Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Reference-free path length interferometry of random media with the intensity moments analysis

Open Access Open Access

Abstract

Stochastic interference of partially coherent light multiple scattered by a random medium is considered. The relationship between the second- and third-order moments of intensity fluctuations in random interference patterns, the coherence function of probe radiation, and the probability density of path differences for the interfering partial waves in the medium are established. The obtained relationships were verified using the statistical analysis of spectrally selected fluorescence radiation emitted by the laser-pumped dye-doped random medium. Rhodamine 6G water solution was applied as the doping agent for the ensembles of densely packed silica grains which were pumped by the CW radiation (532 nm) from the diode-pumped solid state laser. Experimentally observed abrupt decay of the second- and third-order moments of fluorescence intensity fluctuations for the wavelengths ranging from 620 nm to 680 nm is interpreted in terms of amplification of spontaneous emission at large dye concentrations. This paper discusses the new optical probe of random media defined as “the reference-free path length interferometry with the intensity moments analysis”.

© 2017 Optical Society of America

1. Introduction

Light scattering probes of the random media play an important role in various applications in modern material science and biomedicine. The vast variety of physical phenomena accompanying elastic, quasi-elastic, or inelastic light interactions with the probed medium serves as the physical basis for these probes. These interactions modify all or part of the properties (amplitude, frequency, initial phase, and polarization state) of the light waves propagating in the medium. Accordingly, the changes in these properties occurring due to multiple scattering of the probe light in the medium contain the necessary information about the structure, or dynamics, or chemical composition of the examined substance.

The coherence domain probes of the random media are associated with elastic or quasi-elastic light-medium interactions; these terms mean that the propagating waves do not change or insignificantly change the frequency in the sequences of scattering events, whereas random variations in the propagation direction from one scattering event to another can be significant. This physical picture illustrating the features of such interactions corresponds to the so-called discrete scattering model [1, 2] previously applied for the phenomenological description of coherence phenomena in the multiple scattering (see, e.g., [3–7]).

The diffusing wave spectroscopy pioneered by G. Maret with P.-E. Wolf [8] and D.J. Pine et al. [9] is the most popular coherence domain technique used to characterize the scatter dynamics in the complex multiple scattering systems at the spatial scales of the order of the used wavelength. This technique applied in probing of various scattering systems (from biological tissues and soft media [10–15] to sand flows [16]) in the past three decades has undergone essential modifications in part of instrumentation and data processing procedures from the classical single-point detection schemes to the multispeckle arrangement [17–19].

In the “frozen” media with the stable, time-independent scatter positions, the multiple scattering of the coherent light is not accompanied by the temporal decorrelation of the scattered radiation, which exhibits the static speckle modulation. Such media can be considered as multiple-beam interferometers with stochastically distributed path differences between the interfering beams. Speckle patterns modulating the outgoing multiple scattered light are associated with stochastic interference of partial contributions to the light field in the medium. Thus, the probing techniques based on the statistical analysis of such stochastic interference patterns can be identified as the reference-free stochastic interferometry of random media. Additionally, application of the field correlation techniques (see, e.g., [20, 21]) provides promising results in the area of coherence-domain diagnostics of random scattering systems characterized by the stable structure.

Typically, the multiple scattered speckle-modulated light is characterized in the weak scattering limit by the Rayleigh statistics of intensity fluctuations, which results from the superposition of two fully developed non-correlated speckle patterns with equal values of average intensity. These speckle patterns correspond to the independent orthogonally polarized modes of the propagating light in the medium [22–24].

The transition from coherent to partially coherent illumination of multiple scattering random media can cause a remarkable suppression of stochastic interference modulation of the outgoing multiple scattered light in the case when the average path of light propagation in the medium is comparable or exceeds the coherence length of the light. Consequently, varying the coherence length of the probe light or changing the light propagation conditions, and evaluating the speckle contrast or oscillation index in the outgoing light, we can characterize the optical transport properties of the probed medium or visualize its inhomogeneous structure. In particular, the imaging of macroscopically heterogeneous random media using speckle contrast as the visualization parameter and illumination source with the fixed coherence length (a He-Ne laser) was considered in [25]. Evaluation of the optical transport properties (the mean transport free path and the scattering anisotropy parameter [1],) of the random media using the probe light with the tunable coherence length and speckle contrast estimates for the outgoing multiply scattered light was discussed in [26]. In this case, the coherence length of the probe light was varied by changing the pumping current in the light source (the laser diode operating below the generation threshold). The polarization discrimination of the outgoing light was applied for additional characterization of the light depolarization length in the medium. An alternative approach to the coherence length tuning in the reference-free stochastic interferometry is based on frequency modulation of the light sources. In particular, the periodic binary modulation of the output frequency of the laser diode with the tunable modulation depth in combination with the long-exposure capture of the speckle patterns ensures the necessary effect [27].

