The time-dependent exchange-correlation functional for a Hubbard dimer: quantifying non-adiabatic effects

Johanna I. Fuks∗,1, 2 Mehdi Farzanehpour∗,1 Ilya V. Tokatly,1, 3 Heiko Appel,4 Stefan Kurth,1, 3 and Rubio1, 4 1Nano-Bio Spectroscopy group and ETSF, Dpto. F´ısica de Materiales, Universidad del Pa´ısVasco, Centro de F´ısica de Materiales CSIC-UPV/EHU-MPC and DIPC, Av. Tolosa 72, E-20018 San Sebasti´an,Spain 2Department of and Astronomy, Hunter College and the Graduate Center of the City University of New York, New York City, United States 3IKERBASQUE, Basque Foundation for Science, E-48011 Bilbao, Spain 4Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4-6, D-14195 Berlin, Germany (Dated: January 13, 2021) We address and quantify the role of non-adiabaticity (”memory effects”) in the exchange- correlation (xc) functional of time-dependent density functional theory (TDDFT) for describing non-linear dynamics of many-body systems. Time-dependent resonant processes are particularly challenging for available TDDFT approximations, due to their strong non-linear and non-adiabatic character. None of the known approximate density functionals are able to cope with this class of problems in a satisfactory manner. In this work we look at the prototypical example of the resonant processes by considering Rabi oscillations within the exactly soluble 2-site Hubbard model. We con- struct the exact adiabatic xc functional and show that (i) it does not reproduce correctly resonant Rabi dynamics, (ii) there is a sizable non-adiabatic contribution to the exact xc potential, which turns out to be small only at the beginning and at the end of the Rabi cycle when the ground state population is dominant. We then propose a ”two-level” approximation for the time-dependent xc potential which can capture Rabi dynamics in the 2-site problem. It works well both for resonant and for detuned Rabi oscillations and becomes essentially exact in the linear response regime. This new, fully non-adiabatic and explicit density functional constitutes one of the main results of the present work.

