• Keine Ergebnisse gefunden

Radiofrequency Driving of Ultracold Fermions

J Er

= 4

√π V0

Er

34 e−2

V0/Er (2.49)

U Er

= r8

π kaV0

Er

34

. (2.50)

For our simulations in this thesis, we will use a finite chain with open boundary conditions.

The Fermi gas consists of two internal species which can move on a one-dimensional lattice structure and are attractively interacting. The unperturbed Hamiltonian is given by,

H0 =−J XL−1 i=1,σ

(ci,σci+1,σ+h.c.) +U XL

i=1

ni,1ni,2 , (2.51) where c(†)i,σ are the fermionic annihilation (creation) operators of spin σ = {1,2} on site i, and ni,σ the corresponding number operator. J denotes the hopping amplitude, U < 0 the attractive on-site interaction, and L the number of lattice sites. The Fermi-Hubbard model in one dimension can be solved exactlyby Bethe ansatz. We will discuss its ground state phase diagram, together with a brief summary of the Bethe ansatz solution in chapter 3 along with the other numerical and analytical methods used throughout this thesis.

2.5 Radiofrequency Driving of Ultracold Fermions

compared to the atomic cloud, we can also safely assume that the rf-field couples to the en-tire cloud equally strongly, i.e. the Rabi frequency associated with the transition is constant throughout the sample, ΩR = q|E0|h3|r|2i (matrix element of dipole operator which induces transitions between|2iand|3i).

Homogeneous, three-dimensional BCS model for three fermionic species

For the three-state system in chapter 4, by driving Rabi oscillations on the|2i → |3i hy-perfine transition, we effectively break the Cooper pairs and thereby drive the system out of equilibrium. The interaction within the|12i = |1i ⊗ |2imanifold is treated at the mean-field level as outlined before, Eq. 2.22. To add the third state, we include the rf-driving of the form Ω(t) =~ΩRcos(ωrft). Since momentum is conserved in the transfer, we can model the rf-drive asH0(t) = Ω(t)P

k(ck,3ck,2+h.c.). We assume the upper state|3ito be non-interacting, hence the full Hamiltonian describing the system readsH=HBmCSf +H3+H0(t), where

H3 =X

k

k+03−µ3

nk,3 . (2.52)

Upon a unitary transformation|Ψ˜i=e−iγOtˆ |Ψi, the effective Hamiltonian,Heff, governing the dynamics of the transformed state is given by,

i~∂|Ψ˜i

∂t ={e−iγOtˆ HeOtˆ +~γOˆ}|Ψ˜i ≡Heff|Ψ˜i. (2.53) Since, [H, N1] = [H, N2 +N3] = 0, choosing ~γOˆ = [(02 −µ2)−(01 −µ1)]N1, and, in a second transformation~γOˆ =−02N (whereN is the total particle number), allows us to shift the hyperfine levels|1iand|2ion top of each other and move the hyperfine splitting fully into the Hamiltonian describing the upper level. Since initially the upper level is empty, we can take the chemical potential of the upper level to be zero. Furthermore, initially the system is prepared in a balanced mixture of states|1iand|2i, and once their populations are set, the net particle number does not change anymore (within the theoretical modelling; there is no decay channel present). In our numerical simulations we therefore initialise our system in the grand canonical ensemble withµ1 = µ2 = µ(taking the chemical potential into account to set the correct particle number according to Fig. 2.8). During the evolution however, we describe the dynamics in the canonical ensemble and setµ = 0, therefore the chemical potential doesnot explicitly appear in our equations of motion as we will see later in Eq. 3.3.

In order to eliminate the fast oscillating terms in the driving we evoke the rotating-wave approximation, taking the operator γOˆ = −ωrfN3. In this case [ ˆO, HBmCSf ] = [ ˆO, N] = 0 but importantly the driving term doesnotcommute withO. Defining the rf-detuning from theˆ atomic resonance to beδ =ωrf−ωawith~ωa=03−µ302 =03025, we find

5µ3= 0att= 0and remains zero fort >0since we are driving the system far red-detuned from the resonance so the transfer rate to the upper level is very low.