Another way to the reference-free low-coherence interferometry of the random media can be related to the broadband sources of the probe light and the narrow-band spectral selection of the outgoing multiple scattered radiation. In this case, the coherence length of the detected radiation is determined by the selection conditions lcλc2/Δλ (Δλ is the width of the spectral window used for the selection, and λc is the central wavelength). Note that such broadband light source can be embedded in the probed medium. In particular, this approach can be provided by doping the probed medium with the fluorescent dye with follow-up pumping by the CW laser radiation [28, 29].

Interpretation of the empirical data obtained by the above cited modifications of the reference-free low-coherence interferometry is usually based on the approximate analytical or numerical inversion of integral transformation of the probability density function of the path differences Δs for partial contributions to the optical field in the medium. This integral transformation contains the square of the coherence function module (or spectral density) for the probe light and establishes the relationship between the second-order statistical moment of the outgoing light intensity or its derivative such as speckle contrast and coherence length of the probe light for the used illumination and detection conditions. It is worth stressing that higher-order statistical moments of the detected speckle intensity must exhibit higher sensitivity to the variations in the optical transport properties of the examined medium compared to with the generally applied normalized second-order moment I2/I2 and its derivative speckle contrast V=σI/I=I2I2/I. In particular, intensity statistics in the case of the narrow-band spectral selection of fluorescence radiation outgoing from the laser-pumped dye-doped random medium exhibits a strong influence of the wavelength of detected light on the skewness of intensity distributions [28]. The presumable interpretation of this effect is related to a strong wavelength dependence of the ratio of average Δs value to the coherence length associated with amplification of spontaneous fluorescence emission for the long-path partial contribution of optical field in the medium under condition of high dye concentrations [28, 29].

The aim of this work is to consider a novel approach to the analysis of empirical data obtained using the reference-free low-coherence interferometry of the random media in application to the case of position-dependent narrow-band spectral selection of fluorescence radiation from the CW laser-pumped dye-doped random media. In our opinion, this case could be of great interest for numerous practical applications beginning from material science and ending biomedical optics. Additionally, this analysis can be useful for better understanding of statistical properties of partially coherent light multiple scattered by the random media.

2. Second- and third-order statistical moments of intensity fluctuations of multiple scattered partially coherent light: from the discrete scatter model to continuous path length distributions

We will consider the statistical properties of partially coherent light propagating in the multiple scattering random medium using the following approach.

  • 1. The stochastic interference of N partial contributions (plane waves) emitted within the spectral range from λ to λ+Δλ by the continuous-wave source in the random medium is considered; the spectral range (λ,λ+Δλ) is determined by the applied detection system and the condition of quasi-monochromaticity (Δλ<<λ,Δω<<ω¯) is satisfied (ω¯ is the average frequency of interfering partially coherent waves, Δω=2πcΔλ/{λ(λ+Δλ)}2πcΔλ/λ2 is the detection bandwidth, and cis the speed of light).
  • 2. The second- and third-order moments of intensity fluctuations are calculated.
  • 3. The obtained expressions are modified using transition from the discrete path length distribution p1,p2,....,pN for the ensemble of interfering partial waves to continuous path length distribution with the probability density function ρ(s) under the conditionN.
  • 4. The case of stochastic interference of the non-polarized light will be considered assuming statistical independence of partial waves propagating in the co- and cross-polarization channels. Despite its simplicity, this approach allows an adequate description of the variety of interference and polarization effects associated with multiple scattering in the random media (see, e.g., [30]). Therefore, first we shall consider interference of “scalar” waves assuming the same linear polarization states for all interfering waves. Then we shall generalize the obtained result to the case of the non-polarized light.

