• Keine Ergebnisse gefunden

D.2 Discrete Fourier transform

7.1 Theoretical background

In strongly scattering samples light is scattered many times before it leaves the sample again. This results in a turbid and soft appearance, like e.g. that of a cloud in the sky. If light is scattered often enough in a disordered material, each light path can be regarded as a random walk. One can say that light is diffusing.

Mathematically the transport of light can in this case be described by the diffusion equation, which makes the theoretical treatment of those systems possible.

The foundations of DWS go back to the eighties, when Maret and Wolf [107] investigated multiply scat-tered light in dense samples of randomly distributed interacting particles under Brownian motion. They showed that the diffusion approximation is well suited to describe the temporal correlations measured in the fluctuating intensity of speckles originating from a coherent light beam going through the sample.

Since then, DWS has been established and has become a standard tool to retrieve information about the dynamics of many optically dense systems like particle suspensions, emulsions, foams, polymers, gran-ular matter etc. [108]. Some of the theoretical ideas behind DWS are given in the following. We will partly follow the explanations in [109] and [110], which are good reviews including detailed derivations together with important applications.

7.1.1 General form of the autocorrelation function

In DWS we want to see a signal from the motion of the scatterers in a highly scattering turbid sample. Like in single dynamic light scattering (DLS), it is retrieved from measuring the auto correlation function𝑔1(𝑑) of the electric field𝐸(𝑑) =βˆ‘π‘

𝑗=1πΈπ‘—π‘’βˆ’π‘–πœ™(𝑑), which is produced by all𝑁interfering plain waves registered in the detector:

𝑔1(𝑑)≑ βŸ¨πΈβˆ—(𝑑)𝐸(0)⟩

⟨|𝐸2|⟩ = 1

⟨𝐼⟩

⟨ 𝑁

βˆ‘

𝑗,π‘˜=1

πΈπ‘—βˆ—π‘’βˆ’π‘–πœ™π‘—(𝑑)πΈπ‘˜π‘’π‘–πœ™π‘˜(0)

⟩

=

βˆ‘π‘ 𝑗=1

βŸ¨πΌπ‘—βŸ©

⟨𝐼⟩

βŸ¨π‘’βˆ’π‘–Ξ”πœ™π‘—(𝑑)⟩

(7.1) Each of the𝑁waves corresponds to one of the light paths through the sample. For the last equality we define𝐼𝑗 =πΈπ‘—βˆ—πΈπ‘—and use that the field vectors𝐸𝑗and the phaseΦ𝑗 are not correlated. Their ensemble averages βŸ¨β€¦βŸ©can thus be computed individually. Further we assume that the fields of different light paths are uncorrelated, so that only summands with𝑗 =π‘˜remain. This approximation obviously works well for independent particles. But it was also proven to be valid for interacting particles, as long as their interactions are short-ranged compared to mean free path of light [111]. For single scattering, the phase shift for each wave due to the movement of the scattering particle isΞ”πœ™π‘—(𝑑) =π‘žβƒ—β‹…Ξ”βƒ—π‘Ÿπ‘—. In the case of multiple scattering, many scatterers contribute to the phase shift of each light path. As indicated in Figure7.1(to a very good approximation [109]) the total phase shift for one path is the following sum of the phase shifts from all scattering sites:

Ξ”πœ™π‘—(𝑑) =

βˆ‘π‘›(𝑗) 𝑖=1

βƒ—

π‘žπ‘–β‹…Ξ”βƒ—π‘Ÿπ‘–(𝑑) (7.2)

Here𝑛(𝑗)is the number of scatterers contributing to path number𝑗. Theπ‘žβƒ—π‘–are the individual scattering vectors for each scattering particle andΞ”βƒ—π‘Ÿπ‘–is the displacement of that particle. In order to simplify our

7.1 Theoretical background

Figure 7.1:A selection of detected light paths in a multiply scattering medium. Light comes in from the left and is detected on the right hand side. Paths in blue correspond to𝑑= 0and paths in light blue to a later time𝑑. The total phase shift for each light path due to the motion of the scatterers can be computed from the individual phase shiftπ‘žβƒ—π‘–β‹…Ξ”βƒ—π‘Ÿπ‘–(𝑑)at each scatterer. While some correlation remains after time𝑑 for each individual path, among themselves the paths are uncorrelated.

