,” Phys. Rev. Lett., vol. 58, p. 1593, 1987, https://doi.org/10.1103/physrevlett.58.1593.Search in Google Scholar
[62] I. Menda, N. Burič, D. B. Popovič, S. Prvanovič, and M. Radonjič, “Geometric phase for analytically solvable driven time-dependent two-level quantum systems,” Acta Phys. Pol. A, vol. 126, p. 670, 2014.10.12693/APhysPolA.126.670Search in Google Scholar
[63] M. Naghiloo, J. J. Alonso, A. Romito, E. Lutz, and K. W. Murch, “Information gain and loss for a quantum Maxwell’s demon,” Phys. Rev. Lett., vol. 121, 2018, 030604, https://doi.org/10.1103/physrevlett.121.030604.Search in Google Scholar
[64] P. Talkner, E. Lutz, and P. Hänggi, “Fluctuation theorems: work is not an observable,” Phys. Rev. E, vol. 75, 2007, 050102, https://doi.org/10.1103/physreve.75.050102.Search in Google Scholar
© 2020 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- Atomic, Molecular & Chemical Physics
- Control of the half-cycle harmonic emission process for generating the intense and ultrashort single attosecond pulses (SAPs)
- Dynamical Systems & Nonlinear Phenomena
- Influence of varying magnetic field on nonlinear wave excitations in collisional quantum plasmas
- Nonlinear rarefactive isothermal ion acoustic waves in magnetized ultrarelativistic degenerate plasmas
- Rapid Communication
- Comment on “On biological signaling” by G. Nimtz and H. Aichmann, Z. Naturforsch. 75a: 507–509, 2020
- Quantum Theory
- The Rabi problem with elliptical polarization
- A basic definition of spin in the new matrix dynamics
- Solid State Physics & Materials Science
- Pressure induced nodal line semimetal in YH3
- Catalytic removal of methylene blue with different stoichiometric ratios of ZnCuS nanoparticles
- Surface-state energies and wave functions in layered organic conductors
Articles in the same Issue
- Frontmatter
- Atomic, Molecular & Chemical Physics
- Control of the half-cycle harmonic emission process for generating the intense and ultrashort single attosecond pulses (SAPs)
- Dynamical Systems & Nonlinear Phenomena
- Influence of varying magnetic field on nonlinear wave excitations in collisional quantum plasmas
- Nonlinear rarefactive isothermal ion acoustic waves in magnetized ultrarelativistic degenerate plasmas
- Rapid Communication
- Comment on “On biological signaling” by G. Nimtz and H. Aichmann, Z. Naturforsch. 75a: 507–509, 2020
- Quantum Theory
- The Rabi problem with elliptical polarization
- A basic definition of spin in the new matrix dynamics
- Solid State Physics & Materials Science
- Pressure induced nodal line semimetal in YH3
- Catalytic removal of methylene blue with different stoichiometric ratios of ZnCuS nanoparticles
- Surface-state energies and wave functions in layered organic conductors