Analogue of Hamilton-Jacobi theory for the time-evolution operator

Michael Vogl,1, 2 Pontus Laurell,1, 3 Aaron D. Barr,1 and Gregory A. Fiete1, 2, 4 1Department of Physics, The University of Texas at Austin, Austin, TX 78712, USA 2Department of Physics, Northeastern University, Boston, MA 02115, USA 3Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, USA 4Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA (Dated: July 16, 2019) In this paper we develop an analogue of Hamilton-Jacobi theory for the time-evolution operator of a quantum many-particle system. The theory offers a useful approach to develop approximations to the time-evolution operator, and also provides a unified framework and starting point for many well-known approximations to the time-evolution operator. In the important special case of peri- odically driven systems at stroboscopic times, we find relatively simple equations for the coupling constants of the Floquet Hamiltonian, where a straightforward truncation of the couplings leads to a powerful class of approximations. Using our theory, we construct a flow chart that illustrates the connection between various common approximations, which also highlights some missing con- nections and associated approximation schemes. These missing connections turn out to imply an analytically accessible approximation that is the “inverse” of a rotating frame approximation and thus has a range of validity complementary to it. We numerically test the various methods on the one-dimensional Ising model to confirm the ranges of validity that one would expect from the approximations used. The theory provides a map of the relations between the growing number of approximations for the time-evolution operator. We describe these relations in a table showing the limitations and advantages of many common approximations, as well as the new approximations introduced in this paper.

I. INTRODUCTION operator of finite-size systems. Some approximations we can discuss in a unified manner with the framework intro- duced in this paper include the Dyson-Neumann series, Our understanding of the dynamics of quantum many- [44–46] the Magnus expansion,[47–51] Fer’s series,[50–52] particle systems and associated non-equilibrium phenom- Wilcox’ series,[50, 51, 53] the rotating frame approxima- ena has seen rapid growth in recent years [1, 2], which tion, [13] and flow equation methods.[54] has resulted from advances in theory—especially Flo- quet systems [3–15]—and experiment, particularly in the The focus of this paper is on the flow equations for preparation of and characterization of non-equilibrium couplings (which we will introduce shortly) in a Hamil- states.[2, 16–22] As the field of non-equilibrium quantum tonian and their relation to the various approximations systems expands, an increasing amount of effort is being mentioned above. We note that important results were devoted to the study of such systems, driven in part by obtained in prior work by flow equation methods in the the wealth of novel phenomenology that the time domain equilibrium case, [55, 56] and the non-equilibrium case permits. For cold atoms trapped in optical lattices, time- [57] making use of the construct of a Sambe space.[58] In dependent driving has enabled [2, 23] coherent control this work, we find an analogue to Hamilton-Jacobi the- of tunneling [24], induction of phase transitions [25, 26], ory for the time-evolution operator that allows one to generation of effective magnetic fields [27], and the mea- make clear connections between various approximation surement of nontrivial topological invariants [28]. Note- schemes. Our formulation takes shape in the form of flow worthy examples from solid state systems include pho- equations for the couplings in the Hamiltonian.[55, 56] toinduced superconductivity, [29, 30] and hidden or oth- Our discussion culminates in a diagram, Fig.1, that erwise inaccessible orders. [31–33] Even more strikingly, interrelates the different approximations and highlights the time domain also allows entirely novel phases, such spots which symmetry suggests can be filled by consid- as time crystals,[34, 35] and non-equilibrium topological ering another limiting flow equation. In this paper, we phases. [9, 36–42] develop this limiting flow equation and find that the ap- A number of numerical and analytical tools have proximations perform as one would have expected from arXiv:1902.07237v3 [cond-mat.stat-mech] 15 Jul 2019 been developed to understand the principal quantity of the order of approximations seen in the diagram in Fig.1. interest—the time evolution operator—which mathemat- We display a table, TableI, that makes clear to the reader ically is a time-ordered exponential. It is not practical to the range of validity for each approximation, along with summarize all known methods to calculate this quantity, its advantages and shortcomings. In particular, we find so we will set our focus on analytically accessible approx- that the flow equations obtained in Ref.[54] are useful imations that can be used for arbitrary forms of time- when truncated like Wegner’s, [55, 56] and the approxi- dependence. This restriction will therefore exclude the mation we develop in this paper (by completing our flow vast literature on numerical methods, and a recently de- chart) fills in a gap for an analytically accessible approx- veloped exact method [43] to calculate the time-evolution imation in the intermediate time regime (or equivalently 2 for a time-periodic system in the intermediate frequency If one repeatedly reinserts the left side into the right side regime). Our results provide a comprehensive picture of of the equation one obtains a series in powers of H(t). the approximations to the time-evolution operator that Truncated at second order, this series is given as, can be used for time-periodic quantum many-particle sys- Z t Z t Z t1 0 0 3 tems. The approximation we develop here is particularly U(t) = 1H −i dt H(t )− dt1 dt2H(t1)H(t2)+O(H ). useful when kV k  kH0k ∼ ω. That is, when the system 0 0 0 is subject to a drive V that is weak compared to the static (4) Neglecting higher order terms in H(t) is valid if H(t) is part of the Hamiltonian H0 but when the static part H0 is not negligible when compared to the drive frequency small compared to ∼ 1/t. In other words, such an expan- ω. This regime is relevant to e.g. cavity-QED appli- sion is restricted to sufficiently short times. In addition, cations [59, 60] (where strongly interacting photons are truncating the series destroys its unitarity, which is a se- often subject to weak time-periodic drives), and weakly rious drawback. The loss of unitarity occurs already at driven cold atom quantum ratchets.[61] first order in H(t). Our paper is organized as follows. In Sec.II, we give If one is interested in the evolution of eigenstates of a short introduction to each of the common approxi- a constant Hamiltonian H0, it is advantageous to split −iH0t mations mentioned above. In Sec.III, we establish re- U(t) = UI (t)e because in this case the second factor −iH t iφ lations between the different approximations. In Sec.IV only leads to a phase e 0 |ψi = e |ψi. For expecta- we present the new approximation and a diagram of re- tion values of an operator Oˆ the phase factor disappears ˆ † ˆ † ˆ lations in Fig.1 that were suggested by symmetry. In hOi(t) = hψ|U (t)OU(t)|ψi = hψ|UI (t)OUI (t)|ψi and UI Sec.V we compare the different approximations for an encodes all relevant information. It now fulfills Ising model. In section VI we calculate the l2 distance Z t 0 0 2 between the exact and the approximate time-evolution UI (t) = 1H − i dt VI (t ) + O(VI ), (5) operators. Finally, in Sec.VII we present our conclusions. 0 iH0t −iH0t A few technical details are relegated to the appendices. where VI = e (H(t) − H0)e . Such a procedure is called the interaction picture (hence the “I” subscript). The expansion for UI (t) is often used in setting up Feyn- II. SUMMARY OF COMMON man diagrams for scattering problems because for small APPROXIMATIONS TO U(t) perturbations V (t) the procedure works well enough to approximate the S-, S = UI (t = ∞). The time-evolution operator U(t) fulfills i∂ U(t) = H(t)U(t); U(0) = 1 , (1) t H B. Magnus expansion where 1H is the identity on the Hilbert space of the Hamiltonian, H. We have set Planck’s constant, ~ = 1. The broken unitarity in Dyson’s approach is a seri- Eq.(1) can be solved by a simple matrix exponential if ous drawback because spurious terms may appear in [H(t1),H(t2)] = 0 for all t1 and t2. However, when calculations—a problem Magnus[47–51] solved. His way [H(t1),H(t2)] 6= 0 Eq.(1)becomes more complicated and of dealing with this issue was by making the ansatz one has to concatenate matrix exponentials for infinites- U(t) = eΩ(t) for the time-evolution operator and search- imal time steps dt, ing for anti-Hermitian Ω instead of U(t). Inserting this U(t) = e−idtH(t−dt)e−idtH(t−2dt) ··· e−idtH(0). (2) ansatz into Eq.(1) and using the general expression for the derivative of the exponential map he found that, Such a procedure, which sometimes is called Trot- terization, reproduces the formal solution: U(t) = dΩ(t) adΩ −i R t dt0H(t0) = i ad H(t), (6) T e 0 in the limit dt → 0, where T is the time- dt e Ω − 1 order operator and the expression is called a time-ordered where the shorthand adΩ = [Ω,.] was used. exponential. However, in general this is not an analyt- The solution can be found by first solving the equation ically tractable procedure and therefore approximations for adΩ = 0, i.e. to lowest order and then reinserting the are needed. In the remainder of this section we will sum- result to generate higher orders. One finds that, marize a few of the most common approximations. 4 U(t) = eΩ1(t)+Ω2(t)+Ω3(t)+O(H ), Z t A. Dyson-Neumann series 0 0 Ω1(t) = −i dt H(t ), 0 An important approximation to the time ordered ex- 1 Z t Z t1 Ω2(t) = − dt1 dt2[H(t1),H(t2)], ponential is due to Dyson.[44–46] One integrates Eq.(1) 2 0 0 to find, i Z t Z t1 Z t2   Z t Ω3 = dt1 dt2 dt3 [H(t1), [H(t2),H(t3)]] + (t1 ↔ t3) , 0 0 0 6 0 0 0 U(t) = 1H − i dt H(t )U(t ). (3) 0 (7) 3 where (t1 ↔ t3) = [H(t3), [H(t2),H(t1)]] is used as a in other cases. One can, however, make a good argument shorthand. We denote higher orders by Ωn. as to why this statement might be generally true. Similar to the case of the Dyson series, this approxima- The behaviour may be rooted in the fact that the tion only works for sufficiently short times or sufficiently Wilcox approximation assumes that higher order terms small Hamiltonians. However, it is an improvement in eWn appear to the right of lower order terms eWn−1 rather that it is unitary to all orders and therefore its mathe- than the left. There is no a priori reason for either choice. matical structure is more sound. We note that at the This kind of asymmetry or ambiguity does not exist in lowest orders it agrees with the expansion discussed in the Magnus case, which may be the reason it performs Ref.[13]. better. A similar type of asymmetry that is due to an It is worth noting that the Magnus Expansion does not ambiguity of ordering also exists for the Zassenhaus for- 2 t(X+Y ) tX tY − t [X,Y ] converge in all cases, which means that an optimal cut-off mula e = e e e 2 ··· , which is a dual of order exists in these cases. Important work to understand the BCH formula. Observations of this asymmetry have how this affects the time-scales accessible by a description inspired recent attempts to symmetrize the Zassenhaus using a Magnus expansion has been published in Ref.[62]. formula[63], which was able to yield some improvements in numerical accuracy. One may only wonder if a sim- ilar procedure could be used to improve on the Wilcox C. Wilcox expansion expansion. It is also worth noting that the procedure in Ref.[63] Matrix exponentials of complicated operators are dif- introduces its own ambiguities. Rather than having to ficult to calculate and approximate. Wilcox (inspired ask if one should order different terms in the perturba- by Fer’s work, which we will discuss next) [50, 51, 53] tion series from left to right, or right to left, one has to split the Magnus expansion into separate exponentials: ask if one should order them from inside to outside or Ω1(t)+...+Ωn(t) W1(t) Wn(t) m e → e ...e , where Wm = O(H ). from outside to inside. Therefore, the usefulness of the n Terms of order eO(H ) are neglected if the product is Wilcox approximation may be restricted to niche uses W truncated after eWn−1 . Such an expansion would be ad- where separate operator exponentials e i are easier to P W P Ω Ωi vantageous if e i is easier to calculate than e i i . compute than e i . To generate the terms up to Wn one uses the Baker-Campbell-Hausdorf (BCH) formula eAeB = A+B+ 1 [A,B]+O(A2B,AB2) e 2 . One also introduces a dummy D. Fer’s series parameter δ that keeps track of different orders of H. δW (t) δnW (t) Pn δmO (W ) One then finds that e 1 ...e n ≈ e m=1 m i , Fer [50–52] approached the challenge of finding a time- where Om is a of the operators Wi that includes ordered exponential quite differently. His idea was to first all terms of order δm that are generated by the BCH ignore the time-ordering aspect and as a first approxima- P δmΩ expansion. A comparison with e m m gives the set tion take, of equations Ωm = Om(Wi). These can be solved for −i R t dt0H(t0) the Wi. A more detailed description of this procedure is U(t) ≈ U1(t) = e 0 , (10) given in Ref. [50]. One finds to the first two orders, for a general time-dependent Hamiltonian H(t). Unlike Magnus he did not search for corrections to the 3 U(t) = eW1(t)eW2(t)eO(H ), exponent but instead took the time-evolution operator to Z t factorize, U(t) = U1(t)U2(t), and found an equation for 0 0 U (t), W1(t) = −i dt H(t ), (8) 2 0 t t † 1 Z Z 1 i∂tU2(t) = H2(t)U2(t); H2(t) = U1(t) (H(t)−i∂t)U1(t). W2(t) = − dt1 dt2[H(t1),H(t2)]. (11) 2 0 0 One may now for U2 again ignore the time-ordering These terms and higher order terms are related to the aspect and repeat the same procedure. One then finds Magnus expansion. This relation up to order H3 is given the recursive scheme, by 2n U(t) = U1...Un + O(H ), W1(t) = Ω1(t); W2 = Ω2(t); Z t −Fj (t) 0 0 1 (9) Uj(t) = e ; Fj(t) = i dt Hj(t ), W3(t) = Ω3(t) − [Ω1(t), Ω2(t)]. 0 2 † Hj(t) = Uj−1(t)(Hj−1(t) − i∂t)Uj−1(t); H1(t) = H(t). We will see later in an explicit example illustrating (12) that this approximation is not as good as the Magnus The advantage of this procedure over the Wilcox ap- approximation. While we cannot be certain that this is proach is that at each order j infinitely many orders of generally the case, we have observed the same property H(t) are added to the exponents Fj. Including infinitely 4 many orders of H(t) should be expected to result in a to Eq.(1), which achieves the goal of removing V (t) if the more reliable approximation to the time-evolution oper- time interval [0,T ] is comparatively small. A particularly ator. While infinitely many terms are added, the method convenient property of this transformation is that it re- is still controlled. It is found [50] that the terms that are duces to the identity operator at t = T by construction: neglected when dropping the n-th term in the product S1,T (T ) = 1H . n are of order O(H2 ). For these reasons (in our exam- Because of this property the time-evolution operator U ple later) the Fer expansion is found to be more reliable at times T has no contribution from S1,T (t). Therefore, than the Magnus expansion if we break the series off at to learn about times T it is enough to calculate the time small orders, which is the practical thing to do because evolution operator UT in the rotating frame, high orders become complicated in both cases. There- fore, while the convergence radius of the Fer approxima- i∂tUT (t)= H1,T (t)UT (t), (14) tion is smaller than for the Magnus case [47] at small † H1,T (t)= S (H(t) − i∂t)S1,T . (15) orders, even when it does not converge, it is often found 1,T to be more reliable—an effect often-times humorously(!) A solution at times T that ignores the time-ordering summarized by Carrier’s rule: “Divergent series converge aspects of the time evolution operator, faster than convergent series because they don’t have to −i R T dtH (t) converge.” [64] U(T ) ≈ e 0 1,T (16) The disadvantage of the method over the Magnus case is that it is often extremely difficult to calculate the Hj, in in many cases now turns out to be an improvement over many cases even H2. Furthermore, the method—like the the same done for Eq.(1)—particularly this is true if other approaches discussed to this point—is restricted to VT (t) is larger than H¯T [13, 54]. relatively short times. As with Fer’s method, one may iterate this procedure. One may split H1,T (t) in the same way we split H(t). Following this logic, one finds that the time evolution E. Rotating frame approximation operator can be successively approximated. An iterative procedure is given by,

