Paper

Absorption imaging of a quasi-two-dimensional gas: a multiple scattering analysis

, , , and

Published 1 May 2012 © IOP Publishing and Deutsche Physikalische Gesellschaft
, , Citation L Chomaz et al 2012 New J. Phys. 14 055001DOI 10.1088/1367-2630/14/5/055001

1367-2630/14/5/055001

Abstract

Absorption imaging with quasi-resonant laser light is a commonly used technique for probing ultra-cold atomic gases in various geometries. In this paper, we investigate some non-trivial aspects of this method when applying the method to in situ diagnosis of a quasi-two-dimensional (2D) gas. Using Monte Carlo simulations we study the modification of the absorption cross-section of a photon when it undergoes multiple scattering in the gas. We determine the variations of the optical density with various parameters, such as the detuning of the light from the atomic resonance and the thickness of the gas. We compare our results to the known 3D result (the Beer–Lambert law) and outline the specific features of the 2D case.

Export citation and abstractBibTeXRIS

1. Introduction

The study of cold atomic gases has recently shed new light on several aspects of quantum many-body physics [14]. Most of the measurements in this field of research are based on the determination of the spatial density of the gas [5]. For instance, one can use the in situ steady-state atomic distribution in a trapping potential to infer the equation of state of the homogeneous gas [6]. Another example is the time-of-flight method, in which one measures the spatial density after switching off the trapping potential and allowing for a certain time of ballistic expansion. This gives access to the momentum distribution of the gas, and to the conversion of interaction energy into kinetic energy at the moment of the potential switch-off.

To access the atomic density n(r), one usually relies on the interaction of the atoms with quasi-resonant laser light. The most common method is absorption imaging, in which the shadow imprinted by the cloud on a low-intensity probe beam is imaged on a camera. The simplest modelling of absorption imaging is based on a mean-field approach, in which one assumes that the local value of the electric field driving an atomic dipole at a given location depends only on the average density of scatterers. One can then relate the attenuation of the laser beam to the column atomic density along the line-of-sight z. The optical density of the cloud is given by the Beer–Lambert law

where σ is the photon scattering cross-section and Iin (resp. Iout) are the incoming (resp. outgoing) intensity of the probe laser in the plane xy perpendicular to the propagation axis. For a closed two-level atomic transition of frequency ω0 = ck0, σ depends on the wavelength λ0 = 2π/k0 associated with this transition and on the detuning Δ = ω − ω0 between the probe light frequency ω and the atomic frequency:

Here Γ represents the natural linewidth of the transition (i.e. Γ−1 is the natural lifetime of the excited state of the transition). Equation (2) assumes that the intensity of the probe beam is much lower than the saturation intensity of the atomic transition. Quasi-resonant absorption imaging is widely used to measure the spatial distribution of atomic gases after a long time- of-flight, when the density has dropped sufficiently so that the mean-field approximation leading to equation (1) is valid.

One can also use absorption imaging to probe in situ samples, at least in the case when σ n(col) is not very large so that the output intensity is not vanishingly small. This is particularly the case with low-dimensional gases. Consider for example a 2D gas, such that the translational degree of freedom along z has been frozen. For a probe beam propagating along this axis, one can transpose the Beer–Lambert law of equation (1) by simply replacing the column density by the surface density n(2D) of the gas. This 2D Beer–Lambert law can be heuristically justified by treating each atom as a disc of area σ that blocks every photon incident on it. In an area A ≫ σ containing N = An(2D) ≫ 1 randomly placed atoms, the probability that a photon is not blocked by any of the discs is .

In a quasi-2D gas there is, however, an important limitation on the optical densities to which one may apply the Beer–Lambert prediction of equation (1). Already for σ0n(2D) = 1 the mean inter-particle distance is only 0.7 λ0 and one may expect that the optical response of an atom strongly depends on the precise location of its neighbours. More precisely, the exchange of photons between closely spaced atoms induces a resonant van der Waals interaction that significantly shifts the atomic resonance frequency with respect to its bare value ω0. The optical density of the gas at resonance may then be reduced with respect to equation (1), and this was indeed observed in a series of experiments performed with a degenerate 87Rb gas [7, 8].