In the case of superposition of N partially coherent waves emitted by the continuous-wave source, the resultant time-averaged intensity in the arbitrarily chosen detection point can be expressed as

I(t)¯=E(t)E*(t)¯=|E(t)|2¯=|iNEi(t)exp{jφi(t)}|2¯,
where time averaging denoted by the overline symbol is carried out over the time interval T>>2π/Δω, E(t) is the resultant time-dependent electric field in the observation plane, and Ei(t)=Ii(t), φi are the time-dependent amplitude and the phase of the i-th partial wave (Ei(t) is the real non-negative value and Ii(t) is the slowly varying intensity of the partial wave). The phase term φi can be written in the following form: φi(t)=ω¯tksi+ϕ(tksi/ω¯), where ϕ(t) denotes the randomly fluctuating part of the phase caused by the finite bandwidth of emitted radiation, the terms ksi and ksi/ω¯ correspond to the phase and time delay due to propagation of the partial wave between the source and the detector, k is the wavenumber of the propagating light in the medium, si is the propagation path length for the i-th partial wave. Note that the random part of the phase ϕ(tksi/ω¯) is the common term for all N contributions since we consider them as emitted by the same continuous-wave (CW) source. At the same time, the time delay term ksi/ω¯ causes decorrelation between the values of the electric field of the waves propagating along the various paths in the medium. The expression (1) can be written in the following form
I(t)¯=iNiNEi(t)Ei(t)exp[j{φi(t)φi(t)}]¯==iNiNIi(t)Ii(t)exp[j{φi(t)φi(t)}]¯.
Each term in Eq. (2) describes the interference of ith and ith partially coherent waves; the partial coherence is caused by the difference in the time-dependent random phase fluctuations ϕ(tksi/ω¯) and ϕ(tksi/ω¯)of the interfering waves. Considering the sum of the terms Ii(t)Ii(t)exp[j{φi(t)φi(t)}]and Ii(t)Ii(t)exp[j{φi(t)φi(t)}] in the right-hand side of Eq. (2), we can rewrite this equation as
I(t)¯=iNIi(t)¯+iNiiNIi(t)Ii(t)cos[k(sisi)+Δϕ(t,sisi)]¯,
whereΔϕ(t,sisi)=ϕ(tksi/ω¯)ϕ(tksi/ω¯).

It should be taken into account that partial waves propagate into the probed multiple scattering medium along various random paths, and undergo random sequences of statistically independent scattering events. Each scattering event leads to random transformation in the amplitude of the propagating wave Eim+1(t)=AimEim(t) with the factor Aim dependent on the random scattering vector qim=kim+1kim, where kim and kim+1 are wave-vectors of the i-th propagating wave before and after the m-th scattering event. The products Ai1Ai2AiM and Ai1Ai2AiMfor i-th and i-th waves are statistically independent random magnitudes. Thus, we can assume statistical independence of the values Ei(t)=Ii(t), Ei(t)=Ii(t), φi(t), and φi(t) for each pair of interfering partial waves. A similar assumption is one of the basic points of the discrete scattering model, which was successfully used to describe the various interference-based phenomena in the random multiple scattering media (see, e.g., [3, 5–8, 26, 27]).

Therefore the each time-averaged term Ii(t)Ii(t)cos[k(sisi)+Δϕ(t,sisi)]¯ can be transformed in the following way