In the rotating frame approximation, one finds an −i R T dtH (t) U(t) ≈ e 0 n,T , approximate time-evolution operator without having to T =t worry about the time-ordering aspect. This is accom- † Hn,T (t) = Sn,T (t)(Hn−1,T (t) − i∂t)Sn,T (t), plished by removing the time-dependence up to an ar- R −i dtVn,T (t) bitrarily chosen time T by a unitary transformation. In Sn,T = e , general this is a difficult task. However, it turns out that 1 Z T it is possible to do this to a good approximation if one Vn,T (t) = Hn,T (t) − dtHn,T (t); H0,T (t) = H(t). ¯ T 0 splits H(t) = HT + VT (t) into a part that is constant (17) ¯ 1 R T on the interval [0,T ] given by HT = T 0 dtH(t), and a One finds [13, 54] that this approximation offers a ¯ part VT (t) = H(t) − HT that averages to zero over the significant improvement over the Magnus approximation same interval. Here, T is an arbitrarily chosen time at when V (t) is large. However, it is sometimes more cum- which the time evolution operator will be evaluated. The bersome to implement. Also, it is important to stress time dependence can be arbitrary and there is no need that, since T could be chosen arbitrarily, as a final step for a time-periodic drive. one has to set T = t in U(t). However, it is important to note that this type of split- ting occurs naturally in periodically driven systems at stroboscopic times. For this reason this type of approx- III. RELATIONS BETWEEN THE METHODS imation is commonly applied in a Floquet setting. One should also note that statements about the convergence While the approximations discussed in Sec.II may seem of stroboscopic time evolution operators in the literature unrelated to one another at a first glance, there is an apply for a non-periodic case if we take T as an “artifi- overarching reformulation of Eq.(1) that connects them cial” stroboscopic time and ensure to exclude all state- all. ments that consider multiple periods. This point is il- Let us recall the basic idea of Hamilton-Jacobi theory lustrated by realizing that, to calculate U(T ), we do not [65]. In Hamilton-Jacobi theory one arrives at a refor- need to know if H(2T ) = H(T )—this does not have to mulation of classical mechanics by searching for a gen- be true, rather it is enough to know only about times erating function of canonical transformations that make 0 < t < T . the (generally time-dependent) Hamiltonian equal to a To remove the time dependent part of the Hamiltonian constant—thereby removing the focus from Hamilton’s one may apply the unitary transformation, equations. The full information of the dynamics is then absorbed into the generator. Without loss of generality, R t 0 0 −i dt VT (t ) S1,T (t) = e 0 , (13) the constant may be taken to be zero. Specifically, one 5 may consider a generator S(q, P, t) of canonical transfor- which is the quantum analogue of Eq.(18) since both mations from coordinates (q, p) to coordinates (Q, P). equations determine a transformed Hamiltonian. Un- One finds that the Hamiltonian K(Q, P, t) in terms of like the classical version, we introduced an additional the new coordinates is given as parameter s for calculational convenience. The reason is that transformations in the quantum case are operators ∂S K(Q, P, t) = H(q, p, t) + . (18) rather than phase space functions, and thus harder to ∂t determine. In principle, it could be possible to find the The variables then also fulfill the conditions, unitary transformation removing H(t) in a single step, which would obviate the need to introduce s. ∂S ∂S The appropriate boundary conditions are set by p = ; Q = . (19) ∂q ∂P putting H(t, 0) as the original untransformed Hamilto- nian. We may also keep track of the time-evolution op- One may then construct an S such that K = 0, and erator in the original frame. For the first infinitesimal therefore P˙ = Q˙ = 0, by solving step it is, ∂S ∂S H(q, , t) + = 0, (20) U(t) = S(t, δs)Uδs(t), (25) ∂q ∂t and the more general case is found by repeating this after ∂S where we made use of p = ∂q . each infinitesimal transformation. One should stress that the key idea of the Hamilton- Up to this point, the treatment coincides with the use Jacobi formalism is to construct a coordinate transforma- of time-dependent generators in Ref.[66]. We now, how- tion that gets rid of the Hamiltonian and thereby makes ever, choose Σ very different from the Wegner generator the equations of motion trivial. We will do the same here. (which is designed to block diagonalize H). We choose it We will try to find a unitary transformation that gets rid such that it reduces the Hamiltonian H(t) → (1−δs)H(t) of the Hamiltonian so that the equation for the time- by some infinitesimal value δs, evolution operator becomes trivial. This is the central Z t idea of this paper. 0 0 We take Eq.(1) as a starting point and introduce a uni- Σ(t) = −i dt H(s, t ). (26) 0 tary transformation, S = eδsΣ(t), generated by an as yet undetermined quantity Σ(t) that will be chosen to reduce Notice that this generator also leaves a residual term the Hamiltonian H(t) as H(t) → (1 − δs)H(t). Hereby R t δs 0 dt1 [H(t1),H(t)] in Eq.(22). We will discuss it later. δs is infinitesimal and ensures that the exponential can With our specific choice of generator Σ we find that be safely expanded to lowest order. Eq.(24) becomes, We may split the time evolution operator as U0 = † † t S Uδs = [1 − δsΣ(t)]Uδs and act with S(t) = 1 − δsΣ(t) dH(s, t) Z from the left on the Schr¨odingerequation. The time evo- = −H(s, t) + i dt1 [H(s, t1),H(s, t)] , (27) ds 0 lution operator in the new frame Uδs now fulfills the mod- ified equation, which is the equivalent of Eq.(20) in the Hamilton-Jacobi formalism, just for the time evolution operator. The i∂tUδs = (H(t) − iδs∂tΣ(t) − δs [Σ(t),H(t)]) Uδs. (21) analogy becomes more clear if we recognize that both equations determine a transformed Hamiltonian that is One may read off a new Hamiltonian, zero. Let us discuss this in slightly more detail. One can H(t, δs) = H(t) − iδs∂tΣ(t) − δs [Σ(t),H(t)] , (22) directly check that the fixed point H(s, t) = 0, which we where we introduced a second parameter slot for a param- want to reach, is stable. To see this one realizes that near dH(s,t) eter s in addition to the time dependence, which labels the fixed point Eq.(27) reduces to ds = −H(s, t). the behavior of the Hamiltonian along a unitary flow. If dH(s,t) This means that near the fixed point [ ds ,H(s, t)] = we replace H(t) → H(t, s), H(t, δs) → H(t, s + δs) and 0. Thus the quantities display a behaviour like a scalar. Σ(t) → Σ(t, s) we may keep track of how the Hamilto- Therefore, one may apply an ordinary fixed point anal- nian changes under a chain of dynamically determined ysis, according to which the fixed point H(s, t) = 0 is infinitesimal unitary transformations, stable. Furthermore, this fixed point is the only fixed point. Let us briefly see why. H(t, s + δs) = H(t, s) − iδs∂ Σ(t, s) − δs [Σ(t, s),H(t, s)] . t The equation that determines the fixed point is (23) H(s, t) = i R t dt [H(s, t ),H(s, t)]. One may also real- By a Taylor expansion around δs = 0 we see that the 0 1 1 Hamiltonian fulfills the differential equation, ize that Tr(H(s, t)[H(s, t1),H(s, t)]) is zero by the cyclic property of trace, and therefore the Frobenius inner prod- dH(t, s) uct hH(s, t), [H(s, t1),H(s, t)]i = 0. Both sides of the = −i∂ Σ(t, s) − [Σ(t, s),H(t, s)] , (24) ds t equation, therefore, are perpendicular in an operator 6 product sense. This means the equation can only be ful- seem arbitrarily chosen. The heuristic reason, however, filled if both sides are zero. Therefore, the only fixed is that setting s = 0 in the generator already assumes a point is H = 0. Thus, it can be expected that the equa- small value of s. This is only justified if s < 1, so for tions will flow toward H(s, t) = 0 as needed for the anal- consistency we need to set s < 1. The reason we let it ogy to the Hamilton-Jacobi theory to hold. run up to this maximum value is that we want to be able One should note that the unitary transformation that to get as close to a fixed point as possible. For a more gets rid of the Hamiltonian was obtained by multiply- detailed discussion of such an approximation we point to ing infinitely many infinitesimal unitary transformations Ref.[54]. Another way to look at this approximation is R t 0 0 to recognize that it just performs the unitary transform sδ(t, s) = (1−iδs 0 dt H(s, t )). In other words, one may 1 −i R t dtH(t) write, SF = e 0 . This is the same thing that happens in the lowest order of Fer’s approximation. S(t, s + δs) = sδ(t, s)S(t, s) Since in a unitary transformation we do not lose any Z t (28) 0 0 information one can make repeated use of Eq.(31). Con- = S(t, s) − iδs dt H(s, t )S(t, s). catenating these transformations allows us to reconstruct 0 the expansion due to Fer, Eq.(12). In fact, this reformu- One finds via a Taylor expansion that, lation is more powerful than the standard approach due to Fer since his method usually cannot be implemented Z t 0 0 analytically. The necessary unitary transformations are i∂sS(t, s) = dt H(s, t )S(t, s). (29) 0 often hard to calculate. The advantage of our method is that we may make use of the non-perturbative nature The time-evolution operator in the frame after rotation of Fer’s approach but avoid some of its difficulties if we by S is now trivially given as Us=∞ = 1, because the make another non-perturbative approximation, which is Hamiltonian is zero. Therefore, the time evolution oper- taking a truncated ansatz for H(s, t). This allows us to ator in the original frame is just U(t) = S(t), by Eq.(25). do the necessary unitary transform approximately while How do we make practical use of the operator val- keeping infinite orders from the couplings in H(t). The ued Eq. (24)? In it, H(s, t) is a linear operator validity of such an approach will be shown later on in an and therefore may be written as a linear combina- explicit example. tion of operators with coefficients ci(s, t), H(s, t) = The lowest order Wilcox approximation, U(t) ≈ P ˆ W1 W 2 i ci(s, t)Oi. This mathemetical structure in turn e e , also follows naturally from Eq.(31). If one solves R t Eq.(31) while neglecting the H2 term one finds, also implies that −H(s, t) + i 0 dt1 [H(s, t1),H(s, t)] = P 0 ˆ − i gi(t, [cj(s, t )])Oi, where gi has a functional depen- 0 dence on the cj(s, t ) because V (s, t) itself depends on H(s, t) ≈ (1 − s)H(t). (32) the cj(s, t) and it appears under an . One may therefore write Eq. (27) as, Reinserting this result in Eq.(31), one finds the solu- tion dci(s, t) X = − g (t, [c (s, t0)]). (30) ds i j i Z t i H(1, t) = − dt1[H(t1),H(t2)]. (33) 2 0 At a first glance one may wonder whether Eq.(27) is use- ful because it is a complicated functional equation mak- Therefore, to order H the time evolution operator is ing it difficult to obtain H(s, t). Moreover, it may appear given by the unitary transformation that we tried to im- R 0 0 that not only did we not get rid of the problem of having plement, U(t) = e−i dt H(t ), which means that we re- to find a time ordered exponential Eq.(29), but we made produced the exponent W1. Removing H(1, t) by the the issue worse by adding additional complications. same procedure, we reproduce W2. Therefore, finding However, this exercise was a worthwhile time invest- the lowest order Wilcox approximation from Eq.(27) was ment because it allows a very simple way of identifying just as easy. One should remark that we may also repro- approximations. If we assume that we can get rid of H(t) duce the Dyson-Neumann approximation, which is found swift enough that s ≈ 0 we may set H(s, t) ≈ H(0, t) = by a Taylor expansion. H(t) in the generator Σ in Eq.(26). Details on this kind So how do we reproduce the remaining approximations of approximation are given in Ref.[54]. This means that - rotating frame and Magnus approximation by going we are left with back to Eq.(27)? Let us, like in the rotating frame ap- ¯ dH(s, t) Z t proximation in Sec.II E, split H(s, t) = HT (s) + VT (s, t) = −H(t) + i dt [H(t ),H(s, t)] , (31) 1 R T 1 1 into a constant part H¯T (s) = dtH(s, t) and a time- ds 0 T 0 dependent part VT (s, t) = H(s, t) − H¯T (s) that averages where the complication of functional dependences on t is to zero on the interval [0,T ]. Any arbitrary choice of T gone. Now, if we let s run from 0 to 1 we get rid of H(t) is possible; in a last step one will have to set T = t. If we to lowest order. The value of s = 1 at this point may assume that V (t) is dominant in the generator Eq.(26) 7 then Eq.(27) reduces to We would like to stress the added convenience this re- sult is expected to provide for numerical studies with the Z t dH(s, t) flow equation approach. One can now solve a set of dif- = −VT (s, t) + i dt1 [VT (s, t1),H(s, t)] , (34) ds 0 ferential equations for couplings of the generator of stro- boscopic time evolutions, i.e. the Floquet Hamiltonian. where we stress that s ∈ [0, ∞). Now let us show the usefulness this approach provides One should note that Eq.(34) is not an approxima- when we try to find approximation schemes. We can tion but rather a unitary transformation that achieves make the same approximation to Eq.(34) as we previously a different goal than the one previously considered. For did to Eq.(27). Namely, we assume that we can get rid of the specific case of a periodically driven system we dis- V (t) swiftly enough that we may set V (s, t) ≈ V (0, t) = cussed it great detail in Ref.[54]. However, let us quickly V (t) in the generator. In this case Eq.(34) simplifies to summarize. This equation does not remove H¯ (s) but t only VT (s, t) will be removed. The generator Σ = dH(s, t) Z R T = −VT (t) + i dt1 [VT (t1),H(s, t)] . (37) −i 0 dtVT (s, t) has an advantage over the original gen- ds 0 erator because it vanishes at times T . This makes the equation a bit more useful than Eq.(27) because the time Again, to lowest order, V (t) is removed if we let s run from zero to one. This implements the unitary trans- evolution operator U(t) at times t = T now coincides R −i dtVT (t) with the time evolution operator Us=∞(T ) in the rotated formation ST = e , which vanishes at times T . frame. It simply becomes, Therefore, the time evolution operator in this approxi- mation at times t is given as,