expression for𝑔1(𝑑)further, we rewrite the sum over all𝑁different light paths as a sum over the possible number of scatterers𝑛: which is the relative contribution of all light paths with𝑛scatterers to the total intensity. Now the average is not only an ensemble average over particle positions but also an average over the possible individual scattering vectorsπ‘žβƒ—π‘–taking part in paths with𝑛scatterers. To obtain the second equality in7.3, we assume that the successive phase shiftsπ‘žβƒ—π‘–Ξ”π‘Ÿπ‘–(𝑑)are uncorrelated. This is a safe assumption in most disordered samples. Hence the average over allπ‘žβƒ—π‘–is reduced to the averaging over the scattering vector in a single scattering event.

In the following we consider a sample that consists of particles in a solvent or a gas that does not contribute to the scattering. To go further we need to assume that the particle displacementsΞ”βƒ—π‘Ÿπ‘–obey a Gaussian distribution. As discussed in AppendixBthis holds for free Brownian motion and also for Brownian motion in a harmonic potential. In the previous chapter (6.2.4) we have seen that there is some deviation from the Gaussian distribution in glassy systems. However, this deviation is relatively small, so the Gaussian approximation should be sufficiently good. The scalar productπ‘žβƒ—π‘– β‹…Ξ”βƒ—π‘Ÿπ‘– then equally obeys a Gaussian distribution. By using that only moments of even order are nonzero, we obtain

⟨

For the last equality we need the assumption that scattering vectorsπ‘žβƒ—π‘–and corresponding displacements Ξ”βƒ—π‘Ÿπ‘– are not correlated, the factor 13 comes in because we are in 3D. As a consequence, hydrodynamic interactions are neglected from here on. The ensemble average over particle positions is restricted to

βŸ¨Ξ”π‘Ÿ2(𝑑)⟩, so that it gives the mean squared displacement (MSD). An average over absolute values of the scattering vectorπ‘ž remains to be done. As long as the MSD is small compared toπ‘žβˆ’2, which is always true for small lag times𝑑, one can restrict the calculation to the first term of a cumulant expansion [110]:

⟨ Now theπ‘žaverage is restricted to the termβŸ¨π‘ž2⟩. As mentioned above, we only need the distribution ofπ‘ž

for a single scattering event. This is given by the product𝑃(π‘ž)⋅𝑆(π‘ž)of the form factor and the structure factor, which would be the result of a static single light scattering experiment (cf. Equ.1.26). Then we obtain Here we use that2π‘˜0𝑛ef fis the largest possible scattering vector that occurs for backscattering. The wave vector π‘˜0 = (2πœ‹)βˆ•πœ†corresponds to the wavelength of lightπœ†outside the sample,𝑛ef f is the refractive index in the sample. For the second equality in 7.6we use the relation π‘ž = 2π‘˜0𝑛ef fsin(πœƒβˆ•2)between scattering vectorπ‘žand scattering angleπœƒ.

The average 𝑔 = ⟨1 βˆ’ cosπœƒβŸ©is known as the anisotropy factor. For 𝑔 = 1the scattering is isotropic:

The same intensity is scattered toπœƒand toπœ‹βˆ’πœƒ, so that the correlation between incoming and outgoing direction is lost. This is the case if the structure factor can be neglected, i.e. 𝑆(π‘ž) ≃ 1, and the diameter𝜎 of the particles is small compared to the wavelengthπœ†(Rayleigh scattering for𝜎 β‰ͺ πœ†in dilute systems).

Anisotropic scattering is indicated by𝑔 <1. It occurs in dense particle systems where𝑆(π‘ž) β‰ 1and/or for particles with𝜎 ≳ πœ†which must be described as Mie scatterers [112]. They usually scatter significant proportion of the light in forward direction to small anglesπœƒ.