The general subject of the propagation of a light wave in a dense atomic sample, where multiple scattering plays an essential role, has been the subject of numerous experimental and theoretical works (see, e.g., [9, 10] in the context of cold atoms and [11] for a review). Here we present a quantitative treatment of the collective effects that appear when a weak probe beam interacts with a quasi-2D atomic gas. We consider an ensemble of N atoms at rest with random positions and investigate the transmission of quasi-resonant light by the atom sheet. We model the resonance transition between the atomic ground (g) and excited (e) states by a transition. We present two equivalent approaches; the first one is based on the calculation of the field radiated by an assembly of N dipoles, where each dipole is driven by an external field plus the field radiated by the N − 1 other dipoles; the second one uses the standard T matrix formalism of scattering theory. We show that in both cases the optical density of the medium can be determined by solving the same 3N × 3N linear system. A similar formalism has been previously used for the study of light propagation in small 3D atomic samples, in the presence of multiple scattering (see, e.g., [1220]). However, its application to quasi-2D samples has (to our knowledge) not yet been investigated, except in the context of the Anderson localization of light [14] (the 1D case was addressed in [21]). Our numerical calculations are performed for N = 2048 atoms, which is sufficient to reach the 'thermodynamic limit' for the range of parameters that is relevant for experiments. We show, in particular, that even for moderate values of σ0n(col), the optical density is notably reduced compared to what is expected from the Beer–Lambert law (e.g. more than 20% reduction for σ0n(col) = 1). We investigate how the absorption line shape is modified by resonant van der Waals interactions and also show how the result (1) is recovered when one increases the thickness of the gas, for a given column density n(col). Note that our simulations are performed with a much smaller atom number than in real experiments (N is up to 105 in [7, 8]). However, we expect our results to be relevant for addressing experimental realizations, since the light attenuation can be treated in the local density approximation. Indeed in the 'thermodynamic limit' mentioned above, the optical density at a point (x,y) depends only on the spatial density n(x,y,z) along the line of sight and not on the overall density distribution n(x',y',z') in the sample.

The paper is organized as follows. In section 2, we detail the modelling of the atom–light interaction with the two-level approximation and rotating wave approximation (RWA). Then we explain the principle of the calculation of the absorption of a weak probe beam crossing the atom slab (section 3). The ensemble of our numerical results are presented in section 4. Finally, in section 5 we discuss some limitations of our model and present our concluding remarks.

2. Modelling the atom–light interaction

2.1. The electromagnetic field

We use the standard description of the quantized electromagnetic field in the Coulomb gauge [22], and choose periodic boundary conditions in the cubic-shaped quantization volume . We denote by aq,s the destruction operator of a photon with wave vector q and polarization s (sq). The Hamiltonian of the quantized field is

and the transverse electric field operator reads E(r) = E(+)(r) + E(−)(r) with

and E(−)(r) = (E(+)(r)). The wave vectors q are quantized in the volume as qi = 2πni/Li, i = x,y,z, where ni is a positive or negative integer.

2.2. The atomic medium

We consider a collection of N identical atoms at rest in positions rj, j = 1,...,N. We model the atomic resonance transition by a two-level system with a ground state |g〉 with angular momentum Jg = 0 and an excited level of angular momentum Je = 1. We choose as a basis set for the excited manifold the three Zeeman sublevels |eα〉, α = x,y,z, where |eα〉 is the eigenstate with eigenvalue 0 of the component Jα of the atomic angular momentum operator. We denote by ℏω0 the energy difference between e and g. The atomic Hamiltonian is thus (up to a constant)

The restriction to a two-level approximation is legitimate if the detuning Δ between the probe and the atomic frequencies is much smaller than ω0. The modelling of this transition by a transition leads to a relatively simple algebra. The transitions that are used for absorption imaging in real experiments often involve more Zeeman states ( for Rb atoms in [7, 8]), but are more complex to handle [23, 24], and they are thus beyond the scope of this paper. However, we believe that the most salient features of multiple scattering and resonant van der Waals interactions are captured by our simple level scheme.

2.3. The atom–light coupling

We treat the atom–light interaction using the electric dipole approximation (length gauge), which is legitimate since the resonance wavelength of the atoms λ0 is much larger than the atomic size. We write the atom–light coupling as

where Dj is the dipole operator for the atom j. We will use the RWA, which consists in keeping only the resonant terms in the coupling:

where h.c. stands for the Hermitian conjugate. Here D(+)j represents the raising part of the dipole operator for atom j:

where d is the electric dipole associated with the ge transition and is a unit vector in the direction α.

When a single atom is coupled to the electromagnetic field, this coupling results in the modification of the resonance frequency (the Lamb shift) and in the fact that the excited state e acquires a non-zero width Γ

For simplicity we will incorporate the Lamb shift into the definition of ω0. Note that the proper calculation for this shift requires that one goes beyond the two-level and the RWAs. The linewidth Γ, on the other hand, can be calculated from the above expressions for V using the Fermi golden rule.

The RWA provides a very significant simplification of the treatment of the atom–light coupling, in the sense that the total number of excitations is a conserved quantity. The annihilation (resp. creation) of a photon is always associated with the transition of one of the N atoms from g to e (resp. from e to g). This would not be the case if the non-resonant terms of the electric dipole coupling D(+)i·E(−) and D(−)i·E(+) were also taken into account. For a single atom, the small parameter associated with the RWA is Δ/ω0, which is, in practice, in the range 10−6–10−9. For a collection of atoms, the RWA is still an excellent approximation if the shifts introduced by resonant van der Waals interactions are small compared to atomic frequencies. In practice, this requires that the interatomic distances remain very large compared to the atom size (the Bohr radius).

Formally, the use of the electric dipole interaction implies to add to the Hamiltonian an additional contact term between the dipoles (see, e.g., [12, 25]). This term will play no role in our numerical simulations because we will surround the position of each atom by a small excluded volume, which mimics the short-range repulsive interaction between atoms. We checked that the results of our numerical calculations (see section 4) do not depend on the size of the excluded volume, and we can safely omit the additional contact term in this work.