2.5.1 Modelling the Radiofrequency Drive

Heff = HBmCSf +X

k

(k+~ωa−~ωrf)nk,3+~ΩR

2 [erft+e−iωrft]X

k

e−iγtck,3ck,2+eiγtck,2ck,3

≈ HBmCSf (µ= 0) +X

k

(k−~δ)nk,3+ ~ΩR

2 X

k

ck,3ck,2 +ck,2ck,3

, (2.54)

where in the last line we dropped the fast rotating terms∼e±2iωrft6. For notational simplicity we will, in all subsequent chapters, refer to the effective Hamiltonian simply as the Hamilto-nian of the system. This concludes our theoretical modelling of the system. The numerical simulation of the model is discussed in section 3.1.

Three-species Fermi-Hubbard model in a one-dimensional lattice geometry

Figure 2.11:Sketch of the underlying Fermi-Hubbard model with the rf-coupling of the different in-ternal (hyperfine) statesσ ={1,2,3}(here depicted in blue, red and green respectively). The hopping amplitudeJ is taken to be the same in the lower and upper bands, and the hyperfine splitting to the final state is denoted byV3.

Just as in the homogeneous BCS system, the rf-field induces vertical transitions in momen-tum space between the different hyperfine levels of the atom. The difference is the presence of an optical lattice, which breaks the translational symmetry of the system such that ‘real’

momentum is no longer a good quantum number and states are instead labelled by their quasi-momentum living in the first Brillouin zone. The attractive, one-dimensional Fermi-Hubbard model was introduced in section 2.4.2 and is sketched in Fig. 2.11 for clarity. We model the rf-coupling as a perturbationH0(t)to the total Hamiltonian of the system (Eq. 2.51), given by

6This procedure is commonly known as a rotating-wave approximation.

2.5 Radiofrequency Driving of Ultracold Fermions

H0(t) =

z Ω(t)}| {

~Ω23cos(ωrft) XL

i=1

(ci,3ci,2+h.c.)

= Ω(t) XL m=1

(ckm,3ckm,2 +h.c.), (2.55) whereΩ23is the Rabi frequency of the transition,ωrfthe frequency of the rf-field, andkm = L+1 (m = 1, . . . , L) the momentum of the particle, where we have used the Fourier transform for open boundary conditions for numerical convenience. We take the third level of the Hamilto-nian to be a free band,

H3 =−J

L−1X

i=1

(ci,3ci+1,3+h.c.) +V3

XL i=1

ni,3 . (2.56)

This neglects interactions in the final state, which for many atoms can be dominant. However, e.g. for 6Li, any mixture of the lowest three hyperfine states exhibits a broad Feshbach reso-nance. Choosing the|13i=|1i ⊗ |3imanifold as the initial mixture and driving rf-transitions to the initially empty |2i state indeed realises a system with a small scattering length in the final|12i=|1i ⊗ |2istate [115], thus the final state interaction is small and we are mainly in-terested in the dynamics induced by the rf-driving. The energetic splittingV3between the state

|2iand|3iis usually much larger than the kinetic and interaction energy scales, i.e.V3 J, U. Finally, the full model is given byH(t) = H0+H3+H0(t). In chapters 6 and 7 we will present results of the quasi-exact solution of the full, interacting many-body problem H(t)using the time-dependent matrix product state algorithm (see section 3.2 for details).

2.5.2 The non-interacting system

In this subsection, we describe the response of a non-interacting Fermi gas to the rf-drive.

The BCS Hamiltonian (Eq. 2.54) and the Fermi-Hubbard model (Eqs. 2.51, 2.55, 2.56) become diagonal in momentum space in the absence of interactions.

HBCS(t)→X

k

hk(nk,1+nk,2) + (k+~ωa)nk,3+ Ω(t)(ck,3ck,2+h.c.)i

(2.57)

HFH(t)→ XL m=1

h

k(nkm,1+nkm,2) + (k+V3)nkm,3+ Ω(t)(ckm,3ckm,2+h.c.)i

, (2.58) where k = ~2k2/(2m) for the BCS model, and k = −2Jcos(kma) for the Fermi-Hubbard model. We see that the individual momenta fully decouple, and the system can be understood as a series of three-level quantum systems, subject to a periodic drive,

2.5.2 The non-interacting system

H(t) = X

k

Ψk

k 0 0 0 k Ω(t) 0 Ω(t) k+23

Ψk, (2.59)

whereΨk = (ck,1, ck,2, ck,3)7and23 = ~ωa (23 = V3) for the BCS (Fermi-Hubbard) model.

Level|1iis fully decoupled, and the non-trivial dynamics takes place in the two-dimensional {|k,2i,|k,3i} subspace. The transition is ‘vertical’ in momentum space, as exemplary depicted in Fig. 2.12 for the non-interacting Fermi-Hubbard model dispersion.