To transfer the description into a picture of diffusing light waves, we need know the average distance between two scatterers in a light path. It is given as the mean free path𝑙.

𝑙 = 1 It decreases with an increasing density of the scatterers𝜌and an increasing scattering cross section𝜎. A calculation of𝜎[112] involves the form factor𝐹(πœ‘, πœƒ)or𝐹(π‘ž)and additionally the structure factor𝑆(π‘ž) in dense systems. Since the structure peaks of𝑆(π‘ž)increase with increasing density𝜌, the mean free path𝑙 is no more inversely proportional to𝜌in dense systems. For systems with anisotropic scattering 𝑔 < 1it makes sense to define an average lengthπ‘™βˆ—, the so called transport mean free path, after which light has travelled long enough so that the correlation of its direction with the incoming direction is lost.

A natural definition is to divide𝑙by the anisotropy factor𝑔[110]:

π‘™βˆ— = 𝑙 For the second equality one uses Equation7.6. This makes clear thatπ‘™βˆ—depends primarily on the values of𝐹(π‘ž)and𝑆(π‘ž)for largeπ‘ž close to the backscattering directionπ‘ž = 2π‘˜0𝑛ef f.

Now we can rewrite the autocorrelation function 𝑔1(𝑑) in terms of path lengths𝑠 = 𝑛𝑙 instead of the number of scatterers𝑛. We combine Equation7.3with7.4,7.5,7.6and7.8to get:

𝑔1(𝑑) =βˆ‘

Here, eventually we go to the continuum limit and write the sum over all occurring path lengths𝑠as an integral over the distribution𝑃(𝑠). One should note that this integration implies anincoherentsum in a sense that each light path contributes independently to the decay of𝑔1(𝑑) [109]. Hence the correlation function in Equ.7.9is comparable to the incoherent intermediate scattering function𝐹𝑠(π‘ž, 𝑑)for single light scattering, which is equally a function of the MSD (in the Gaussian approximation, cf. Equ.1.23).

7.1 Theoretical background It has been found that coherent parts make up for a fraction ofπ‘βˆ’1withπ‘β‰ˆ (π‘˜0π‘™βˆ—)(π‘˜0𝐿)[109], where𝐿is the minimum distance that light must travel through the sample. Thus forπ‘™βˆ— ≫ πœ†coherent contributions to the decay of𝑔1(𝑑)are negligible.

It is also possible to rewrite𝑔1(𝑑) in terms of a diffusion coefficient𝐷 = βŸ¨Ξ”π‘Ÿ2(𝑑)βŸ©βˆ•(6𝑑). Now one can also include the dependence of𝐷on particle interactions by taking𝐷(π‘ž) =π·π‘ β„Ž(π‘ž)βˆ•π‘†(π‘ž), whereβ„Ž(π‘ž) = 𝐻(π‘ž)βˆ•π»(∞)describes the hydrodynamic interactions and𝐷𝑠 is the short time diffusion coefficient. In this case one can omit the approximation applied in Equ.7.4, that requiredΞ”βƒ—π‘Ÿandπ‘žβƒ—to be independent:

𝑔1(𝑑) = With the use of visible light, the latter approximation is valid in systems with big particles (𝜎 >2Β΅m) [109]. For those one can equally use Equation7.9, that directly connects𝑔1(𝑑)to the MSD. For particles of a size𝜎= 412nm the approximation leads to a deviation of about20 %[109] in the decay time. Thus, one should have in mind that the error might be even larger for smaller particles.

It is convenient to hide the specialties of the particle system in a characteristic decay time𝜏0: 𝜏0 =[

(π‘˜0𝑛ef f)2𝐷𝑠]βˆ’1

= 6𝑑⋅[

(π‘˜0𝑛ef f)2βŸ¨Ξ”π‘Ÿ2(𝑑)⟩]βˆ’1

, (7.11)

Then one obtains a more clear and catchy form for the autocorrelation function:

𝑔1(𝑑) = the correlation decreases by a factor ofexp(βˆ’2π‘‘βˆ•πœ0). Note thatexp(βˆ’2π‘‘βˆ•πœ0) is exactly the correlation function for single scattering (DLS) at an angle of90β—¦(π‘ž=√2π‘˜0𝑛ef f).