3. Interaction of a probe laser beam with a dense quasi-two-dimensional (2D) atomic sample

We present in this section the general formalism that allows one to calculate the absorption of a quasi-resonant laser beam by a slab of N atoms. We address this question using two different approaches. The first one maps the problem onto the collective behaviour of an assembly of N oscillating dipoles [12]. The equation of motion for each dipole is obtained using the Heisenberg picture for the Hamiltonian presented in section 2. It contains two driving terms, one from the incident probe field and the other from the field radiated by all the other dipoles at the location of the dipole under study. The steady state of this assembly of dipoles is obtained by solving a set of 3N linear equations. The second approach uses the standard quantum scattering theory [26], which is well suited for perturbative calculations and partial resummations of diagrams. We suppose that one photon is incident on the atomic medium and use resummation techniques to take into account the multiple scattering events that can occur before the photon emerges from the medium. The relevant quantity in this approach is the probability amplitude Tii that the outgoing photon is detected in the same mode as the incident one [15, 18], and we show that Tii is obtained from the same set of equations as the values of the dipoles in the first approach.

3.1. Wave propagation in an assembly of driven dipoles

In this section, we assume that the incident field is prepared in a coherent state corresponding to a monochromatic plane wave EL epsilon ei(kzωt). We choose the polarization epsilon to be linear and parallel to the x-axis (). Since we consider a transition, this choice does not play a significant role and we checked that we recover essentially the same results with a circular polarization. Note that the situation would be different for an atomic transition with larger Jg and Je since optical pumping processes would then depend crucially on the polarization of the probe laser.

The amplitude EL is supposed to be small enough that the steady-state populations of the excited states ej,α are small compared to unity. This ensures that the response of each atomic dipole is linear in EL; this approximation is valid when the Rabi frequency dEL/ℏ is small compared to the natural width Γ or the detuning Δ.

Using the atom–light coupling (6), the equations of motion for the annihilation operators aq,s in the Heisenberg picture read

This equation can be integrated between the initial time t0 and the time t, and the result can be injected in the expression for the transverse field to provide its value at any point r:

where Efree stands for the value obtained in the absence of atoms. We now take the quantum average of this set of equations. In the steady-state regime the expectation value of the dipole operator Dj(t) can be written as dj e−iωt + c.c., and the average of Efree(r,t) is the incident field EL epsilon ei(kzωt) + c.c. We denote the average value of the transverse field operator in r as , and we obtain after some algebra (see, e.g., [12, 27])

where we set uj = k(r − rj) (with k ≈ k0),

and

The function gα,α'(kr) is identical to the one appearing in classical electrodynamics [28], when calculating the field radiated in r by a dipole located at the origin.

We proceed similarly for the equations of motion for the dipole operators D(−)j and take their average value in steady state. The result can be put in the form [12]

where the reduced detuning δ = 2Δ/Γ has been defined in equation (2) and uj,j' = k(rj' − rj). This can be written with matrix notation

where the 3N vectors |X〉 and |Y 〉 are defined by

and where the complex symmetric matrix [M] has its diagonal coefficients equal to δ + i and its off-diagonal coefficients (for j ≠ j') given by gα,α'(ujj'). This matrix belongs to the general class of Euclidean matrices [29], for which the (i,j) element can be written as a function F(ri,rj) of points ri in the Euclidean space. The spectral properties of these matrices for a random distribution of the ri's (as will be the case in this work, see section 4) have been studied in [2932].

Equation (15) has a simple physical interpretation: in the steady state each dipole dj is driven by the sum of the incident field EL and the field radiated by all the other dipoles. This set of 3N equations was first introduced by L L Foldy in [33] who named it, together with equation (12), 'the fundamental equations of multiple scattering'. Indeed for a given incident field, the solution of (16) provides the value of each dipole dj, which can then be injected in (12) to obtain the value of the total field at any point in space.

3.2. Absorption signal

From the expression of the average value of the dipoles, we now extract the absorption coefficient of the probe beam and the optical density of the gas. We suppose that the N atoms are uniformly spread in a cylinder of radius R along the z-axis and located between z = −ℓz/2 and z = ℓz/2. We can consider two experimental setups to address this problem. The first one, presented in figure 1(a), consists in measuring after the atomic sample the total light intensity with the same momentum as the incident probe beam. This can be achieved by placing a lens with the same size as the atomic sample, in the plane z = ℓ' > ℓz/2 just after the sample. The light field at the focal point of the lens F gives the desired attenuation coefficient. We refer to this method as 'global', since the field E(F) provides information over the whole atomic cloud. One can also use the setup sketched in figure 1(b), which forms an image of the atom slab on a camera and provides a 'local' measurement of the absorption coefficient. In real experiments local measurements are often favoured because trapped atomic sample are non-homogeneous and it is desirable to access the spatial distribution of the particles. However, for our geometry with a uniform density of scatterers, spatial information on the absorption of the probe beam is not relevant. Therefore we only present the formalism for global measurements, which is simpler to derive and leads to slightly more general expressions. We checked numerically that we obtained very similar results when we modelled the local procedure.