PACS numbers: 31.15.ee,42.65.-k,71.15.Mb

I. INTRODUCTION uses the instantaneous density as input for an approxi- mate ground-state functional. Thus, this approximation completely neglects both the history and the initial-state Due to the favorable balance between efficiency dependence of the exact functional. and accuracy, time-dependent density functional theory (TDDFT) is becoming the theory of choice to describe The successes and failures of the adiabatic approxima- the interaction of many-electron systems with external tion to describe linear response phenomena have been ad- electromagnetic fields of arbitrary intensity, shape and dressed in many works [1, 2, 15–17]. However, much less time dependence. Within this theory the is known about the performance of adiabatic TDDFT for are expressible as functionals of the time-dependent den- general dynamics beyond linear response. In some of our sity. Similarly to static DFT, in TDDFT one can de- past studies [8, 11, 18, 19] on one-dimensional model sys- fine an auxiliary non-interactiong Kohn-Sham (KS) sys- tems we have shown that adiabatic xc functionals fail to tem which reproduces the exact time-dependent dynam- describe dynamical processes where the density changes ics of the density. It is the propagation of this aux- significantly in time (e. g. in photo-physical and chem- iliary system in an (unknown) effective local potential ical processes where valence electrons are promoted to which makes TDDFT computationally powerful. How- empty states). There are few cases where the exact time- ever, despite of the great success of the theory in de- dependent xc potential is known and can thus be used to scribing optical properties of a large variety of molecules test approximations [11, 20, 21]. In these works it has and nanostructures [1–3], the available approximations been shown numerically that novel dynamical steps ap- arXiv:1312.1667v1 [cond-mat.other] 5 Dec 2013 for the exchange-correlation (xc) potential exhibit seri- pear in the xc potential which are fundamental to capture ous deficiencies in the description of non-linear processes, the proper resonant versus non-resonant dynamics and long range charge transfer [4–6] and double excitations charge localization. While the construction of accurate [7–9], to mention a few. approximations to the exact universal xc functional of The theoretical challenge is to improve the available TDDFT for Coulomb systems remains a challenge, sim- functionals in order to capture the nonlocality both in ple model Hamiltonians constitute a convenient frame- space and time of the exact xc functional which depends work to gain insights into the properties of the exact on the entire history of the density, the initial (interact- TDDFT functional. ing) many-body state and the initial KS state [10–14]. In the present work we exploit the possibilities of a We note that almost all TDDFT calculations today use solvable lattice model – the 2-site Hubbard model [22– an adiabatic approximation for the xc potential, which 24] – to address the impact of non-locality in time in 2 the exchange correlation functional of TDDFT. Specif- oscillations in interacting systems as well as the main ically, we study resonant Rabi oscillations, a prototyp- difficulties of describing Rabi dynamics within TDDFT. ical example of non-linear external field driven dynam- The many-body time-dependent Schr¨odingerequation, ics where the population of states changes dramatically in time. We first derive here the exact ground-state i∂t|ψ(t)i = H(t)|ψ(t)i, (2) Hartree-exchange-correlation (Hxc) functional for the 2- describes the evolution of the system from a given initial site model using the Levy-Lieb constrained search[25–27]. state |ψ i. Since the Hamiltonian (1) is independent of This functional, when used in a TDDFT context with 0 , the spin structure of the wave function |ψ(t)i is the instantaneous time-dependent density as input, con- fixed by the initial state. In the following we study the stitutes the exact adiabatic approximation which can be evolution from the ground state of the Hubbard dimer used as a reference to quantify the role of memory effects. and therefore it is sufficient to consider only the singlet By carefully studying and quantifying the dynamics pro- sector of our model. duced by TDDFT with the adiabatic Hxc potential we In the absence of an external potential, v = 0, the demonstrate that it fails both quantitatively and qualita- 1,2 stationary singlet eigenstates of the Hamiltonian (1) take tively to describe Rabi oscillations. In the second part of the form this work we apply an analytic density-potential map for   lattice systems [10, 28] to derive an explicit, fully non- † † † † † † † †  |gi = Ng cˆ1↑cˆ1↓ +c ˆ2↑cˆ2↓ + β+ cˆ1↑cˆ2↓ − cˆ1↓cˆ2↑ |0i(3a), adiabatic xc density functional which correctly captures √ † † † †  all features of Rabi dynamics in the Hubbard dimer. This |e1i = 1/ 2 cˆ1↑cˆ1↓ +c ˆ2↑cˆ2↓ |0i, (3b) functional is one of the main results of this paper.   |e i = N cˆ† cˆ† +c ˆ† cˆ† + β cˆ† cˆ† − cˆ† cˆ†  |0i, The paper is organized as follows: in Sec. II we in- 2 e2 1↑ 1↓ 2↑ 2↓ − 1↑ 2↓ 1↓ 2↑ troduce the physics of the Rabi effect for the Hubbard (3c) dimer, showing how the dipole moment and state occu- pations evolve with time during the course of resonant Rabi oscillations. In Sec. III we address the same prob- Here |0i is the vacuum state, |gi is the ground state, and lem from a TDDFT perspective. In particular we use |e1,2i are two excited singlet states. The Ng/e2 = (2 + the exact adiabatic xc functional as a reference to quan- 2 −1/2 2β±) are normalization factors and the coefficients tify memory effects. In the Sec. IV we consider the exact β± are defined as interacting system in a two-level approximation which al- p 2 2 lows us to derive a new approximate Hxc potential as an β± = (U ± 16T + U )/4T. (4) explicit functional of the time-dependent density. The excellent performance of this approximation is demon- The energy eigenvalues corresponding to the eigenstates strated and explained. We end the paper with our con- (3) are clusions in Sec. V. In the Appendix we derive the exact E = 2T β , (5a) ground state xc potential for the Hubbard dimer using g − the Levy-Lieb constrained search. Ee1 = U, (5b)

Ee2 = 2T β+ . (5c)

II. RABI OSCILLATIONS FOR TWO-SITE To simplify notations, we rewrite the external potential HUBBARD MODEL part in Eq. (1) in the form X ∆v (v nˆ + v nˆ ) = (ˆn −nˆ )+C(t)(ˆn +n ˆ ) (6) We consider the dynamics of two electrons on a Hub- 1 1σ 2 2σ 2 1 2 1 2 bard dimer, that is, a two-site interacting Hubbard model σ with on-site repulsion U and hopping parameter T . The P wheren ˆi = nˆiσ is the operator of the number of Hamiltonian of the system reads σ particles on site i, ∆v = v1 −v2 is the difference of on-site potentials, and C(t) = (v (t)+v (t))/2. The last term in ˆ X  † †  1 2 H = − T cˆ1σcˆ2σ +c ˆ2σcˆ1σ + U (ˆn1↑nˆ1↓ +n ˆ2↑nˆ2↓) Eq. (6) corresponds to a spatially uniform potential. This σ term can be trivially gauged away and will be ignored X + (v1(t)ˆn1σ + v2(t)ˆn2σ) , (1) in the following without loss of generality. Nontrivial σ physical effects come only from the external potential ∆v which is coupled to the difference of on-site densities. † ˆ wherec ˆiσ andc ˆiσ are creation and annihilation operators The quantity d =n ˆ1 −nˆ2 can be interpreted as the dipole for a spin-σ electron on site i, respectively. Then ˆiσ = moment of our simplified model of a diatomic system † ˆ cˆiσcˆiσ are the operators for the spin-σ density at site i, and its expectation value d(t) = hψ(t)|d|ψ(t)i uniquely and the v1,2(t) are time-dependent on-site potentials. We determines the on-site densities n1(t) and n2(t) if the use ~ = e = 1 throughout this work. Energies are given total number of particles is fixed. In the following, in in units of the hopping parameter T . As we will see, this particular for TDDFT, we will use the dipole moment simple model captures most qualitative features of Rabi d(t) as the basic “density variable”. 3