Summary of approximations

Most important to obtain an explicit expression for𝑔1(𝑑)is the assumption of independent light paths, which should be well-fulfilled forπ‘™βˆ— ≫ πœ†. Somehow connected to this is the assumed independence of successive scattering events in each light path. If𝑙 ≫ πœ†and if the correlation between particle positions (seen e.g. in𝑔(π‘Ÿ)) has decayed at a distance of 𝑙, this should be valid as well. Possibly problematic for glassy systems could be the assumption of a Gaussian distribution for the displacementsΞ”βƒ—π‘Ÿπ‘–of the scatterers. To obtain Equation7.9, the scattering vectorπ‘žin a single scattering event must be uncorrelated to the displacement of that scatterer. Collective movements, that occur e.g. in hydrodynamic interactions, are therefore ignored. By usage of Equation7.10one could at least approximately include hydrodynamic interactions, but this goes beyond the scope of this thesis. Another approximation is done in Equation7.5, where the lag time𝑑must to be small enough, so that βŸ¨Ξ”π‘Ÿ2(𝑑)⟩ β‰ͺ πœ†2. However, in most of the cases, the average path length is much larger thanπ‘™βˆ— so that the factorπ‘ βˆ•π‘™βˆ— in the exponent of Equations7.9 and7.10leads to a decay of𝑔1(𝑑)long before the MSD becomes comparable to the wavelengthπœ†.

7.1.2 Diffusion equation and path length distribution𝑃(𝑠)

In the diffusion approximation (for length scales larger thanπ‘™βˆ—) one can describe the energy densityπ‘ˆ(βƒ—π‘Ÿ, 𝑑) of light with the diffusion equation

πœ•π‘ˆ(βƒ—π‘Ÿ, 𝑑)

πœ•π‘‘ =π·π‘™βˆ‡2π‘ˆ(βƒ—π‘Ÿ, 𝑑) with 𝐷𝑙= π‘π‘™βˆ—

3 . (7.13)

The diffusion coefficient𝐷𝑙contains the speed of light in the medium𝑐 =𝑐0βˆ•π‘›ef fand the transport mean free pathπ‘™βˆ—, the distance after which the direction of light is randomized, the division by3accounts for the number of space dimensions. For a solution of Equ.7.13one also needs initial and boundary conditions.

The idea is to compute the outgoing flux𝐽⃗out(βƒ—π‘Ÿout, 𝑑)at a positionβƒ—π‘Ÿout at the boundary of the sample after an instantaneous light pulse has started to diffuse in the sample at time 𝑑 = 0. Since light needs the time𝑑=π‘ βˆ•π‘to reach the position of the detector atβƒ—π‘Ÿout, the distribution𝑃(𝑠) is directly proportional to 𝐽⃗out(βƒ—π‘Ÿout, 𝑑=π‘ βˆ•π‘). Because the light path is only randomized after a distance ofπ‘™βˆ—the starting point for the light pulse is not directly at the border of the sample. Instead one defines it to start at a distance𝑧0 β‰ˆπ‘™βˆ— inside the sample. Therefore, as an example, the initial condition for a sample with a planar boundary at 𝑧= 0and a parallel beam coming from𝑧 <0is given by:

π‘ˆ(𝑧, 𝑑= 0) =π‘ˆ0𝛿(π‘§βˆ’π‘§0, 𝑑) (7.14)

Boundary conditions are usually obtained by requiring that the incoming net flux into the sample is zero [110]. This results in the condition

( π‘ˆ+ 2

3π‘™βˆ— Μ‚βƒ—π‘›β‹…βˆ‡π‘ˆ) |

||βƒ—π‘Ÿβˆˆπœ•π‘‰ = 0, (7.15)

whereπœ•π‘‰ means the surface of the sample and̂⃗𝑛is a unit vector which is always orthogonal to that surface and directing out of the sample. Other boundary conditions, e.g. absorption of light at the surface, were investigated, but Equ.7.15gave the best agreement with experiments [109]. With a solutionπ‘ˆ(βƒ—π‘Ÿ, 𝑑)for the full diffusion problem, the path length distribution is given by