Figure 1. Refer to the following caption and surrounding text.

Figure 1. Two possible setups for measuring the absorption of an incident probe beam by a slab of atoms using a lens of focal f. (a) Global probe. (b) Local probe.

Standard image

We assume that the lens in figure 1(a) operates in the paraxial regime, i.e. its focal length f is much larger than its radius R. We relate the field at the image focal point of the lens to the field in the plane z = ℓ' just before the lens:

where the integral runs over the lens area. Since the incident probe beam is supposed to be linearly polarized along x, we calculate the x component of the field in F. Plugging the value of the field given in equations (12) and (17), we obtain the transmission coefficient

This result can be simplified in the limit of a large lens by using an approximated value for the integral appearing in (19). We suppose that kℓ' ≫ 1 so that the dominant part in gx,α is the eiu/u contribution to h1. More precisely, the domain in the lens plane contributing to the integral for the dipole j is essentially a disc of radius centred on (xj,yj). When this small disc is entirely included in the lens aperture, i.e. the larger disc of radius R centred on x = y = 0, we obtain

We use the result (20) for all atoms, which amounts to neglecting edge effects for the dipoles located at the border of the lens, and we obtain

with n(col) = N/πR2 and where the coefficient Π is defined by

This coefficient captures the whole physics of multiple scattering and resonant van der Waals interactions among the N atoms. Indeed one takes into account all possible couplings between the dipoles when solving the 3N × 3N system [M]|X〉 = |Y 〉. Once is known the optical density is obtained from

As an example, consider the limit of a very sparse sample where multiple scattering does not play a significant role (σ0n(col) ≪ 1). All non-diagonal matrix elements in [M] are then negligible and [M] is simply the identity matrix, times i + δ. Each Xj,x solution of the system (16) is equal to eikzj/(i + δ), and we obtain as expected

3.3. Light absorption as a quantum scattering process

In order to study the attenuation of a weak probe beam propagating along the z-axis when it crosses the atomic medium, we can also use quantum scattering theory. The Hamiltonian of the problem is

and we consider the initial state where all atoms are in their ground state and where a single photon of wave vector and polarization is incident on the atomic medium

with . The state |Ψi〉 is an eigenstate of H0 with energy ℏω. The interaction of the photon with the atomic medium, described by the coupling V , can be viewed as a collision process during which an arbitrary number of elementary scattering events can take place. Each event starts from a state and corresponds to

  • (i)  
    The absorption of the photon in mode q,s by atom j, which jumps from its ground state |j:g〉 to one of its excited states |j:eα〉. The state of the system is then
    where |vac〉 stands for the vacuum state of the electromagnetic field. The subspace spanned by the states has dimension 3N.
  • (ii)  
    The emission of a photon in the mode (q',s') by atom j, which falls back into its ground state.

Finally, a photon emerges from the atomic sample, and we want to determine the probability amplitude to find this photon in the same mode |k,epsilon〉 as the initial one.

The T matrix defined as

where 0+ is a small positive number that tends to zero at the end of the calculation, provides a convenient tool to calculate this probability amplitude. Generally,

gives the probability amplitude to find the system in the final state |Ψf〉 after the scattering process. The states |Ψi〉 and |Ψf〉 are eigenstates of the unperturbed Hamiltonian H0, with energy Ei. Here we are interested in the element Tii of the T matrix, corresponding to the choice |Ψf〉 = |Ψi〉. Using the definition (28), we find

We now have to calculate the (3N× (3N) matrix elements of the operator 1/(z − H), with z = ℏω + i0+, entering into (30). We introduce the two orthogonal projectors P and Q, where P projects on the subspace with zero photon, and Q projects on the orthogonal subspace. We thus have

We define the displacement operator

and use the general result [22]

where the effective Hamiltonian Heff is

For the following calculations, it is convenient to introduce the dimensionless matrix [M] proportional to the denominator of the right-hand side of (34):

It is straightforward to check5 that for z → ℏω this matrix coincides with the symmetric matrix appearing in (16). Indeed the matrix elements of R(z) are

which can be calculated explicitly. For j = j', the real part of this expression is the Lamb shift that we reincorporate in the definition of ω0, and its imaginary part reads

For j ≠ j', the sum over (q,s) appearing in (37) is the propagator of a photon from an atom in rj in internal state |eα〉, to another atom in rj' in internal state |eα'〉. This is nothing but (up to a multiplicative coefficient) the expression that we have already introduced for the field radiated in rj' by a dipole located in rj:

where the tensor gα,α' is defined in equations (13) and (14).

Suppose now that the atoms are uniformly distributed over the transverse area LxLy of the quantization volume. We set n(col) = N/(LxLy) and we rewrite the expression (30) of the desired matrix element Tii as

where the coefficient Π has been defined in (22). The result (40) combined with (21) leads to

which constitutes the 'optical theorem' for our slab geometry, since it relates the attenuation of the probe beam to the forward scattering amplitude Tii.