Since the dipole moment operator dˆis odd under reflec- tion (interchange of site indices), it has nonzero matrix elements only between states of different parity. In par- 1 ticular, dˆ connects the ground state |gi of Eq. (3a) only 0 to the first excited state |e1i d(t)

-1 d(t) ˆ 2 dge = hg|d|e1i = , (7) q 2 1 1 + β+ pg

p(t) p ˆ e1 while the matrix element of d between the ground state pe2 ˆ and the second excited state vanishes, hg|d|e2i = 0. 0 Now we are ready to discuss Rabi oscillations in the 10 20 30 40 50 Hubbard dimer. Let us consider the evolution of the t systems from its ground state |ψ(0)i = |gi under the action of a time periodic potential Figure 1. (Color online) Rabi oscillations for resonant laser ∆v(t) = 2E0 sin(ωt). (8) ω = ω0 = 2.56 T . Upper panel: dipole moment d(t). 2 Lower panel: Population of ground state pg = |hg|ψi| (solid red),first excited state p = |he |ψi|2 (dotted orange) and The Rabi regime of dynamics occurs when the frequency e1 1 second excited state p = |he |ψi|2 (dashed green). Time is ω of a sufficiently weak driving field approaches the fre- e2 2 given in units of 1/T , where T is the hopping parameter. quency ω0 of the main dipole . In our case this corresponds to the frequency ω ∼ ω0 = Ee1 − Eg close to the energy difference between ground and first excited For a singlet state both KS particles occupy the same states, and the amplitude E  ω /d . 0 0 ge one-particle KS orbital, which is described by two on- Fig. 1 shows resonant dynamics of the dipole moment site amplitudes ϕ1(t) and ϕ2(t). Therefore the time- and state populations obtained by the numerical propa- dependent KS equations reduce to a single 2 × 2 one- gation of Eq. (2) for a moderately strong interaction U = particle Schr¨odingerequation of theform T = 1, frequency ω = ω0 = 2.56, dge = 1.23, amplitude E0 = 0.1 and fixed electron number N = n1 +n2 = 2. We ∆vs 2 2 i∂tϕ1 = −T ϕ2 + ϕ1, (9a) see that the populations pg = |hg|ψi| and pe1 = |he1|ψi| 2 of the ground and the first exited state oscillate between ∆v i∂ ϕ = −T ϕ − s ϕ . (9b) zero and one, while the second excited state stays prac- t 2 1 2 2 2 tically unpopulated, pe2 = |he2|ψi| ≈ 0. The dipole moment shows fast oscillations at the driving frequency As our dynamics starts from the ground state, Eq. (9) has ω superimposed with slow oscillations of the envelope at to√ be solved with the initial condition ϕ1(0) = ϕ2(0) = the Rabi frequency ΩR = dgeE0. The maximal value of 1/ 2 which corresponds to the noninteracting KS ground the dipole moment |dmax| = dge = 1.23 is reached at 1/4 state. By definition the KS potential ∆vs(t) entering and 3/4 of the Rabi cycle when the ground and the first Eq. (9) produces a prescribed (interacting) dipole mo- excited states have equal populations of 1/2. ment. In the present case this KS potential can be found The main characteristic feature of the Rabi regime is explicitly as a functional of the density d(t) [10, 28], a strong variation of the state populations. It is this ¨ 2 feature which makes the description of Rabi oscillations d + 4T d ∆vs[d] = −q . (10) one of the most difficult cases for TDDFT [18, 29]. In 4T 2 (4 − d2) − d˙2 the rest of this paper we discuss the TDDFT approach to the Rabi dynamics for our simple two-site system. It is important to note that the functional ∆vs[d] is given by Eq. (10) only if the system evolves from, and remains sufficiently close to, the ground state. More precisely, it is III. TIME-DEPENDENT KOHN-SHAM shown in Ref. [10] that the functional form of Eq. (10) is EQUATIONS FOR A HUBBARD DIMER valid as long as the condition | arg(ϕ1) − arg(ϕ2)| < π/2 is satisfied during the course of the evolution. If the In the present two-electron case the KS system corre- opposite inequality holds, the overall sign on the right sponds to two non-interacting particles which reproduce hand side of Eq. (10) has to be changed from − to the time dependent dipole moment d(t) of the interact- +. Moreover, the sign changes every time the line ing system. The KS Hamiltonian has the form of Eq. (1) | arg(ϕ1) − arg(ϕ2)| = π/2 is crossed. In terms of the but with no interaction (U = 0) and the external poten- dipole moment, crossing this line corresponds to a van- tial ∆vs is chosen such that the correct time-dependent ishing expression under the square root in Eq. (10) [30]. density of the interacting system is reproduced. The above behavior can be viewed as a manifestation 4 of the initial state and history dependence in TDDFT advantage of the present simple model is that we know [10, 31]. explicitly the functional dependence of the exact ground- The exact KS potential can be calculated by inserting state Hxc potential (see Appendix), i. e., we do not need into Eq. (10) the exact dipole moment d(t) obtained from any a priori knowledge of the time-dependent density. a numerical solution of the many-body Schr¨odingerequa- To test the performance of the adiabatically-exact tion (2). In order to get the Hartree-exchange-correlation functional in the regime of Rabi oscillations we propa- ad (Hxc ) potential we subtract the physical external poten- gate self-consistently the KS equations with vHxc [d](t) tial ∆v from the KS potential, for the same parameters as in Sec. II.