𝑃(𝑠) ∝|||𝐽⃗(βƒ—π‘Ÿout, 𝑑)|||=𝐷𝑙|||Μ‚βƒ—π‘›β‹…βˆ‡π‘ˆ(βƒ—π‘Ÿout, 𝑑)|||= 𝑐

2π‘ˆ(βƒ—π‘Ÿout, 𝑑=π‘ βˆ•π‘) (7.16) Here one makes use of Fick’s law of diffusion and the boundary condition in Equation7.15. Since the autocorrelation function𝑔1(𝑑)is basically a Laplace transform of𝑃(𝑠)(cf. equs.7.9and7.10) it is often convenient to solve the diffusion problem in Laplace space and obtain directly an expression for𝑔1(𝑑) [109].

7.1.3 Transmission geometry

The detailed derivation of the path length distributions for arbitrary geometries and the final computation of𝑔1 uses nontrivial mathematics and goes beyond the scope of this description. Only the formulas for the two most important scattering geometries (see Fig.7.2), that were also used in this work, are presented here including comments.

In transmission geometry, a parallel beam of coherent laser light uniformly illuminates one side of a slab with a thickness𝐿in𝑧direction and an infinite extent in the other two directions. In experiments this idealized situation is well approximated using a beam width much larger thanπ‘™βˆ—and a sample that is larger than the beam width by more than10π‘™βˆ—. As mentioned in the preceding paragraph, light is considered to

7.1 Theoretical background

Figure 7.2: Idealized transmission and backscattering geometry for a slab of thickness𝐿with infinite extent perpendicular to the incoming beam. The coherent parallel light is considered to be homogeneous with an infinite width, like an idealized plane wave. Precise angles of the detector with respect to the incoming beam are arbitrary. Sketches of typically detected light paths are given in red.

start diffusing from a plane at𝑧0β‰ˆπ‘™βˆ—. However, the exact value of𝑧0is not very important since𝐿 ≫ π‘™βˆ— is required anyway. At least𝐿 >10π‘™βˆ—is necessary for the diffusion approximation to be valid. Using the characteristic decay time of the particle system𝜏0, one obtains the autocorrelation function [109]:

𝑔1(𝑑) =

As the approximation suggests, the shape of this function somewhat resembles an exponential decay. But particularly for large𝑑, deviations from exponential behaviour are observed. The characteristic decay time is𝜏0(π‘™βˆ—βˆ•πΏ)2. Recalling that𝑔1(𝑑) decreases by a factorexp(βˆ’2π‘‘βˆ•πœ0)for each step of the random walk, one infers that the mean number of steps isβŸ¨π‘›βˆ—βŸ©= 1

2(πΏβˆ•π‘™βˆ—)2. This corresponds to an average path length ofβŸ¨π‘ βŸ© = 𝐿2βˆ•(2π‘™βˆ—). An easy way to derive this is to consider the growing mean squared displacement 6𝐷𝑙𝑑 of the photons diffusing in the sample: In order to diffuse an average distance of𝐿, which is the sample thickness, light needs the time𝐿2βˆ•(6𝐷𝑙) =𝐿2βˆ•(2𝑐 π‘™βˆ—). A division by the speed of light𝑐yields the average lengthβŸ¨π‘ βŸ©of a light path. Since all light paths must travel the distance𝐿in𝑧direction, the distribution𝑃(𝑠)has a sharp maximum at the average path lengthβŸ¨π‘ βŸ©.

From the definition of𝜏0in Equation7.11one can see that each scatterer moves an average distance of Ξ”π‘Ÿ0 =√ transmis-sion geometry is very sensitive to the dynamics in the sample. An average translation of the scatterers by only a fractionπ‘™βˆ—βˆ•πΏof the wavelength is enough to destroy the correlation. By adjusting the thick-ness of the sample cell the sensitivity can be adjusted, for instance starting from a decay for an average displacement of∼ 50nm in a thin cell down to several nanometers in a thick cell.