−iH(∞,T )T U(t) = Us=∞(T )|T =t = e . (35) R 0 0 T =t U(t)≈ e−i dt H1(t ); (38) † Since T could be chosen arbitrarily this poses no re- H1(t)= ST (t)(H(t) − i∂t)ST (t)|T =t, (39) striction and we were able to set T = t in a last step. With Eq.(34) one may find the generator H(∞, t) of the which is the same as the rotating frame approximation, time evolution. and where we set T = t because T could be chosen ar- Let us pause for a moment and realize that this choice bitrarily to match t. Repeatedly applying the flow equa- of unitary transformation completely removed the need tions produces the full expansion Eq.(17). to calculate a time-ordered exponential. One now just One should stress again that the advantage of the flow has to calculate a mundane matrix exponential. However, equation approach is that one may make a truncated interpreting Eq.(34) in the same way as an equation of ansatz for H(s, t) and therefore calculate an approximate the form Eq.(30) did for the couplings ci, we have traded rotating frame approximation in cases where an exact the complications of a time-ordered exponential for flows matrix exponential may not be calculated. That is, we of couplings with a complicated functional dependence. may take advantage of the non-perturbative nature of the Nevertheless, even in the functional form, Eq.(30), is rotating frame transformation in more cases. In Ref.[54] useful when describing many-body driven systems be- such a truncation for one model was discussed and one cause one can make an ansatz for H(s, t) and one only has may see the advantage this approach still has over a Mag- to solve a finite set of equations numerically for the cou- nus expansion. We will see this explicitly for an example plings in H(s, t). Semi-analytic calculations with such problem later in this work. an expression for the time-evolution operator are then Now let us see how the lowest orders of the Magnus possible because a matrix exponential is much more ac- approximation can be obtained from this approach. As cessible than a time ordered exponential. The method is in the Wilcox approximation case, we solve Eq.(37) while particularly useful when dealing with periodically driven dropping the commutator term to find, systems because H(s = ∞) is then the Floquet Hamilto- ¯ nian. H(s, t) ≈ H + (1 − s)V (t). (40) We should also note that Eq.(34) simplifies signif- icantly for a specific class of Floquet systems. The If we reinsert this into Eq.(37) and perform an integra- functional dependence on couplings in Eq.(30) vanishes tion by parts we find that completely for such a special case. Indeed if H(t) is 1 Z T 1 Z T Z t monochromatically driven, i.e. has the form H(t) = H(1,T ) = dtH(t)+ dt dt1[H(t),H(t1)]. H + eiωtH + e−iωtH with H = H† , then for strobo- T 0 2T 0 0 0 + − + − (41) scopic times T = 2π/ω (where ω is the drive frequency) The matrix exponential U(T ) ≈ e−iH(1,T )T is then the there are no functional dependences. Rather, by a com- second order result of the Magnus expansion. Lastly, the parison of coefficients one finds the set of equations Dyson series to the low orders can be found, similar to dH0 2 1 the Wilcox approximation, by expanding U(T ) to order = [H+,H−] + [H0,H+ − H−], H2. ds ω ω (36) dH 1 We are now in a position to draw a diagram in Fig.1 ± = −H ± [H ,H − H ]. ds ± ω ± 0 ∓ that relates the different approximations we discussed. 8