∆vHxc = ∆vs − ∆v. (11)

Resonant The time dependence of the exact ∆vHxc which corre- 1 sponds to the dipole moment d(t) presented in Fig. 1 0 (i. e., to the regime of resonant Rabi oscillations, de- d(t) scribed in Sec. II) is shown in the top panel of Fig. 3. -1 In practice the exact Hxc functional is unknown and one has to rely on approximations. The simplest and the Detuned 1 most common approximation in TDDFT is based on the adiabatic assumption for xc effects. Below we present and 0 test the adiabatic approximation for our model system. d(t) -1 (t) d (t) 10 20 30 40 50 60 A. Adiabatically exact functional t To construct the adiabatic approximation for the Hubbard dimer we first find the exact ground-state Figure 2. (Color online) Upper panel : d(t) (solid blue) Hxc functional by the Levy-Lieb constrained search, i.e., in the presence of a laser of frequency ω = ω0 = 2.56 T, we perform an exhaustive search over the space of all al- compared to dad(t) (dotted red) propagated using the ex- gs lowed two-particle wave functions Ψ that yield a given act ground state functional ∆vHxc [d] in the presence of a laser resonant with the adiabatically-exact linear response fre- dipole moment d to find the Hohenberg-Kohn energy LR quency ωad = 2.60 T. Lower panel: d(t) for slightly detuned functional FHK[d], laser ω = ω0 + 0.03 (solid blue) compared to dad(t) using LR ˆ ˆ ω = ωad + 0.03 (dotted red). Time is given in units of the FHK[d] = minhΨ|T + U|Ψi, (12) Ψ→d inverse of the hopping parameter T . ˆ ˆ where T and U are operators of the kinetic energy and The results of the propagation confirm a general con- the interaction energy, i. e., the first and the second terms clusion of Ref. [18] about the presence of an artificial dy- in the Hamiltonian (1), respectively. namical detuning in the description of Rabi oscillations The exact ground state Hxc potential is given by the using adiabatic functionals. In Fig. 2 we compare the derivative of the Hxc energy with respect to the dipole evolution of the exact dipole moment d(t) (blue) with the moment, dipole moment dad(t) (red) obtained from KS equations with the adiabatically-exact Hxc potential for resonant gs ∂ s  ∆vHxc [d] = 2 FHK[d] − T [d] (13) (upper panel) and slightly detuned (lower panel) applied ∂d lasers. The upper panel shows the dynamics at resonant where T s[d] is the kinetic energy functional that is de- conditions when the frequency ω of the driving field is fined by Eq. (12) with U = 0. More details on this con- equal to the frequency ωres of the main dipole resonance. struction can be found in the Appendix. In the exact interacting system this frequency is obvi- In the adiabatically-exact approximation the exact ously ωres = ω0, while in the approximate TDDFT it ground-state Hxc potential of Eq. (13) is used in the is approximation-dependent, and should be determined time-dependent KS equations, i.e., the Hxc potential at consistently as the frequency ωres = ωLR of the corre- time t is calculated by inserting the instantaneous value sponding linear response resonance. At first sight, the of d(t) into the ground-state functional function dad(t) resulting from TDDFT with the adiabatic Hxc potential (upper panel in red on Fig. 2) looks quali- ∆vad [d](t) = ∆vgs [d(t)]. (14) tatively similar to the exact d(t) (upper panel in blue on Hxc Hxc Fig. 2). However, there is a deep difference in the under- We note that the adiabatically-exact Hxc potential lying microscopic dynamics. The physical system returns has also been found numerically for real-space one- to its initial state after two periods of the dipole mo- dimensional two-electron systems in Refs. [11, 19, 20] us- ment’s envelope, which corresponds to the Rabi period of 2π TR = = 51.10. In contrast, the microscopic period ing the iterative procedure introduced in Ref. [20]. A big E0dge 5 of the KS system with the adiabatic Hxc potential coin- cides with that of the dipole moment, which is the char- 1 acteristic feature of detuned Rabi oscillations. In fact, the KS Rabi dynamics is always internally detuned by gs 0.5 the presence of the adiabatic potential ∆vHxc [d(t)] which depends on the instantaneous density [18]. While this important difference is hidden in the case of resonant dy-