Ii(t)Ii(t)cos[k(sisi)+Δϕ(t,sisi)]¯==Ii(t)¯Ii(t)¯cos[k(sisi)+Δϕ(t,sisi)]¯=Ii(t)¯Ii(t)¯cos[Δϕ(t,sisi)]¯cos[k(sisi)]Ii(t)¯Ii(t)¯sin[Δϕ(t,sisi)]¯sin[k(sisi)]==Ii(t)¯Ii(t)¯Q2(sisi)+S2(sisi)cos[k(sisi)Ψ(sisi)],
where
Q(sisi)=1T0Tcos[Δϕ(t,sisi)]dt;S(sisi)=1T0Tsin[Δϕ(t,sisi)]dt;tg{Ψ(sisi)}=S(sisi)/Q(sisi),
and the averaging is carried out over the time interval T>>2π/Δω. In further consideration, we will use the notations Ii(t)¯=E0i and Ii(t)¯=E0i. Also, the time-independent constant value I(t)¯ will be denoted as I. The term |g(sisi)|=Q2(sisi)+S2(sisi) is related to the modulus of the first-order coherence function of the considered partially coherent light. The value Ψ(sisi) is related to the principal part of argument of the coherence function, which in general is the complex function. It is obvious that the function |g(sisi)| is equal to 1 in the absence of phase fluctuations and falls to 0 in the case of expressed high-frequency phase fluctuations. On the other hand, the functions |g(sisi)| and Ψ(sisi) can be obtained using the inverse Fourier transformation of spectral density of emitted radiation [31]. In our case, we consider the interfering waves within the wavelength interval from λ to λ+Δλ. Assuming the uniform distribution of spectral density of the emitted light in the given spectral interval, we must expect the coherence function in the well-known form
g(sisi)=sin{π(sisi)/lc}/{π(sisi)/lc}exp{jk(sisi)},
where lc is the coherence length of emitted light. Following [31] (see section 4.3), we can conclude that in this case the term Ψ(sisi) is equal to 0. Therefore, we can rewrite Eq. (2) in the following form
I=iNE0i2+iNiiNE0iE0icos{k(sisi)}|g(sisi)|.
Here we suggest that the following condition takes place: |g(sisi)|=|g(sisi)|. Eq. (4) describes the following physical picture: the emitted quasi-monochromatic waves propagate in the “frozen” random medium with the fixed scattering sites. These waves interfere with each other. As a result, we obtain the random constant value of I at the detection point. Any changes in the ensemble of the propagation paths (s1,....,sN) cause the changes in the interference terms in the right-hand side of Eq. (4), and, correspondingly, the changes in the detected intensity I. The propagation paths can vary due to variations in the position of the detection point, or the positions of scattering sites in the random medium. With the given coherence function of the propagating radiation, the contribution of the pair of waves i,i to the interference term is less, the more is the path length difference sisi.

Let us consider the following scheme of the medium probing: the random values I are sequentially measured at various positions of the detection point, and the obtained set of random values (I1,....,IP) is used to calculate the second- and third-order statistical moments of spatial intensity fluctuations. In this case the averaging procedures should be carried out over the ensemble of all the possible path lengths si. The second-order moment of the spatial intensity fluctuation can be written as

I2=(iNE0i2+iNiiNE0iE0icos{k(sisi)}|g(sisi)|)2,
where the angle brackets denote the ensemble averaging over the set of possible si values. Similarly, the third moment of intensity can be expressed as
I3=(iNE0i2+iNiiNE0iE0icos{k(sisi)}|g(sisi)|)3
Assuming the unit amplitudes E0i=E0i=1 of the interfering statistically independent partial waves and introducing the notation Δsii=sisi, we can rewrite Eqs. (5) and (6):
I2=(N+2mN(N1)/2cos(kΔsm)|g(Δsm)|)2,
I3=(N+2mN(N1)/2cos(kΔsm)|g(Δsm)|)3,
where the index m is associated with the difference ii and varies from 1 to N(N1)/2.

Appropriateness of replacing the random terms E0i in Eqs. (5) and (6) by the deterministic unit value can be validated considering the general statistical properties of sums of the random phasors E0iexp(jφi). In the case of statistical independence of the random variables E0i,φi and large values of the phase terms φi(|φi|>>π), the sum iE0iexp(jφi) of a large number of random phasors is characterized by identical Gaussian distributions of probability densities of real and imaginary parts. These parts are independent on statistical properties of amplitude terms E0i. This results from the central limit theorem (see, e.g [31], section 1.5.6). Therefore we must expect that probability density distributions of the random values I=|iE0iexp(jφi)|2, which correspond to the various probability density distributions of E0i, are identical if their mean values I are equal. This conclusion is illustrated (Fig. 1) by the results of statistical modeling of the probability density distributions ρ(I/I) for the two random sums I=|iE0iexp(jksi)|2. The first sum is characterized by the uniform probability density distribution of E0i with the average value E0i equal to 1.0, and the standard deviation σE0i equal to 0.57. The second sum is characterized by the “deterministic” probability density ρ(E0i)=δ(E0i1.0). Regarding the distributions ρ(I/I), we can see that these two cases are identical. We also took into account the suppression of interference due to partial coherence of the contributions E0iexp(jksi) by introducing the factors g{(sisn)/lc} into the products E0iE0nexp(jksi)exp(jksn). The coherence function g{(sisn)/lc} was taken in the form g{(sisn)/lc}=sin{π(sisn)/lc}/{π(sisn)/lc}, which corresponds to the conditions of our experiment (see below the section 4).