One may see by the symmetry of the diagram in Fig.1 Going back to solve Eq.(45) exactly, we recognize that that there are some approximations still missing, which one may concatenate the reverse rotating frame transfor- we marked in a dashed box. They will be the topic of mations, which we will denote by Si. The time evolution the next section. operator at times T can then be approximated by a prod- uct of these Si,

2 IV. LARGE CONSTANT PART IN THE U(t) ≈ S0(T )S1(T )...Sn(T )|T =t + O(V ); H0,T (t) = H(t), HAMILTONIAN AND THE “REVERSE” 1 Z T −iH¯j,T t ROTATING FRAME APPROXIMATION Sj,T (t) = e ; H¯j,T = dtHj(t), T 0 † ¯ We would like to find a different flow equation to com- Hj(t) = Sj−1(t)(Hj−1(t) − Hj−1,T )Sj−1(t), plete the rest of the diagram in the left-hand side of Fig.1. (46) This time we let H¯ dominate in the generator, so that where again we were able to evaluate U at T = t since T Σ = −iH¯T (s)t and we find flow equations, can be chosen arbitrarily. The approximation in Eq.(46) is expected to work well in the limit of H¯T  VT . dH(s, t) = −H¯ (s) + it[H¯ (s),H(s, t)]. (42) As in the previous cases let us make yet another ap- ds T T proximation that follows the same structure as before. Namely, we first solve Eq.(45) for the case that we can Here, Eq.(42) is not to be interpreted as an approx- neglect the commutator and find, imation, but rather it is constructed such that it re- moves all constant terms in the Hamiltonian. This is, H(s, t) ≈ (1 − s)H¯T + VT (t). (47) H(∞, t) = VT (∞, t). More precisely, the time evolu- tion operator in the transformed frame (at s = ∞) is Reinserting this result one finds that, R T {e dtV (∞,T )} = 1 +O(V 2) because R T dtV (s, t) = 0. H 0 T t Therefore, to order O(V 2) we can neglect the contribu- H(1, t) ≈ V (t) + i [H¯ ,V (t)]. (48) T 2 T T tion of the time evolution operator in the transformed frame. After setting T = t (since T was arbitrarily cho- Therefore, we may approximate by doing a time average sen) the time evolution is given as i Z T 2 H¯1 ≈ dt t[H¯T ,VT (t)] U(t) = S(∞,T )|T =t + O(V ), (43) 2T 0 (49) Z T Z t1 with i = − dt1 dt2[H¯T ,VT (t2) − VT (t1)], 2T 0 0 ∂sS(s, T ) = −iT H¯T (s). (44) R T where 0 dtVT (t) = 0 and the Cauchy formula for re- Unlike the previous methods, Eq.(27) and Eq.(34), no peated integration was used. One may use this result integral appears on the right hand side of Eq.(42), which in Eq.(46) to get an approximation to the time evolu- means that the coupling constants fulfill simple differ- tion operator. The result Eq.(46) reproduces the Wilcox ential equations (and not complicated functional equa- series in the limit of small VT (t) to order VT (t). tions). However, we did not eliminate the need to calcu- Now that all approximations are in place, we have com- late a time ordered exponential. Therefore, in its current pleted the approximation diagram in Fig.1. We now form the method is not ideal. demonstrate the accuracy of these methods in various Let us make the same approximation we made in both limits on a model system. previous cases. If we assume that we may get rid of H¯ swift enough that s ≈ 0, then we may set H¯ (s) ≈ H¯ (0) = ¯ H in the generator Σ and the flow equations simplify as, V. DRIVEN ISING MODEL

dH(s, t) ¯ ¯ = −HT + it[HT ,H(s, t)], (45) To illustrate the quality of the approximations devel- ds oped in the previous section and presented in Fig.1, we where one lets s run from 0 to 1 to remove H¯ to lowest will apply them to a one-dimensional spin chain and com- order in t. That is, we are now implementing a unitary pare them with exact diagonalization results. We will transformation S(t) = e−itH¯ . Because this (in the sense consider the driven Ising model, that we reduce out the constant part of the Hamilto- X z z z x nian) does the reverse of a rotating frame transformation H(t) = (Jzσi σi+1 + Bz(t)σi + Bx(t)σi ), we dub this the “reverse” rotating frame approximation. i (50) Note that while we could calculate the unitary trans- Bx(t) = B sin(ωt); Bz(t) = B cos(ωt), formations exactly if H¯ is not too complicated, our for- x,y,z x,y,z mulation has the advantage that it enables us to use a where [σi , σj ] = 0 for i 6= j and on-site they fulfill truncated ansatz if needed. the Pauli algebra for spin-1/2 particles. 9

FIG. 1. This illustration shows the relation between the different approximations discussed in the text. A new approximation that is implied by symmetry is shown in the dashed box on the left side of the figure. We supply these in later sections of the manuscript. In the downward direction the approximations become progressively worse. To the left the approximations are expected to work better for larger constant parts of the Hamiltonian H¯ , and on the right for larger time-dependent parts of the Hamiltonian V (t). The crossed out arrows signify that the result cannot be recovered without going to higher order in the approximate Wilcox series.