The emergence of resonant van der Waals interactions is straightforward in this approach. Let us consider for simplicity the case when only N = 2 atoms are present. The effective Hamiltonian Heff is a 6 × 6 matrix that can be easily diagonalized and its eigenvectors, with one atom in |e〉 and one in |g〉, form in this particular case an orthogonal basis, although Heff is a non-Hermitian [34, 35]. For a short distance r between the atoms (kr ≪ 1), the leading term in h1(u) and h2(u) is u−3 and the energies (real parts of the eigenvalues) of the six eigenstates vary as ∼ ± ℏΓ/(kr)3 (resonant dipole–dipole interaction). The imaginary parts of the eigenvalues, which give the inverse of the radiative lifetime of the states, tend either to Γ or 0 when r → 0, which correspond to the superradiant and subradiant states for a pair of atoms, respectively [36].

For N > 2 the eigenvectors of the non-Hermitian Euclidean matrix Heff are, in general, non-orthogonal, which complicates the use of standard techniques of spectral theory in this context [31, 32]. More precisely, one could think of solving the linear system (16), or equivalently calculating Tii in equation (30), by using the expansion of the column vector |Y 〉 defined in equation (17) on the left (|αj〉) and right (〈βj|) eigenvectors of Heff. Then one could insert this expansion into the general expression of the matrix element Tii, to express it as a sum of the contributions of the various eigenvalues of Heff. However, the physical discussion based on this approach is made difficult by the fact that since Heff is non-Hermitian, the {|αj〉} and the {|βj〉} bases do not coincide. Hence the weight 〈βj|Y 〉〈Y |αj〉 of a given eigenvalue in the sum providing the value of Tii is not a positive number, and this complicates the interpretation of the result.

As a final comment, we note that our approach does not make use of the quite common concept of self-energy. In the context of the propagation of waves in a random medium [11], the self-energy Σ appears by averaging the relation G = G0 + G0V G over disorder, where G = 1/(z − H) and G0 = 1/(z − H0). The average of G then satisfies , where Σ is an infinite sum of irreducible diagrams. Here we choose a different but equivalent route, which consists in solving exactly the scattering problem for N atoms and 1 photon for a given realization of the disorder, i.e. a given draw {rj} of the positions of the atoms, obtain the attenuation of the sample, and subsequently average this result over several realizations of the disorder (see section 4).

3.4. Beyond the sparse sample case: 3D versus 2D

For a sparse sample, we have already calculated the optical density at first order in density (equation (24)) and the result is identical for a strictly 2D gas and a thick one. The approach based on quantum scattering theory is well suited to go beyond this first-order approximation and look for differences between the 2D and 3D cases. The basis of the calculation is the series expansion of equation (34), which gives

Consider the case of a resonant probe δ = 0 for simplicity. The result obtained for a sparse sample in equation (24) corresponds to the first term [P/(z − H0)] of this expansion. Here we investigate the next order term and explain why one can still recover the Beer–Lambert law for a thick (3D) gas, but not for a 2D sample.

Double scattering diagrams for a thick sample (kz ≫ 1). We start our study by adding the first term (n = 1) in the expansion (42) to the zeroth-order term already taken into account in equation (24). This amounts to taking into account the diagrams where the incident photon is scattered on a single atom and those where the photon 'bounces' on two atoms before leaving the atomic sample. Inserting the first two terms of the expansion (42) into (40), we obtain

We now have to average this result on the positions of the atoms j and j'. There are N(N − 1) ≈ N2 couples (j,j'). Assuming that the gas is dilute so that the average distance between two atoms (in particular |zj − zj'|) is much larger than k−1, the leading term in gxx is the eiu/u contribution of h1(u) in equations (13) and (14). We thus arrive at

where the average is taken over the positions r and r' of two atoms. We first calculate the average over the xy coordinates and we obtain (cf equation (20))

For a thick gas (kz ≫ 1) the bracket in this expression has an average value of ≈1/2. Indeed the function to be averaged is equal to 1 if z < z', which occurs in half the cases, and it oscillates and averages to zero in the other half of the cases, where z > z'. We thus obtain the approximate value of the transmission coefficient:

where we recognize the first three terms of the power series expansion of , corresponding to the optical density .

Double scattering diagrams for a 2D gas (ℓz = 0). When all atoms are sitting in the same plane, the evaluation of the second-order term (and the subsequent ones) in the expansion of Tii in powers of the density is modified with respect to the 3D case. The calculation starts as above and the second term in the bracket of equation (43) can now be written as

If we keep only the terms varying as eiu/u in h1 and h2 (equation (14)), we can calculate analytically the integral in (47) and find the same result as in 3D, i.e. iσ0n(2D)/4. If this was the only contribution to (47), it would lead to the Beer–Lambert law also in 2D, at least at second order in density. However, one can check that a significant contribution to the integral in (47) comes from the region u = kr < 1. In this region, it is not legitimate to keep only the term in eikr/kr in h1, h2, since the terms in eikr/(kr)3, corresponding to the short-range resonant van der Waals interaction, are actually dominant. Therefore the expansion of the transmission coefficient in powers of the density differs from (46), and one cannot recover the Beer–Lambert law at second order in density. Calculating analytically corrections to this law could be done following the procedure of [12]. Here we will use a numerical method to determine the deviation with respect to the Beer–Lambert law (see section 4.2).