Δv 0 namics of the dipole moment, it is revealed immediately when the driving frequency is a bit shifted (detuned) from -0.5 the exact resonance. The dipole moments d(t) and dad(t) for a slightly detuned driving field with ω = ωres + 0.03 ΔvHxc are presented in the lower panel on Fig. 2. The exact ad -1 ΔvHxc - ΔvHxc dipole moment d(t) develops a “neck” at t ∼ TR showing that the actual physical period is indeed TR ≈ 50. On 10 20 30 40 50 60 the other hand, the function dad(t) is practically unaf- t fected by the external detuning because the KS system, being already strongly detuned internally, is insensitive Figure 3. (Color online) Time-dependent Hxc potential to small external variations of the driving frequency. This ∆vHxc (t) (in units of the hopping parameter T ) (solid blue) qualitative failure of the adiabatic approximation clearly and its non-adiabatic contribution defined as ∆vHxc (t) − gs demonstrates the important role of xc memory effects in ∆vHxc [d(t)] (in units of the hopping parameter T ) (dotted the correct description of Rabi oscillations. red). Time is given in units of 1/T . To further quantify non-adiabatic effects in the Rabi regime we extract a non-adiabatic contribution to the total Hxc potential. Namely, we subtract the adiabatic of an effective two-level system. gs potential ∆vHxc [d(t)] evaluated at the exact dipole mo- Let us assume that the second excited state |e2i is not ment d(t) from the exact ∆vHxc (t) defined by Eqs. (11) participating in the dynamics. We write the many-body and (10). In Fig. 3 we present the non-adiabatic part Schr¨odingerequation (2) in the two level approximation of Hxc potential together with the exact ∆vHxc (t). The as non-adiabatic contribution to ∆vHxc (t) turns out to be more than double the amplitude of the external potential ∆v(t) i∂tψg(t) = Egψg(t) + dge ψe(t) , (15a) and in fact as large as the Hxc potential itself during a 2 significant part of the Rabi-cycle. Not surprisingly, the ∆v(t) i∂ ψ (t) = E ψ (t) + d ψ (t) , (15b) non-adiabatic effects are small at the beginning and at t e e1 e ge 2 g the end of the Rabi cycle when the ground-state popu- lation is dominant and the system is close to the linear where ψg(t) = hg|ψ(t)i, and ψe(t) = he1|ψ(t)i are the response regime. But they grow fast when the system is projections of the time-dependent wave function onto the driven away from the ground-state and they remain large ground and first excited state, respectively. for a large part of the Rabi-cycle. It is interesting to no- By rotating the basis we can represent Eq. (15) in the tice that, centered around TR/2, there is a long period of form of a Schr¨odingerequation for one particle on an time during which the amplitude of the adiabatic effects effective “two-site lattice”. In other words, Eq. (15) is remains almost constant (see Fig. 3). unitarily equivalent to Eq. (9) with hopping constant Apparently a better approximation for the xc potential ω0/2 and external potential given by dge∆v/2. Using is needed to capture non-adiabatic effects relevant to de- this mapping and the KS potential ∆vs of Eq. (10) we scribe Rabi oscillations. In the next section we propose can immediately write the external potential ∆v of the an explicit non-adiabatic density functional based on a interacting system as a functional of the dipole moment two-level description of the interacting system. d   p ¨ 2 2L (−1) d + ω0d IV. TIME-DEPENDENT XC POTENTIAL IN ∆v [d] = q  , (16) dge 2 2 2 ˙2 THE TWO-LEVEL APPROXIMATION ω0(dge − d ) − d