This model was chosen because it has much of the One should note that in this case one was able to fully structure present in more complicated time-dependent write down an analytical result. problems because [V (t1),V (t2)] 6= 0 in general, which is a common feature of many systems of interest. Below we will derive expressions for all the different approxi- 2π B. Magnus series and Wilcox series mations (shown in Fig.1) that are valid at times T = ω . If we use Eq.(7) and Eq.(8) we find that at t = T , A. Dyson-Neumann series X z z Ω1 = W1 = −iT Jz σi σi+1, Inserting our Hamiltonian, Eq.(50), in Eq.(4) we find i that the following definitions are useful, 2   (53) T B X y B z z X Ω2 = W2 = i σ − Jz(σi+1 + σi−1) . H¯ = J σzσz , π i 2 z i i+1 i i 2 y Vy = B σi , (51) The approximate time evolution operators in the Magnus X y z z expansion UM and in the Wilcox approximation UW are, Vyz = 2JzB σi (σi+1 + σi−1), i U (T ) = eΩ1+Ω2 ; U (T ) = eW1 eW2 . (54) and the time evolution operator is approximately given M W as, One was able to find analytical expressions for the expo- T 2  i i  nents of the different contributions to the time evolution U(T ) ≈ 1 − iT H¯ − H¯ 2 + V − V . (52) 2 π yz π y operator. 10

C. Rotating frame approximation D. Reverse rotating frame approximation

Here we may use Eq.(37) to find H(1, t). The proce- Let us first find H(1, t) according to Eq.(45). The pro- dure is as follows. We start with a Hamiltonian that has cedure is as follows. We start with a Hamiltonian that the form of the original Hamiltonian, Eq.(50), but with has the form of the original Hamiltonian, Eq.(50), but arbitrary couplings. We then insert the Hamiltonian in with arbitrary couplings. We then insert the Hamiltonian the flow equations and add newly generated couplings to in the flow equations and add newly generated couplings the Hamiltonian. This could be stopped at some point to the Hamiltonian. This could be stopped at some point but here, because of the relatively simple structure of but here we are able to reach a point when no new cou- z x the external drive V (t) = Bz(t)σi +Bx(t)σi , we are able plings are generated. The couplings that contribute are x z z z y z x z z to reach a point when no new couplings are generated. the five {σi , σi , σi σi+1, σi σi±1, σi σi−1σi+1}. Therefore, The couplings that contribute are found to be the nine the Hamiltonian has the form, {σx,y,z, σx,y,zσx,y,z, σxσy,z , σyσz }. To be more precise, i i i+1 i i±1 i i±1 X x z y z y z the Hamiltonian in the rotating frame has the form, HRR = Cxσi + Czσi + Cyz(σi σi+1 + σi σi−1) i (58) z z x z z X x y z x x + Czzσi σi+1 + Cxzzσi σi−1σi+1. HR = [Cxσi + Cyσi + Czσi + Cxxσi σi+1 i The flow equations that correspond to this are given x y x y y y + Cxy(σi σi+1 + σi σi−1) + Cyyσi σi+1 (55) in the Appendix B, and their solutions as well. After x z x z z z averaging the couplings Eq.(B3) in the reverse rotating + Cxz(σi σi+1 + σi σi−1) + Czzσi σi+1 y z y z frame over one period, we find that the couplings are, + Cyz(σi σi+1 + σi σi−1)]. Bω2 sin 8πJz  ¯ ω The flow equations for the couplings, as well as the re- Cz = Czz = 0; Cyz = , 4π (ω2 − 16J 2) sults, are given in Appendix A. Averaging the results in z . (59) 2 2 4πJz  Eq.(A3) over one period, we find that at stroboscopic Bω sin ω Cx = Cxzz = 2 2 . times we have a Floquet Hamiltonian with couplings ap- 4π (ω − 16Jz ) proximately given as, Therefore, the time evolution operator at stroboscopic times is approximately, Cx = Cxy = Cxz = 0,

2 4B  2 8B  −iHT¯ −iHRRT 3Jz 3Jzω J2 ω 3Jzω J2 ω U(T ) ≈ e e , (60) Cxx = + 2 − 2 16 8B 128B ¯ 8B  4B  where H is given in Eq.(51). The first exponential factor J ωJ J ωJ ¯ + z 1 ω − z 1 ω , in Eq. (60), e−iHT , should be interpreted as the trans- 16B 2B formation to the frame in which HRR and the couplings   8B  Jz 1 8B JzωJ1 in Eq. (59) are valid. Hence Eq. (60) is valid only at C = + J J − ω , yy 4 2 z 2 ω 16B s = 1.   4B  8B  In addition we note that for this approximation we 1 8B JzωJ2 ω JzωJ2 ω Cyz = JzJ1 + − , were able to give analytical expressions for the exponents 2 ω 4B 16B in the time evolution operator. 9J 1 8B  3J ω2J 8B  C = z − J J + z 2 ω zz 16 2 z 2 ω 128B2 E. Fer approximation 3J ω2J 4B  J ωJ 4B  − z 2 ω + z 1 ω , 8B2 2B     In the Fer approximation the flow equations, Eq.(31), ω 1 4B 4B generate infinitely many terms. In its traditional form Cy = − ωJ0 − BJ1 , 4 4 ω ω the Fer approximation would therefore not be applica- 4B  ble for such a system. Our method using flow equations, C = BJ , z 2 ω however, allows one to truncate those terms and include (56) only terms that appeared in the rotating frame approxi- mation and in the reverse rotating frame approximation. where Jn is the n-th Bessel function of the first kind. The time evolution operator in this case is just, That is, we take X x y z x x HF = [Cxσ + Cyσ + Czσ + Cxxσ σ U(T ) ≈ e−iHRT , (57) i i i i i+1 i + C (σxσy + σxσy ) + C σyσy with the couplings above. Similar to the Magnus case, xy i i+1 i i−1 yy i i+1 (61) x z x z z z we found an analytical expression for the exponent in the + Cxz(σi σi+1 + σi σi−1) + Czzσi σi+1 y z y z x z z time evolution operator. + Cyz(σi σi+1 + σi σi−1) + Cxzzσi σi−1σi+1]. 11

1.0 The flow equations one finds for this ansatz are given in Dyson-Neumann Magnus Wilcox Appendix C. While they are analytically accessible, the RotFrame ReverseRotFrame explicit expressions for the couplings are far too compli- 0.8 Fer cated to be illuminating. Exact flow, trunc. ansatz 0.6

F. Truncated exact flow equations distance from0.4 exact 2 Since the Hamiltonian we consider in Eq.(50) has the iωt −iωt 0.2 form H = H0 + e H+ + e H−, we may make use normed l of Eq. (36) to derive exact flow equations, which can be 0.0 treated very conveniently numerically. Much like in the 5 10 15 20 25 30 35 case of Fer’s approximation this will generate infinitely ω many terms, which is why we took the same truncated ansatz, FIG. 2. Plot of the l2 distance, Eq.(64), between the differ- ent approximations and the exact time evolution operator for X x y z x x H0,+,− = [Cxσi + Cyσi + Czσi + Cxxσi σi+1 strong driving B = 4, Jz = 1 and L = 14 sites. Evidently, all i approximations perform better as the large frequency limit is x y x y y y approached. + Cxy(σi σi+1 + σi σi−1) + Cyyσi σi+1 x z x z z z + Cxz(σi σi+1 + σi σi−1) + Czzσi σi+1 + C (σyσz + σyσz ) + C σxσz σz ], yz i i+1 i i−1 xzz i i−1 i+1 The distance that is found via exact diagonalization of (62) systems with up to 16 sites using the QuSpin package.[67] for all three parts of the Hamiltonian. The resulting flow All evolution operators are evaluated at stroboscopic equations are sufficiently opaque that we do not exhibit times and the Hilbert space has dimension D. When them. both operators are unitary, l (U ,U ) ∈ [0, 1]. Zero cor- The result from a numerical analysis is an effective 2 A E responds to perfect agreement (UA = UE), and unity Hamiltonian of the form, corresponds to maximally separated unitary operators. X y z x x y y One should note that this measure sometimes overesti- H0 = [Cyσi + Czσi + Cxxσi σi+1 + Cyyσi σi+1 i (63) mates errors. This may occur, for example, when evaluat- z z y z y z ing the time evolution of local observables. For instance + Czzσ σ + Cyz(σ σ + σ σ )]. i i+1 i i+1 i i−1 [68, 69], found that local observables for a Trotterized Other terms from the truncated ansatz vanish up to time evolution like Eq.(2) are more robust to an increase numerical accuracy. However, they appear during the in time-step size dt than previously thought. The reason flow. While this method does not offer us analytic expres- for this is that local observables when evolved for suffi- sions for the couplings it still has advantages over brute ciently short times occupy only a small part of the total force exact diagonalization. One important advantage is Hilbert space. The method above as an error estimate that the method is scalable: one may include as many has contributions from all parts of the Hilbert space and terms in the ansatz as desired and therefore arrive at dif- therefore may overestimate the error for certain parts of ferent levels of numerical costs. Such an ansatz may be the Hilbert space. We therefore like to think of it as a motivated by physical considerations or mathematically worst case estimate or as an estimate for global properties by perturbation theory, such as we used. Furthermore, of the time evolution. While it is very interesting to find by using this method one has an explicit expression in out how well these different methods work for different terms of operators and may therefore do semi-analytical observables we will not attempt to discuss such behavior follow-up work. that is specific to a certain operator but rather discuss the generic properties captured in Eq.(64). We wish to determine how well the different approx- VI. COMPARING ALL THE imations perform as a function of frequency for differ- APPROXIMATIONS ent strengths of couplings. Let us first look at the limit of large driving strength, B. From Fig.2, one may see In this section we compare the validity of the differ- that, as expected, the Dyson-Neumann approximation ent approximations discussed in the previous sections by (dashed blue or in print dashed gray with circle markers) calculating the l2 distance, has the worst performance, and even reaches values above q 1, because it is not unitary. The reverse rotating frame 1 † l2(UA,UE) = √ tr((UA − UE)(UA − UE) ), (64) approximation (solid orange or in print solid gray with 2 D rectangle markers) performs poorly, too, which comes as between the various approximate time evolution oper- no surprise because it neglects B2 terms. We may choose ators, UA, and the exact time evolution operator, UE. the Magnus approximation over the Wilcox approxima- 12