Remark 1. For a 3D gas there are also corrections to the second term in equation (45) due the 1/r3 contributions to h1 and h2. However these corrections have a different scaling with the density and can be made negligible. More precisely their order of magnitude is ∼n(3D)k−3, to be compared with the value ∼n(col)k−2 of the second term in equation (45). Therefore, one can have simultaneously n(3D)k−3 ≪ 1 and n(col)k−2 ≳ 1 if the thickness ℓz of the gas along z is ≫1/k.

4. Absorption of light by a slab of atoms

In order to study quantitatively the optical response of a quasi-2D gas, we have performed a Monte Carlo calculation of the transmission factor given in equation (21) and of the related optical density . We start our calculation by randomly drawing the positions of N atoms, then we solve numerically the 3N × 3N linear system (16) and, finally, we insert the result for the N dipoles into the expression of .

The atoms are uniformly distributed in a cylinder of axis z, with a radius R and a thickness ℓz. The largest spatial densities considered in this work correspond to a mean inter-particle distance ≈k−1. Around each atom we choose a small excluded volume with a linear size a = 0.01 k−1. We varied a by a factor of 10 around this value and checked that our results were essentially unchanged. Apart from this excluded volume we do not include any correlation between the positions of the atoms. This choice is justified physically by the fact that, in the case of large phase space densities which motivates our study, the density fluctuations in a 2D Bose gas are strongly reduced and the two-body correlation function g2(r,r') is such that g2(r,r) ≈ 1 [37].

In this section, we first determine the value of N that is needed to reach the 'thermodynamic limit' for our problem: for a given thickness ℓz, should not be an independent function of the number of atoms N and the disc radius R, but should depend only on the ratio N/πR2 = n(col). We will see that this imposes to use a relatively large number of atoms, typically N > 1000, for the largest spatial densities considered here. All subsequent calculations are performed with N = 2048. We then study the dependence of on the various parameters of the problem: the column density n(col), the thickness of the gas ℓz and the detuning Δ. In particular, we show that for a given n(col) we recover the 3D result (1) when the thickness ℓz is chosen sufficiently large.

4.1. Reaching the 'thermodynamic limit'

We start our study by testing the minimal atom number that is necessary to obtain a faithful estimate of the optical density. We choose a given value of n(col) = N/πR2 and we investigate how depends on N either for a strictly 2D gas (ℓz = 0) or for a gas extending significantly along the third direction (ℓz = 20 k−1). We consider a resonant probe for this study (Δ = 0). We vary N by multiplicative steps of 2, from N = 8 up to N = 2048 and we determine how large N must be so that is a function of n(col) only.

The results are shown in figures 2(a) and (b), where we plot as a function of N. We conduct this study for four values of the density n(col), corresponding to σ0n(col) = 0.5,1,2 and 4. Let us consider first the smallest value σ0n(col) = 0.5. For each value of N we perform a number of draws that is sufficient to bring the standard error below 2 × 10−3 and we find that the calculated optical density is independent of N (within standard error) already for N ≳ 100, for both values of ℓz. Consider now our largest value σ0n(col) = 4; for a strictly 2D gas (ℓz = 0), reaches an approximately constant value independent of N for N ≳ 1000. For σ0n(col) = 4 and a relatively thick gas (ℓz = 20 k−1, blue filled circles in figure 2(b)), reaching the thermodynamic limit is more problematic since there is still a clear difference between the results obtained with 1024 and 2048 atoms. This situation thus corresponds to the limit of validity of our numerical results. In the rest of this paper, we will show only results obtained with N = 2048 atoms for column densities not exceeding σ0n(col) = 4. The number of independent draws of the atomic positions (at least 8) is chosen such that the standard error for each data point is below 2%.

Figure 2. Refer to the following caption and surrounding text.

Figure 2. Variation of the optical density calculated from (21) as a function of the number of atoms N, for ℓz = 0 (a) and ℓz = 20 k−1 (b) and for four values of the density: σ0n(2D) = 0.5 (black empty squares), 1 (red empty circles), 2 (green diamonds) and 4 (blue filled circles). The bars indicate the standard deviations. The dotted lines give the value obtained for our largest value of N (N = 2048). The results have been obtained at resonance (Δ = 0).

Standard image

4.2. Measured optical density versus the Beer–Lambert prediction

We now investigate the variation of the optical density as a function of the column density of the sample n(col), or equivalently of the Beer–Lambert prediction . We suppose in this section that the probe beam is resonant (Δ = 0), and we address the cases of a strictly 2D gas (ℓz = 0) and a thick slab (ℓz = 20 k−1).

Consider first the case of a strictly 2D case, ℓz = 0, leading to the results shown in figure 3(a). We see that differs significantly (∼25%) from already for around 1. A quadratic fit to the calculated variation of for σ0n(2D) < 1 (continuous red line) gives

The discrepancy between and increases when the density increases: for , the calculated is only ≈1.4. For such a large density the average distance between nearest neighbours is ≈k−1 and the energy shifts due to the dipole–dipole interactions are comparable to or larger than the linewidth Γ. The atomic medium is then much less opaque to a resonant probe beam than in the absence of dipole–dipole coupling.