In general, the Hxc functional ∆vHxc [d] can be found where the integer p counts how many times the square via Eq. (11) if we know the external potential as a func- root turns into zero during the evolution. The fac- tional of d(t). The presence of interactions makes the tor (−1)p accounts for the sign changes explained after problem of finding the functional ∆v[d] highly nontrivial Eq. (10). even in our simple model. Fortunately in some cases like In order to find the Hxc potential ∆vHxc [d] as a func- the Rabi oscillations the problem is simplified dramati- tional of the dipole moment we substitute the external cally because the behavior of the system is close to that potential ∆v of Eq. (16) and the KS potential of Eq. (10) 6 into Eq. (11): where the d(0) and d˙(0) are fixed by the KS-initial state d¨+ 4T 2d d(0) = 2(|ϕ (0)|2 − |ϕ (0)|2), (21a) ∆v2L [d] = − 1 2 Hxc q ˙ ∗ 4T 2 (4 − d2) − d˙2 d(0) = −4T Im[ϕ1(0)ϕ2(0)]. (21b)   After this preliminaries we can plug the Hxc potential p ¨ 2 (−1) d + ω0d Eq. (17) into the KS equations and propagate them self- − q  (17) dge 2 2 2 ˙2 consistently to test the performance of our non-adiabatic ω0(dge − d ) − d approximation. It is, however, clear that the functional 2L This expression is one of the main results of the present ∆vHxc [d], by construction, should exactly reproduce the paper. It provides us with an explicit fully non-adiabatic results of the two-level approximation to the full interact- density functional which, by construction, should cor- ing problem. Therefore TDDFT with the Hxc potential rectly describe the Rabi oscillations. It is worth em- of Eq. (17) is as accurate as the two-level approximation phasizing that the functional Eq. (17) contains history itself. dependence via the integer p in the second term. One can easily check that the non-linear functional ∆v2L of Eq. (17) produces the exact dynamic xc ker- Hxc Resonant nel in the linear response regime. The formally exact 1 Hxc functional in the linear response can be written as 0 follows d(t)

LR  −1 −1  -1 ∆vHxc [d](ω) = χs (ω) − χ (ω) d = fxc(ω)d, (18) 1 Detuned where χs(ω) and χ(ω) are the density response functions for the KS and the interactiving system, respectively. By 0 definition the term in parentheses is the exact exchange- d(t) d(t) correlation kernel fxc(ω). Since the eigenfunctions for -1 d2L(t) the Hubbard dimer are known, Eqs. (3) and (5), we can write the exact response functions χ(ω) in the Lehmann 10 20 30 40 50 60 representation [32], and substitute it into Eq. (18). The t result takes the following form   Figure 4. (Color online) Upper panel: d(t) (solid blue) for res- LR ω0 1 1  2 ∆vHxc [d](ω) = T − 2 + − 2 ω d. onant laser frequency ω = ω0 = 2.56 T compared to two-level 2dge 4T 2ω0dge 2L 2L approximation d (t) (dashed brown) using ∆vHxc , Eq. (17), (19) and same laser frequency. Lower panel: d(t) (solid blue) and 2L It is now straightforward to see that this equation is iden- d (t) (dashed red) for slightly detuned laser ω = ω0 + 0.03 T tical to the linearized version of the approximate func- (same detuning as lower panel on Fig. 2). Time is given in 2L tional ∆vHxc [d] defined by Eq. (17) with p = 0. In units of 1/T . other words our approximation becomes exact in the lin- ear regime. This nice property is not accidental because In Fig. 4 we compare the exact resonant Rabi dy- the functional of Eq. (17) is based on the two-level ap- namics of the dipole moment with the one obtained in proximation. In the linear response regime, the symmet- the two-level approximation or, alternatively, by solv- ric Hubbard dimer becomes an effective two-level system ing self-consistently the KS equations with the potential because the dipole transition matrix element between the 2L ∆vHxc [d]. The exact and approximate dipole moments ground state |gi and the second excited state |e2i van- are practically on top of each other. The non-adiabatic ishes. functional of Eq. (17) excellently reproduces Rabi oscilla- A subtle property of the non-adiabatic functional tions for a resonant excitation. Apparently it also works Eq. (17) is the dependence on the second time deriva- perfectly for detuned Rabi dynamics provided the de- ¨ ¨ tive d of the dipole moment. The presence of d does not tuning is not too large. Another nice property of this mean that the xc potential assumes a dependence on the approximation is that it becomes essentially exact for future. In general the existence theorem for TDDFT on a sufficiently weak non-resonant driving potential which a lattice [10] requires the second time derivative of the corresponds to the linear response regime. density to be continuous. Therefore d¨ can be calculated as a left limit for any time greater than the initial time, t > 0. At t = 0 the value of d¨(0) is determined by the V. CONCLUSION initial value of the external potential as follows q We use a Hubbard dimer to analyze, both qualitatively ¨ 2 2 2  ˙2 2 d(0) = −dde∆v(0) ω0 dge − d (0) − d (0) − ω0d(0), and quantitatively, the non-adiabatic features present in (20) the TDDFT functional. For this model system the exact 7