1.0 functions of the couplings. We see this in Fig.4 for three Dyson-Neumann Magnus different frequencies. As expected we find the Dyson- 0.8 Wilcox Neumann series performs the worst across the board, RotFrame ReverseRotFrame and the Wilcox approximation is second worst in most 0.6 Fer cases. The Magnus approximation, as expected, is typ- Exact flow, trunc. ansatz ically third worst. For increased magnetic driving we 0.4 see that the exact but truncated flow equations and the distance from exact

2 rotating frame approximation are the most reliable with 0.2 the results for the truncated flow equations being slightly better.

normed l 0.0 In general we can see that the exact but truncated 20 40 60 80 flow equations for a wide range of variables (yet not all) ω yield the most reliable results. The reason it does not al- ways perform better is the arbitrarily chosen truncation

FIG. 3. Plot of the l2 distance between the different approxi- point. We find that the “reverse” rotating frame approx- mations and the exact time evolution operator for weak driv- imation we introduce is most useful in the intermediate ing B = 1, and strong static interaction Jz = 4, for a chain frequency regime at comparatively strong constant parts of L = 14 sites. in the Hamiltonian. To make the results more accessible we provide Table I summarizing the results for the first iteration of each procedure. tion because the Magnus approximation is more accurate. It should be emphasized that the checkmarks in the A recurring theme we find is that approximations that table do not capture that the Magnus approximation make life simpler generally perform more poorly: multi- is vastly better than the Wilcox approximation or the ple less complicated matrix exponentials in the Wilcox Dyson-Neumann approximation, but it should serve as a case would have been easier to calculate than one com- qualitative guide on which method to use. We would also plicated matrix exponential in the Magnus case. The Fer like to stress that the reverse rotating frame approxima- approximation performs slightly worse than the rotating tion makes a regime easier to access analytically when it frame approximation, which is likely due to the need to is not covered by the other approximations that are ana- truncate it at an arbitrary point. From the analytically lytically tractable. One should also note that it can very accessible approximations, the rotating frame approxi- easily be combined with a first order Magnus approxi- mation performs best. However, even it is outperformed mation, which would turn the two red checkmarks in the by the exact flow equations, including the case of a trun- high frequency or short time regime green because this cated ansatz. This example demonstrates that the flow reintroduces the order V 2 terms that were neglected. equations are indeed especially useful when looking for Floquet Hamiltonians. Next we consider the case of strong static parts in VII. CONCLUSION the Hamiltonian. The plot is given in Fig.3. We find that for the Dyson-Neumann series, for almost the full In this work, we introduce an analogue of Hamilton- range of values considered, unitarity is completely bro- Jacobi theory for the time-evolution operator of a quan- ken and l2 does not even appear within the range [0, 1]. tum many-particle system. The theory offers a useful ap- The reverse rotating frame approximation does best for proach to develop approximations to the time-evolution these large couplings. For most of the range of values operator, and also provides a unified framework and the Fer approximation performs similarly. As we found starting point for many well-known approximations to with a strong drive, the Magnus approximation outper- the time-evolution operator. forms the Wilcox approximation. The rotating frame ap- In the process we found a novel approximation to the proximation is only a slight improvement over the Mag- time-evolution operator, which is accurate if the constant nus approximation. The result from the truncated but part of the Hamiltonian is large compared to the time- exact flow equations for the range ω & 20 is compara- dependent part. This approximation may be useful in ble to the best approximations. For the range of values cavity QED applications as discussed earlier or more gen- ω . 20—presumably because of the truncation scheme— erally in cases where the constant part of Hamiltonian is it becomes uncontrolled. It is worth mentioning that, for large enough that the Magnus expansion will be an in- the truncated but exact flow equations, fewer couplings sufficient approximation despite a small external driving contribute to the effective Hamiltonian than in the Fer strength compared to the driving frequency. We were also case. Some of the couplings that appear in the Fer case able to show that one set of flow equations we derived in are zero for truncated flow equations, which to some ex- a prior work turns out to be especially powerful since tent explains the shorter range of validity of the approx- it offers the best approximation, even when truncated, imation. to the time-evolution operator while still being numeri- Let us next look at how the approximations behave as cally easily accessible. Unlike time ordered exponentials, 13

FIG. 4. Plot of the l2 distance between the different approximations and the exact time evolution operator as a function of the couplings for a chain with L = 16 sites. The upper row has Jz = 1 fixed with B being varied. The lower row has B = 1 fixed and Jz is varied. For both cases the left column has ω = 5, the middle column ω = 15 and the right column ω = 25. however, it also facilitates easy access to the Floquet ACKNOWLEDGMENTS Hamiltonian since coefficients in the Floquet Hamilto- nian can be calculated directly, which opens the road to We are grateful for funding under NSF DMREF Grant semi-analytic discussions of systems that are otherwise no DMR-1729588 and NSF Materials Research Science inaccessible. We hope that the flowchart we provided in and Engineering Center Grant No. DMR-1720595. Dur- Fig.1 will guide an understanding of the connections be- ing the writing of the manuscript, P.L. was supported tween different popular approximations. In addition, we by the Scientific Discovery through Advanced Comput- hope TableI we provided will make it easy for a reader ing (SciDAC) program funded by the US Department of to appropriately choose the right approximation for any Energy, Office of Science, Advanced Scientific Computing problem encountered. Research and Basic Energy Sciences, Division of Mate- rials Sciences and Engineering. GAF acknowledges sup- port from a Simons Fellowship, and a QuantEmX grant from ICAM and the Gordon and Betty Moore Founda- tion through Grant GBMF5305.

[1] Anatoli Polkovnikov, Krishnendu Sengupta, Alessandro [6] Achilleas Lazarides, Arnab Das, and Roderich Moessner. Silva, and Mukund Vengalattore. Colloquium: Nonequi- Periodic thermodynamics of isolated quantum systems. librium dynamics of closed interacting quantum systems. Phys. Rev. Lett., 112:150401, 2014. Rev. Mod. Phys., 83:863–883, 2011. [7] William Berdanier, Michael Kolodrubetz, Romain [2] Andr´eEckardt. Colloquium: Atomic quantum gases in Vasseur, and Joel E. Moore. Floquet dynamics of periodically driven optical lattices. Rev. Mod. Phys., boundary-driven systems at criticality. Phys. Rev. Lett., 89:011004, 2017. 118:260602, 2017. [3] Marin Bukov, Luca D’Alessio, and Anatoli Polkovnikov. [8] X. Wen and J.-Q. Wu. Floquet conformal field theory. Universal high-frequency behavior of periodically driven arXiv:1805.00031, 2018. systems: from dynamical stabilization to Floquet engi- [9] Vedika Khemani, Achilleas Lazarides, Roderich Moess- neering. Advances in Physics, 64(2):139–226, 2015. ner, and S. L. Sondhi. Phase structure of driven quantum [4] Jon H. Shirley. Solution of the Schr¨odingerequation with systems. Phys. Rev. Lett., 116:250401, 2016. a Hamiltonian periodic in time. Phys. Rev., 138:B979– [10] Pedro Ponte, Anushya Chandran, Z. Papi, and Dmitry A. B987, 1965. Abanin. Periodically driven ergodic and many-body lo- [5] R. Moessner and S. L. Sondhi. Equilibration and order calized quantum systems. Ann. Phys. (N.Y.), 353:196 – in quantum Floquet matter. Nat. Phys., 13:424, 2017. 204, 2015. 14

−1 −1 −1 t ∼ ω  H0,V t ∼ ω ∼ H0,V t ∼ ω  H0,V HF Remarks any V  H0 H0  V any V  H0 H0  V any V  H0 H0  V Time ordered exponential           Inaccessible analytically, Exact flow equations           numerics harder than (removing H or H ) 0 time ordered exponential Inaccessible analytically, Exact flow equations           numerics comparable to (removing V ) time ordered exponential Inaccessible analytically, Truncated exact flow numerics require a few           (removing V ) flow equations and one matrix exponential If truncated exponents accessible analytically, Fer expansion           numerics require easier flow equations but two matrix exponentials HF accessible by analytical treatments sometimes without truncation, Rotating frame           numerics require easier flow equations and one matrix exponential Exponents accessible analytically sometimes without truncation, Reverse rot. frame           numerics require easier flow equations but two matrix exponentials HF accessible analytically, Magnus approximation           numerics require one matrix exponential Exponents accessible by analytical means, Wilcox approximation           numerics require two matrix exponential Dyson-Neumann           Analytically accessible

TABLE I. Summary of performance quality of various approximations discussed in the text.