Figure 3. Refer to the following caption and surrounding text.

Figure 3. Variations of the optical density as function of the Beer–Lambert prediction for ℓz = 0 (a) and ℓz = 20 k−1 (b). The black dotted line is the straight line of slope 1. In (a) the continuous red line is a quadratic fit with μ = 0.22 to the data points with . The calculations are done for N = 2048, Δ = 0 and the bars indicate standard deviations.

Standard image

Consider now the case of a thick sample, ℓz = 20 k−1 (figure 3(b)). The calculated optical density is then very close to the Beer–Lambert prediction over the whole range we studied. This means that in our chosen range of optical densities, the mean-field approximation leading to is satisfactory if the sample thickness exceeds a few optical wavelengths λ = 2π/k.

It is interesting to characterize how the optical density evolves from the value for a strictly 2D gas to the expected value from the Beer–Lambert law as the thickness of the gas increases. We show in figure 4 the variation of as a function of ℓz for three values of the column density corresponding to and 4. An exponential fit to these data for 2 k−1 ⩽ ℓz ⩽ 20 k−1 gives a good account of the variation observed over this range, and provides the characteristic thickness ℓc needed to recover the Beer–Lambert law. We find that ℓc ≈ 3.0 k−1 for , ℓc ≈ 3.5 k−1 for and ℓc ≈ 4.4 k−1 for .

Figure 4. Refer to the following caption and surrounding text.

Figure 4. Variation of with the thickness ℓz of the gas for various column densities corresponding to (red squares), 2 (green circles), 4 (blue diamonds). The continuous lines are exponential fits to the data. The dotted lines give the Beer–Lambert result. The calculations are done for N = 2048, Δ = 0, and the bars indicate standard deviations.

Standard image

Remark 2. For the largest value of column density considered here (n(col)σ0 = 4), we find that increases slightly above the value of when ℓz is chosen larger than 20 k−1 (the upper value considered in figure 4). We believe that this is a consequence of the edge terms that we neglected when approximating equation (19) by (21). These terms become significant for , because for our atom number N = 2048, the sample radius R ≈ 55 k−1 is then not very large compared to its thickness for ℓz ≳ 20 k−1. In order to check this assumption, we also calculated numerically the result of equation (19) (instead of equation (21)) for practical values of the parameters (position and radius) of the lens represented in figure 1(a). The results give again , but now with remaining below . Since our emphasis in this paper is rather placed on the 2D case, we will not explore this aspect further here.

4.3. Absorption line shape

Resonant van der Waals interactions manifest themselves not only in the reduction of the optical density at resonance, but also in the overall line shape of the absorption profile. To investigate this problem we have studied the variations of with the detuning of the probe laser. We show in figure 5(a) the results for a strictly 2D gas (ℓz = 0) for n(col)σ0 = 1,2 and 4. Several features show up in this series of plots. First we note a blue shift of the resonance, which increases with n(2D) and reaches Δ ≈ Γ/4 for σ0n(2D) = 4. We also note a slight broadening of the central part of the absorption line, since the full-width at half-maximum, which is equal to Γ for an isolated atom, is ≃1.3Γ for n(col)σ0 = 4. Finally, we note the emergence of large, non-symmetric wings in the absorption profile. This asymmetry is made more visible in figure 5(b), where we show with full blue squares the same data as in figure 5(a) for n(col)σ0 = 4, but now plotting as a function of δ. For a detuning δ = ± 15, the calculated optical density exceeds the Beer–Lambert prediction by a factor of 4.1 (resp. 2.8) on the red (resp. blue) side.

Figure 5. Refer to the following caption and surrounding text.

Figure 5. (a) Variation of with the reduced detuning δ = 2Δ/Γ of the probe laser in the case of a 2D gas (ℓz = 0), for three values of the column density σ0n(col) = 1 (red squares), 2 (green circles), and 4 (blue diamonds). (b) Blue full squares: same data as in (a), now plotted for as function of δ. Blue open squares: for a thick gas (ℓz = 20 k−1). Black stars: for a 2D gas (ℓz = 0) and a large exclusion region around each atom (a = k−1). All data in (b) correspond to σ0n(2D) = 4. The calculations are done with N = 2048 atoms and the bars indicate standard deviations.