Kohn-Sham potential is analytic and moreover, the exact restrict the search in Eq. (12) to singlet wavefunctions ground state functional can be found by Levy-Lieb con- only. As a basis for the singlet sector we use the eigen- strained search. The later is propagated self-consistently states of Eq. (3). Then the most general singlet state to study the performance of the adiabatic approxima- may be written as tion. We show that non-adiabaticity is crucial to prop- erly capture the physics of resonant and nearly detuned |Ψi = A1|gi + A2|e1i + A3|e2i, (A.1) Rabi oscillations. The observed non-adiabatic features where we can, without loss of generality, choose the coef- grow as the population of the excited state is rising to its ficients Ai to be real. In the chosen basis, the expectation ˆ ˆ ˆ maximum, becoming even larger than the external poten- value of H0 = T + U takes the simple form tial. Lack of these features in adiabatic functionals causes ˆ ˆ ˆ 2 2 2 them to fail to describe Rabi dynamics, missing both fre- hΨ|H0|Ψi = hΨ|T + U|Ψi = EgA1 + Ee1 A2 + Ee2 A3 quency and amplitude of the physical dipole moment. (A.2) Taking advantage of the fact that under the action of a where the eigenvalues Eg, Ee1 , and Ee2 of the basis func- resonant laser the system behaves as an effective two-level tions are given by Eq. (5). one, we derive an explicit non-adiabatic functional that The expansion coefficients Ai in Eq. (A.1) are not inde- accurately reproduces resonant and slightly detuned Rabi pendent. The normalization condition of the wavefunc- oscillations. This fully non-adiabatic functional incorpo- tion |Ψi leads to rates explicitely the initial-state dependence and becomes 2 2 2 exact in the linear response regime. A1 + A2 + A3 = 1 . (A.3) The present work was focussed on the TDDFT de- In the constrained search we also have to make sure that scription of Rabi oscillations in a minimal model sys- we are only searching over wavefunctions which yield a tem. While the construction of the quasi-exact TDDFT given “density” d. This gives a second condition on the for this minimalistic system was non-trivial, the non- coefficients which reads adiabatic part of the Hxc potential has been found to have a relatively simple structure (see Fig. 3): while non- A1 + β+A3 d = hΨ|nˆ1 − nˆ2|Ψi = 4 A2 (A.4) adiabaticity is small close to the beginning and the end q 2 of the Rabi cycle, its amplitude is signifcant but almost 1 + β+ constant throughout a large middle part of the Rabi cy- cle. This might be a useful observation when aiming to where β+ is given by Eq. (4). We can use Eqs. (A.3) construct non-adiabatic TDDFT functionals applicable and (A.4) to eliminate two of the coefficients, say A2 to realistic systems. and A3, in the constrained search which then becomes a minimization in a single variable, i.e., We acknowledge financial support from the European Research Council Advanced Grant DYNamo (ERC-2010- FHK(d) = min hΨ(A1, d)|Tˆ + Uˆ|Ψ(A1, d)i . (A.5) AdG-267374), Spanish Grant (FIS2010-21282-C02-01), A1 Grupos Consolidados UPV/EHU del Gobierno Vasco (IT578-13), Ikerbasque and the European Commission In general, this minimization has to be carried out nu- HK projects CRONOS (Grant number 280879-2 CRONOS merically. In Fig. 5 we show F as function of the dipole HK CP-FP7). moment for various values of U. We note that F (d) is always minimal at d = 0. For large values of U the slope of F HK(d) changes rapidly as one crosses from negative to positive values of d. For vanishing interaction U = 0, the Appendix: Exact Hohenberg-Kohn functional of the minimization can be carried out fully analytically. The two-site Hubbard model by constrained search resulting functional, the non-interacting kinetic energy, reads

In this Appendix we briefly describe how one can con- r 2 ! ˆ d struct the exact Hohenberg-Kohn functional (12) for two Ts(d) = minhΨ|T |Ψi = 2T 1 − 2 1 − . (A.6) Ψ→d 8 electrons in the two-site Hubbard model by carrying out the constrained search as suggested by Levy [25, 26] and The Hartree-exchange-correlation energy then is given by Lieb [27].