7.1.4 Backscattering geometry

The path length distribution𝑃(𝑠)for the backscattering geometry is a lot broader than that for the trans-mission case. Both very short light paths with a length of a fewπ‘™βˆ— and very long paths are possible.

Incoming beam and detector are on the same side of the sample (see Fig.7.2). The exact angle at which light is collected is not important. But it is substantial that the incoming parallel beam is expanded much

wider thanπ‘™βˆ—. The detector needs to focus at a spot near the centre of the illuminated area. If the latter two conditions are not fulfilled, the path length distribution is qualitatively different. For instance, if the detector is focused to the border of the illuminated area, the contribution of short paths becomes smaller while that of large paths becomes longer.

For a computation of𝑔1(𝑑)one again assumes that the diffusing light starts from a plane at𝑧=𝑧0 =π›Ύπ‘™βˆ— in the sample. The exact value of𝛾 β‰ˆ 2(see [110]) depends on the sample and the sample cell (e.g.

on the particle sizes, thickness of glass walls, reflectivities etc.). Additionally, there is a dependence of𝛾 on the polarization of incoming and detected light, which is measurable by putting a polarizer in front of the detector. If the polarizations of incoming and detected light are perpendicular the very short paths coming from one single backscattering are not detected, because single scattering preserves the polarization. This yields an increase in𝛾. On the other hand, if incoming and detected light have mainly parallel polarizations,𝛾decreases compared to a measurement without a polarizer.

For the longer paths𝑠 ≫ π‘™βˆ—, one can show that𝑃(𝑠) βˆπ‘ βˆ’3βˆ•2. Therefore,𝑃(𝑠)only decays algebraically for𝑠 β†’βˆž. This means that also longer paths contribute to the decay of𝑔1, not only the more frequent short paths. Since many scatterers contribute to the phase shift in long paths, very small displacements Ξ”βƒ—π‘Ÿπ‘–are enough to destroy their correlation. Thus, the long paths are responsible for the decay of𝑔1(𝑑) at small lag times 𝑑. On the other hand, the longer lag times are dominated by the correlations of the short paths.

For a sample of thickness𝐿, the exact solution of the diffusion equation yields [109]:

𝑔1(𝑑) = accu-rate normalization𝑔1(𝑑 = 0) = 1. Since paths of many different lengths𝑠contribute, the decay upon increasing the lag time𝑑is slower than exponential. By comparison ofexp(βˆ’π›Ύβˆš

6π‘‘βˆ•πœ0)with the decay exp(βˆ’2π‘‘βˆ•πœ0)for90β—¦ single scattering, one comes to the conclusion that for𝑑 >3βˆ•2𝛾2𝜏0the correlation for backscattering DWS is actually larger than that for single scattering. It is however clear that correla-tions of multiply scattered light must decay faster. Therefore, the result given in Equation7.19can only be valid for small lag times𝑑 β‰ͺ 𝜏0. This is not only true for the approximation, but also for the exact solution of the diffusion equation. The reason is that for𝑑 > 𝜏0 the correlations mainly stem from the shortest paths. But for short paths, many of the approximations mentioned above (see7.1.1), including the diffusion approximation, are not valid.

Still DWS in backscattering geometry is a very useful tool especially for glassy samples. It allows to probe the dynamics at many length scales at once. Small movements of the particles contribute to the decay of𝑔1(𝑑)at short lag times. But even quite large displacements do not yield a total decay, because the dominating short light paths are still correlated. Using the definition of𝜏0in Equation7.11together with the approximated𝑔1(𝑑)from above, we see that𝑔1(𝑑)has decayed to1βˆ•π‘’with an average displacement of the scatterers by:

Ξ”π‘Ÿ0= (π›Ύπ‘˜0𝑛ef f)βˆ’1 =πœ†β‹…(2πœ‹π‘›ef f𝛾)βˆ’1 (7.20) This is only a factorβˆΌπ›Ύearlier than in a single scattering experiment. A total decay requires an average displacement by about one wavelengthπœ†.