[11] Pedro Ponte, Z. Papi´c, Fran¸cois Huveneers, and [17] Jean Dalibard, Fabrice Gerbier, Gediminas Juzeli¯unas, Dmitry A. Abanin. Many-body localization in periodi- and Patrik Ohberg.¨ Colloquium: Artificial gauge poten- cally driven systems. Phys. Rev. Lett., 114:140401, 2015. tials for neutral atoms. Rev. Mod. Phys., 83:1523–1543, [12] Dmitry A. Abanin, Wojciech De Roeck, and Fran¸cois 2011. Huveneers. Exponentially slow heating in periodically [18] D. N. Basov, R. D. Averitt, and D. Hsieh. Towards driven many-body systems. Phys. Rev. Lett., 115:256803, properties on demand in quantum materials. Nat. Mat., 2015. 16:1077, 2017. [13] Dmitry A. Abanin, Wojciech De Roeck, Wen Wei Ho, and [19] J. Zhang and R.D. Averitt. Dynamics and control in Fran¸coisHuveneers. Effective Hamiltonians, prethermal- complex transition metal oxides. Annu. Rev. Mater. Res., ization, and slow energy absorption in periodically driven 44(1):19–43, 2014. many-body systems. Phys. Rev. B, 95:014112, Jan 2017. [20] D. N. Basov, Richard D. Averitt, Dirk van der Marel, [14] Takashi Mori, Tomotaka Kuwahara, and Keiji Saito. Rig- Martin Dressel, and Kristjan Haule. Electrodynamics of orous bound on energy absorption and generic relaxation correlated electron materials. Rev. Mod. Phys., 83:471– in periodically driven quantum systems. Phys. Rev. Lett., 541, 2011. 116:120401, 2016. [21] Claudio Giannetti, Massimo Capone, Daniele Fausti, [15] Dominic V. Else, Bela Bauer, and Chetan Nayak. Michele Fabrizio, Fulvio Parmigiani, and Dragan Mi- Prethermal phases of matter protected by time- hailovic. Ultrafast optical spectroscopy of strongly corre- translation symmetry. Phys. Rev. X, 7:011026, 2017. lated materials and high-temperature superconductors: [16] Immanuel Bloch, Jean Dalibard, and Wilhelm Zwerger. a non-equilibrium approach. Adv. Phys., 65(2):58–238, Many-body physics with ultracold gases. Rev. Mod. 2016. Phys., 80:885–964, 2008. 15

[22] M Gandolfi, G L Celardo, F Borgonovi, G Ferrini, [39] Takuya Kitagawa, Takashi Oka, Arne Brataas, Liang Fu, A Avella, F Banfi, and C Giannetti. Emergent ultra- and Eugene Demler. Transport properties of nonequilib- fast phenomena in correlated oxides and heterostruc- rium systems under the application of light: Photoin- tures. Phys. Scr., 92(3):034004, 2017. duced quantum Hall insulators without Landau levels. [23] N. Goldman, J. Dalibard, M. Aidelsburger, and N. R. Phys. Rev. B, 84:235108, 2011. Cooper. Periodically driven quantum matter: The case [40] Zhenghao Gu, H. A. Fertig, Daniel P. Arovas, and Assa of resonant modulations. Phys. Rev. A, 91:033632, Mar Auerbach. Floquet spectrum and transport through an 2015. irradiated graphene ribbon. Phys. Rev. Lett., 107:216601, [24] C. Sias, H. Lignier, Y. P. Singh, A. Zenesini, D. Ciampini, 2011. O. Morsch, and E. Arimondo. Observation of photon- [41] Mark S. Rudner, Netanel H. Lindner, Erez Berg, and assisted tunneling in optical lattices. Phys. Rev. Lett., Michael Levin. Anomalous edge states and the bulk-edge 100:040404, Feb 2008. correspondence for periodically driven two-dimensional [25] Alessandro Zenesini, Hans Lignier, Donatella Ciampini, systems. Phys. Rev. X, 3:031005, 2013. Oliver Morsch, and Ennio Arimondo. Coherent control [42] Gonzalo Usaj, P. M. Perez-Piskunow, L. E. F. Foa Torres, of dressed matter waves. Phys. Rev. Lett., 102:100403, and C. A. Balseiro. Irradiated graphene as a tunable 2009. Floquet topological insulator. Phys. Rev. B, 90:115423, [26] Andr´eEckardt, Christoph Weiss, and Martin Holthaus. 2014. Superfluid-insulator transition in a periodically driven [43] Pierre-Louis Giscard and Christian Bonhomme. Gen- optical lattice. Phys. Rev. Lett., 95:260404, 2005. eral solutions for quantum dynamical systems driven [27] M. Aidelsburger, M. Atala, S. Nascimb`ene,S. Trotzky, by time-varying Hamiltonians: applications to NMR. Y.-A. Chen, and I. Bloch. Experimental realization of arXiv:1905.04024, 2019. strong effective magnetic fields in an optical lattice. Phys. [44] F. J. Dyson. The radiation theories of Tomonaga, Rev. Lett., 107:255301, Dec 2011. Schwinger, and Feynman. Phys. Rev., 75:486–502, Feb [28] M. Aidelsburger, M. Lohse, C. Schweizer, M. Atala, J. T. 1949. Barreiro, S. Nascimb`ene,N. R. Cooper, I. Bloch, and [45] F. J. Dyson. The S matrix in quantum electrodynamics. N. Goldman. Measuring the Chern number of Hofstadter Phys. Rev., 75:1736–1755, Jun 1949. bands with ultracold bosonic atoms. Nat. Phys., 11:162– [46] F. J. Dyson. Divergence of perturbation theory in quan- 166, February 2015. tum electrodynamics. Phys. Rev., 85:631–632, Feb 1952. [29] D. Fausti, R. I. Tobey, N. Dean, S. Kaiser, A. Dienst, [47] S. Blanes, F. Casas, J.A. Oteo, and J. Ros. The Mag- M. C. Hoffmann, S. Pyon, T. Takayama, H. Takagi, and nus expansion and some of its applications. Phys. Rep., A. Cavalleri. Light-induced superconductivity in a stripe- 470(56):151 – 238, 2009. ordered cuprate. Science, 331(6014):189–191, 2011. [48] E.B. Fel’dman. On the convergence of the Magnus expan- [30] Andrea Cavalleri. Photo-induced superconductivity. sion for spin systems in periodic magnetic fields. Phys. Contemp. Phys., 59(1):31–46, 2018. Lett. A, 104(9):479 – 481, 1984. [31] L. Stojchevska, I. Vaskivskyi, T. Mertelj, P. Kusar, [49] Wilhelm Magnus. On the exponential solution of differ- D. Svetin, S. Brazovskii, and D. Mihailovic. Ultrafast ential equations for a linear operator. Communications switching to a stable hidden quantum state in an elec- on Pure and Applied Mathematics, 7(4):649–673, 1954. tronic crystal. Science, 344(6180):177–180, 2014. [50] S Klarsfeld and J A Oteo. Exponential infinite- [32] A. Lerose, J. Marino, A. Gambassi, and A. Silva. Quan- product representations of the time-displacement oper- tum many-body Kapitza phases of periodically driven ator. Journal of Physics A: Mathematical and General, spin systems. arXiv:1803.04490, 2018. 22(14):2687, 1989. [33] Y. A. Gerasimenko, I. Vaskivskyi, and D. Mi- [51] F M Fern´andez.On the Magnus, Wilcox and Fer operator hailovic. Long range electronic order in a metastable equations. European Journal of Physics, 26(1):151, 2005. state created by ultrafast topological transformation. [52] F Fer. Bull. Classe Sci. Acad. R. Belg., 44:818, 1958. arXiv:1704.08149, 2017. [53] R. M. Wilcox. Exponential operators and parameter dif- [34] J. Zhang, P. W. Hess, A. Kyprianidis, P. Becker, A. Lee, ferentiation in quantum physics. Journal of Mathematical J. Smith, G. Pagano, I.-D. Potirniche, A. C. Potter, Physics, 8(4):962–982, 1967. A. Vishwanath, N. Y. Yao, and C. Monroe. Observation [54] Michael Vogl, Pontus Laurell, Aaron D. Barr, and Gre- of a discrete time crystal. Nature, 543:217–220, 2017. gory A. Fiete. Flow equation approach to periodically [35] Soonwon Choi, Joonhee Choi, Renate Landig, Georg driven quantum systems. Phys. Rev. X, 9:021037, May Kucsko, Hengyun Zhou, Junichi Isoya, Fedor Jelezko, 2019. Shinobu Onoda, Hitoshi Sumiya, Vedika Khemani, Curt [55] Franz Wegner. Flow-equations for Hamiltonians. Ann. von Keyserlingk, Norman Y. Yao, Eugene Demler, der Phys. (Leipzig), 3:77, 1994. and Mikhail D. Lukin. Observation of discrete time- [56] S. Kehrein. The Flow Equation Approach to Many- crystalline order in a disordered dipolar many-body sys- Particle Systems. Springer Tracts in Modern Physics. tem. Nature, 543:221–225, 2017. Springer Berlin Heidelberg, 2007. [36] Andrew C. Potter, Takahiro Morimoto, and Ashvin Vish- [57] Albert Verdeny, Andreas Mielke, and Florian Mintert. wanath. Classification of interacting topological Floquet Accurate effective Hamiltonians via unitary flow in Flo- phases in one dimension. Phys. Rev. X, 6:041001, 2016. quet space. Phys. Rev. Lett., 111:175301, Oct 2013. [37] Takashi Oka and Hideo Aoki. Photovoltaic Hall effect in [58] Hideo Sambe. Steady states and quasienergies of a graphene. Phys. Rev. B, 79:081406, 2009. quantum-mechanical system in an oscillating field. Phys. [38] Netanel H. Lindner, Gil Refael, and Victor Galitski. Rev. A, 7:2203–2213, Jun 1973. Floquet topological insulator in semiconductor quantum [59] M Hafezi, M. D. Lukin, and J. M. Taylor. Non- wells. Nat. Phys., 7:490, 2011. equilibrium fractional quantum Hall state of light. New. 16