Standard image

In order to gain a better understanding of these various features, we give in figure 5(b) two additional results. On the one hand we plot with empty blue squares the variations of for a thick gas (ℓz = 20 k−1) with the same column density n(col)σ0 = 4. There are still some differences between and in this case, as already pointed out in [18], but they are much smaller than in the ℓz = 0 case. This indicates that the strong deviations with respect to the Beer–Lambert law that we observe in figure 5(a) are specific 2D features. On the other hand, we plot with black stars the variations of for a 2D gas (ℓz = 0) in which we artificially increased the exclusion radius around each atom up to a = k−1 instead of a = 0.01 k−1 (blue full squares) for the other results in this paper. This procedure, which was suggested to us by Robin Kaiser, allows one to discriminate between effects due to isolated pairs of closely spaced atoms, and many-body features resulting from multiple scattering of photons among larger clusters of atoms. The comparison of the results obtained for a = 0.01 k−1 and a = k−1 suggests that the blue shift of the resonance line, which is present in both cases, is a many-body phenomenon, whereas the large-amplitude wings with a blue–red asymmetry, which occurs only for a = 0.01 k−1, is rather an effect of close pairs.

This asymmetry in the wings of the absorption line in a 2D gas can actually be understood in a semi-quantitative manner by a simple reasoning. We recall that for two atoms at a distance r ≪ k−1, the levels involving one ground and one excited atom have an energy (real part of the eigenvalues of Heff) that is displaced by ∼ ± ℏΓ/(kr)3. A given detuning δ can thus be associated with a distance r between the two members of a pair that will resonantly absorb the light. To be more specific let us consider a pair of atoms with kr ≪ 1, and suppose for simplicity that it is aligned either along the polarization axis of the light (x) or perpendicularly to the axis (y). In both cases the excited state of the pair that is coupled to the laser is the symmetric combination . If the pair is aligned along the x-axis, this state has an energy ℏω0 − 3ℏΓ/2(kr)3; hence it is resonant with red detuned light such that δ = −3/(kr)3. If the pair axis is perpendicular to x, the state written above has an energy ℏω0 + 3ℏΓ/4(kr)3; hence it is resonant with blue detuned light such that δ = 3/2(kr)3. This clearly leads to an asymmetry between red and blue detuning; indeed the pair distance r needed for ensuring resonance for a given δ > 0, rblue = (3/2|δ|)1/3 k−1, is smaller than the value rred = (3/|δ|)1/3 k−1 for the opposite value −δ. Since the probability density for the pair distance at small r is in 2D for randomly drawn positions, we expect the absorption signal to be stronger for −δ than for +δ. In a 3D geometry the variation of the probability density at small r is even stronger ( ), but it is compensated by the fact that the probability of occurrence of pairs that are resonant with blue detuned light is dimensionally increased. For example, in our simplified modelling where the pair axis is aligned with the reference axes, a given pair will be resonant with blue detuned light in 2/3 of the cases (axis along y or z) and resonant with red detuned light only in 1/3 of the cases (axis along x). This explains why the asymmetry of the absorption profile is much reduced for a 3D gas in comparison to the 2D case.

5. Summary and outlook

We have presented in this paper a detailed analysis of the scattering of light by a disordered distribution of atoms in a quasi-2D geometry. The particles were treated as fixed scatterers and their internal structure was modelled as a two-level system, with a J = 0 ground state and a J = 1 excited state. In spite of these simplifying assumptions the general trend of our results is in good agreement with the experimental finding of [7], where a variation of the measured optical density similar to that of figure 3 was measured.

Several improvements to our modelling can be considered in order to achieve quantitative agreement with theory and experiment. The first one is to include the relatively complex atomic structure of the alkali-metal species used in practice, with a multiply degenerate ground state; this could be done following the lines of [23, 24]. A second improvement consists in taking into account the atomic motion. This is in principle a formidable task, because it leads to a spectacular increase in the dimension of the relevant Hilbert space. This addition can, however, be performed in practice in some limiting cases, for example if one assumes that the particles are tightly bound in a lattice [38, 39]. When the atom–light interaction is used only to probe the spatial atomic distribution of the gas, neglecting the particle motion should not be a major problem. Indeed the duration of the light pulse is quite short (∼10 μs only). Each atom scatters only a few photons in this time interval and its displacement is then smaller than the mean interatomic spacing for the spatial densities encountered in practice. The acceleration of the atoms under the effect of resonant van der Waals interaction should also have a minor effect under relevant experimental conditions.

Finally, another aspect that could be valuably studied is the interaction of the gas with an intense laser beam [40]. One could thus validate the intuitive idea that saturation phenomena reduce the effects of resonant van der Waals interactions [8, 41], and are thus helpful in providing a faithful estimate of the atomic density from the light absorption signal. A first step in this direction consists in implementing the tools developed in [42, 43], which allow one to study the modification of the coherent propagation of light in a dilute medium, due to the nonlinear optical response of the scatterers. In particular, it would be interesting to investigate whether the instabilities found in [43] play a significant role in absorption imaging of dense atomic clouds.

Acknowledgments

We thank I Carusotto, Y Castin, K Günter, M Holzmann, R Kaiser, W Krauth and S P Rath for helpful discussions and comments. L Chomaz acknowledges financial support from Direction Générale de l'Armement. This work was supported by Institut Francilien de Recherche sur les Atomes Froids and Agence Nationale de la Recherche (the project BOFL).

Footnotes

  • As to the derivation leading from equation (10) to equation (12), one must take into account the non-resonant terms that are usually dropped in the RWA, in order to ensure proper convergence of the sum (37) and obtain the tensor gαα'.

undefined