The Hilbert space for two fermions on two sites is of EHxc (d) = FHK(d) − Ts(d) (A.7) dimension six and separates into a singlet and a triplet sector of dimension three each. Since for any value of and the corresponding Hxc potential can be easily ob- d, the ground state of Hˆ0 = Tˆ + Uˆ is a singlet, we may tained by differentiation.

[1] M. Marques, N. T. Maitra, F. Nogueira, E. Gross, functional theory, Lecture Notes in Physics, Vol. 837 and A. Rubio, Fundamentals of time-dependent density 8

[13] P. Elliott and N. T. Maitra, Phys. Rev. A 85, 052510 10 (2012). FHK [d] [14] M. Ruggenthaler, S. E. B. Nielsen, and R. van Leeuwen, 8 Phys. Rev. A 88, 022512 (2013). [15] C. Ullrich, Time-Dependent Density-Functional Theory 6 (Oxford University Press, Oxford, 2012). [16] M. Marques and A. Rubio, Phys. Chem. Chem. Phys. [d] 4 U=1 11, 4436 (2009). HK F 2 U=10 [17] “See the special issue of chemical physics, 2011, 391 (1), U=100 guest-edited by R. Baer, L. Kronik, S. K¨ummel.”. 0 [18] J. I. Fuks, N. Helbig, I. V. Tokatly, and A. Rubio, Phys. Rev. B 84, 075107 (2011). -2 [19] J. I. Fuks, P. Elliott, A. Rubio, and N. T. Maitra, J. Phys. Chem. Lett. 4, 735 (2013). -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 [20] M. Thiele, E. K. U. Gross, and S. K¨ummel,Phys. Rev. d Lett. 100, 153004 (2008). [21] J. D. Ramsden and R. W. Godby, Phys. Rev. Lett. 109, 036402 (2012). Figure 5. Exact Hohenberg Kohn density functional FHK[d] [22] D. J. Carrascal and J. Ferrer, Phys. Rev. B 85, 045110 (in units of the hopping parameter T ) for different Hubbard (2012). strength’s U. [23] C. A. Ullrich, J. Chem. Phys. 125, 234108 (2006). [24] C. Verdozzi, Phys. Rev. Lett. 101, 166401 (2008). [25] M. Levy, Proc. Natl. Acad. Sci. USA 76, 6062 (1979). [26] M. Levy, Phys. Rev. A 26, 1200 (1982). (Springer, Heidelberg, 2012). [27] E. Lieb, Int. J. Chem. 24, 243 (1983). [2] M. E. Casida, Journal of Molecular Structure: [28] Y. Li and C. A. Ullrich, J. Chem. Phys. 129, 044105 THEOCHEM 914, 3 (2009). (2008). [3] E. Livshits and R. Baer, Phys. Chem. Chem. Phys. 9, [29] M. Ruggenthaler and D. Bauer, Phys. Rev. Lett. 102, 2932 (2007). 233001 (2009). [4] N. T. Maitra, J. Chem. Phys. 122, 234104 (2005). [30] Since the whole right hand side in this equation is propor- [5] O. Gritsenko and E. J. Baerends, J. Chem. Phys. 121, tional to the time derivative of the denominator, turning 655 (2004). the square root into zero does not lead to a divergence be- [6] D. J. Tozer, J. Chem. Phys. 119, 12697 (2003). cause the numerator vanishes at the same time. Also, for [7] N. T. Maitra, F. Zhang, R. J. Cave, and K. Burke, J. any physical evolution the expression under the square Chem. Phys. 120, 5932 (2004). root always stays nonnegative as |d˙| for a lattice is prop- [8] N. Helbig, J. Fuks, I. Tokatly, H. Appel, E. Gross, and erly bounded from above [10, 24, 33]. A. Rubio, Chem. Phys. 391, 1 (2011). [31] N. T. Maitra, in Time-Dependent Density Functional [9] P. Elliott, S. Goldson, C. Canahui, and N. T. Maitra, Theory, Lecture Notes in physics, Vol. 706 (Springer, Hei- Chem. Phys. 391, 110 (2011). delberg, 2011). [10] M. Farzanehpour and I. V. Tokatly, Phys. Rev. B 86, [32] G. F. Giuliani and G. Vignale, Quantum theory of the 125130 (2012). electron liquid (Cambridge University Press, Cambridge, [11] P. Elliott, J. I. Fuks, A. Rubio, and N. T. Maitra, Phys. 2005). Rev. Lett. 109, 266404 (2012). [33] I. V. Tokatly, Phys. Rev. B 83, 035127 (2011). [12] N. T. Maitra, K. Burke, and C. Woodward, Phys. Rev. Lett. 89, 023002 (2002).