J. Phys., 15:063001, 2013. [65] H. Goldstein, C.P. Poole, and J.L. Safko. Classical Me- [60] J. E. Reiner, W. P. Smith, L. A. Orozco, H. M. Wiseman, chanics. Addison Wesley, 3rd edition, 2002. and Jay Gambetta. Quantum feedback in a weakly driven [66] C. Tomaras and S. Kehrein. Scaling approach for the cavity QED system. Phys. Rev. A, 70:023819, Aug 2004. time-dependent Kondo model. EPL (Europhysics Let- [61] M. Heimsoth, C. E. Creffield, and F. Sols. Weakly driven ters), 93(4):47011, 2011. quantum coherent ratchets in cold-atom systems. Phys. [67] Phillip Weinberg and Marin Bukov. QuSpin: a Python Rev. A, 82:023607, Aug 2010. Package for Dynamics and Exact Diagonalisation of [62] Tomotaka Kuwahara, Takashi Mori, and Keiji Saito. Quantum Many Body Systems part I: spin chains. Sci- Floquet-Magnus theory and generic transient dynamics Post Phys., 2:003, 2017. in periodically driven many-body quantum systems. An- [68] Markus Heyl, Philipp Hauke, and Peter Zoller. Quan- nals of Physics, 367:96 – 124, 2016. tum localization bounds Trotter errors in digital quan- [63] Ana Arnal, Fernando Casas, and Cristina Chiralt. On the tum simulation. arXiv:1806.11123, 2018. structure and convergence of the symmetric Zassenhaus [69] Lukas M. Sieberer, Tobias Olsacher, Andreas Elben, formula. Computer Physics Communications, 217:58 – Markus Heyl, Philipp Hauke, Fritz Haake, and Peter 65, 2017. Zoller. Digital Quantum Simulation, Trotter Errors, and [64] John P. Boyd. The Devil’s invention: Asymptotic, su- Quantum Chaos of the Kicked Top. arXiv:1812.05876, perasymptotic and hyperasymptotic series. Acta Appli- 2018. candae Mathematica, 56(1):1–98, Mar 1999.

Appendix A: Flow equations for rotating frame

We find that the flow equations for the rotating frame transformation of Eq. (55) for the Ising model are

dC (s) dC (s) x = 2C (s)F (t) − B sin(ωt); y = 2C (s)F (t) − 2C (s)F (t), ds y 2 ds z 1 x 2 dC (s) dC (s) z = −2C (s)F (t) − B cos(ωt); xx = 4C (s)F (t), ds y 1 ds xy 2 dC (s) dC (s) xy = −2C (s)F (t) + 2C (s)F (t) + 2C (s)F (t); yy = 4C (s)F (t) − 4C (s)F (t), (A1) ds xx 2 xz 1 yy 2 ds yz 1 xy 2 dC (s) dC (s) xz = 2C (s)F (t) − 2C (s)F (t); zz = −4C (s)F (t), ds yz 2 xy 1 ds yz 1 dC (s) yz = −2C (s)F (t) − 2C (s)F (t) + 2C (s)F (t), ds xz 2 yy 1 zz 1

B−B cos(ωt) B sin(ωt) where F1(t) = ω and F2(t) = ω and with the initial conditions

Cxx(0) = Cxy(0) = Cyy(0) = 0,

Cxz(0) = Cyz(0) = Cy(0) = 0, (A2) Czz(0) = Jz; Cx(0) = B sin(ωt),

Cz(0) = B cos(ωt). 17

Solving these equations one finds ! 2B sin ωt  C (1, t) = J sin2(ωt) sin4 2 , xx z ω ! ! ωt 2B sin ωt  4B sin ωt  C (1, t) = J sin sin(ωt) sin2 2 sin 2 , xy z 2 ω ω ! !! 1 2B sin ωt  4B sin ωt  C (1, t) = J 2 sin(2ωt) sin4 2 + sin(ωt) sin2 2 , xz 4 z ω ω ! ωt 4B sin ωt  C (1, t) = J sin2 sin2 2 , yy z 2 ω ! ! 1 ωt 8B sin ωt  ωt ωt 4B sin ωt  C (1, t) = J sin3 sin 2 + J sin cos2 sin 2 , yz 2 z 2 ω z 2 2 ω ! ! ! 1 ωt 8B sin ωt  4B sin ωt  C (1, t) = J 8 sin4 cos 2 + 8 sin2(ωt) cos 2 + 4 cos(ωt) + 3 cos(2ωt) + 9 , zz 16 z 2 ω ω ! !! 1 4B sin ωt  ωt 4B sin ωt  C (1, t) = 2B sin(ωt) cos 2 − ω cos sin 2 , x 4 ω 2 ω ! ! 1 2B sin ωt  ωt 4B sin ωt  C (1, t) = ω sin2 2 − B sin sin 2 , y 2 ω 2 ω ! !! 1 ωt 4B sin ωt  4B sin ωt  C (1, t) = ω sin sin 2 + 2B(cos(ωt) − 1) cos 2 . z 4 2 ω ω (A3) 18

Appendix B: Reverse rotating frame flow equations The initial conditions are

Cxx(0) = Cxy(0) = Cyy(0) = Cxz(0), The flow equations that are found with the Hamilto- nian given by (58) when inserted in (45) are = Cyz(0) = Cxzz(0) = Cy(0) = 0, Cx(0) = B sin(ωt); Cz(0) = B cos(ωt); Czz(0) = Jz. dC (s) (C2) yz = −2J tC (s) − 2J tC (s), ds z x z xzz The solutions are too tedious to write out and therefore dC (s) dC (s) not given. zz = −J ; z = 0, (B1) ds z ds dC (s) dC (s) xzz = x = 4J tC (s), ds ds z yz with initial conditions

Cyz(0) = Cxzz(0) = 0; Czz(0) = Jz, (B2) Cx(0) = B sin(ωt); Cz(0) = B cos(ωt).

Solving this we find that

Czz(1, t) = 0; Cz(1, t) = B cos(ωt), 2 Cx(1, t) = B cos (2Jzt) sin(ωt), 2 (B3) Cxzz(1, t) = −B sin (2Jzt) sin(ωt), 2 Cyz(1, t) = B cos (2Jzt) sin(ωt).

Appendix C: Fer approximation flow equations

The flow equations that are found with the Hamilto- nian given by (61) when inserted in (31) and are found as dC (s) xx = 4C (s)f (t), ds xy z dC (s) xy = −2C (s)f (t) + 2C (s)f (t) + 2C (s)f (t), ds xx z xz x yy z dC (s) yy = 4C (s)f (t) − 4C (s)f (t), ds yz x xy z dC (s) xz = −2C (s)f (t) + 2J tC (s) + 2C (s)f (t), ds xy x z y yz z dC (s) yz = −2J tC (s) − 2C (s)f (t) − 2J tC (s) ds z x xz z z xzz − 2Cyy(s)fx(t) + 2Czz(s)fx(t), dC (s) zz = −4C (s)f (t) − J , ds yz x z dC (s) xzz = 4J tC (s), ds z yz dC (s) x = −B sin(tω) + 2C (s)f (t) + 4J tC (s), ds y z z yz dC (s) y = −2C (s)f (t) − 4J tC (s) + 2C (s)f (t), ds x z z xz z x dC (s) z = −B cos(tω) − 2C (s)f (t), ds y x (C1) B(1−1 cos(tω)) B sin(tω) where fx(t) = ω and fz(t) = ω .