-1 -0.5 0 0.5 1

ka/π

k V3J, U

σ= 2 σ= 3

Figure 2.12:The lower and upper bands of the non-interacting system in the|23i=|2i⊗|3imanifold.

The two free bands are separated by the hyperfine splitting23=V3 J, U.

Up to overall constant terms, the effective Hamiltonian of this two-level system takes the formH23 =−1223σz+ Ω(t)σx, which we recognise as the Hamiltonian describing a two-level atom driven by a laser field. Within the rotating-wave approximation, the dynamics can be solved analytically8. The drive induces (off)resonant Rabi oscillations given by

hnk,3(t)i= Ω223

2eff sin2(1

2Ωefft), (2.60)

where Ωeff = p

2232 is the generalised, effective Rabi frequency and ~δ = ~ωrf23

the detuning of the rf-field from the bare 2-3 transition. Fig 2.13 shows the evolution of the

7To keep the notation as simple as possible, and for this section only, we will label the momentum ask, to mean eitherk(BCS) orkm(Fermi-Hubbard).

8One can go beyond the rotating wave approximation, applying Floquet theory to this time-periodic problem.

For more details we refer the reader to [116].

2.5 Radiofrequency Driving of Ultracold Fermions

upper level populationhnk,3(t)ifor various detunings. On resonanceδ = 0and the population coherentlyoscillates between the states{|k,2i,|k,3i}. After a timet = π/Ωeff =π/Ω23, for a system initialised in the lower|2ilevel, the populations are completely reversed andhnk,3i= 1 (π-pulse). At finite detuning the Rabi oscillations become faster, Ωeff > Ω23, albeit with a reduced amplitude. The overall amplitude of the oscillations has a Lorentzian dependence on the detuning, where its width is given by the bare Rabi frequencyΩ23of the problem9. Finally, it is important to stress the coherent nature of the transfer. Starting initially in the|k,2istate, the Hamiltonian evolves the system into a coherentsuperpositionof statesψ(t) =α(t)|k,2i+ β(t)|k,3i at time t. As we will see in the later part of this thesis (chapters 6 and 7) adding interactions to the system allows for decoherence and dephasing effects, which reduce or even hinder the coherent transfer of particles between the levels.

0 2π 4π

R

t

0 0.5 1.0

h n

k,3

( t ) i

δ=3ΩR

δ=1ΩR

δ= 0ΩR

δ= 2ΩR

Figure 2.13:Time evolution of the upper level populationhnk,3(t)i. The population undergoes coher-ent Rabi oscillations. Shown are curves for both red (δ <0) and blue (δ >0) detunings. For off-resonant driving,δ6= 0, we observe an increase in the effective modulation frequency accompanied with a reduc-tion in the amplitude of the oscillareduc-tions, in accordance with Eq. 2.60.

9Note that in section 2.5 we introduced the Rabi frequency asR(BCS) and 23 (Fermi-Hubbard). Here we choose the latter notation but the equations hold nonetheless forbothsystems discussed in this thesis.

Chapter 3

Methods

In this chapter we will detail the numerical and analytical methods used throughout this thesis to study the non-equilibrium dynamics of correlated Fermi gases. In section 3.1 we introduce the time-dependent BCS theory and derive its equations of motion. Supplemented by a global self-consistency condition the coupled, non-linear equations of motion are solved with a fourth-order Runge-Kutta method, which allows us to explore the time evolution of the BCS state in chapters 4 and 5. Section 3.2 is devoted to the theoretical background of matrix product state techniques and how they can be used to simulate low-dimensional, interacting quantum many-body systems. We will discuss their range of validity and highlight its algorithmic strength, allowing us access to the quasi-exact time-evolution of driven Hubbard models, as studied in chapters 6 and 7. In these chapters, we study the response of the system to an rf-drive, by mon-itoring the transfer and population of the final state. Within linear response theory, discussed in section 3.3, the transfer rate to the final state can be related to the spectral function of the correlated, initial state, which we use to gain insight into the underlying excitation spectrum of the Fermi-Hubbard model. Our findings are corroborated by exact results from the Bethe ansatz technique, applied to the attractive Fermi-Hubbard model, which is introduced in section 3.4.

While the full theoretical framework of Bethe ansatz technique is not part of this thesis, we will briefly discuss the central ideas and highlight the obtained excitation spectrum, relevant for our analytical understanding of the central results of chapter 6.