 figure: Fig. 1

Fig. 1 The simulated distributions ρ(I/I) in the cases of the deterministic value |E0i|=1 (1, 3) and random uniform distribution of the phasor amplitudes (2, 4). The number of statistically independent phasors used in the simulation procedure is equal to 500; the sample size applied for the reconstruction of ρ(I/I) is equal to 10000. The cases (1, 2) correspond to coherent summation of the random phasors (sism/lc=1.0⋅10−3) and the cases (3, 4) correspond to the partially coherent summation (sism/lc=1.0).

Download Full Size | PDF

Note that the simulated distributions ρ(I/I) exhibit the expected peculiarities of the coherence-dependent statistical properties of I (the negative exponential form of the probability density in the case of total coherence of contributions and the gradual narrowing of the probability density distribution ρ(I)δ(II) with transition to the incoherent summation of contributions). It should be mentioned that independence of phasor amplitudes from time is the necessary condition for the above consideration. This condition is met in the case of application of the intensity-stabilized CW source. Also, it can be provided using the appropriate detection time T>>2π/Δω~τc, there τc is the correlation time of amplitude fluctuations. In the case of amplitude fluctuations with the correlation time comparable to T, the probability density distributions I/I will deviate from the shapes presented in Fig. 1. As it is shown below (see section 4), the condition T>>2π/Δω is reliably fulfilled in our experiment. Thus, in our case we can assume validity of the above consideration and application of the deterministic value E0i=1 instead of the randomly distributed amplitudes of contributing phasors should not challenge the results of further analysis.

Equations (7) and (8) are transformed into the following forms

I2=N2+4NmN(N1)/2cos(kΔsm)|g(Δsm)|++4(mN(N1)/2cos(kΔsm)|g(Δsm)|)2,I3=N3+6N2mN(N1)/2cos(kΔsm)|g(Δsm)|++12N(mN(N1)/2cos(kΔsm)|g(Δsm)|)2++8(mN(N1)/2cos(kΔsm)|g(Δsm)|)3.
In the right-hand side the second term of these expressions is equal to zero due to statistical independence of the values Δsm. The analysis of the fourth term in the latter expression shows that only the terms cos(kΔsm)cos(kΔsm)cos(kΔsm)|g(Δsm)||g(Δsm)||g(Δsm)| with the indices m,m,m satisfying the following condition of the cyclic permutation m=ip,m=po,m=oi with m0,m0,m0 have non-zero values in the course of ensemble averaging. These values are equal to (1/4)|g(Δsm)||g(Δsm)||g(Δsm)|. The total amount of these terms is evaluated as N(N1)(N1) and the fourth term can be expressed in the following form 2N(N1)(N1)(|g(Δsm)|)3.

Finally, considering the transition from discrete to continuous distribution of sivalues under the condition N, we obtain the following expressions for the normalized second- and third-order moments in case of stochastic interference of the polarized partial waves

M2=I2I2=1+N1N0|g(Δs)|2ρ(Δs,Δs)d(Δs)1+0|g(Δs)|2ρ(Δs,Δs)d(Δs).
M3=I3I3=1+3N1N0|g(Δs)|2ρ(Δs,Δs)d(Δs)++2(N1)2N2(0|g(Δs)|ρ(Δs,Δs)d(Δs))31+30|g(Δs)|2ρ(Δs,Δs)d(Δs)+2(0|g(Δs)|ρ(Δs,Δs)d(Δs))3.
Here ρ(Δs,Δs) is the probability density function of the path length differences Δs evaluated as [26]
ρ(Δs,Δs)={20ρ(s+Δs)ρ(s)ds,Δs0;0,Δs<0,
where Δs is the ensemble-averaged path length difference. Note that I2/I2 approaches 2, and I3/I3 approaches 6 in the case of totally coherent light lc (this is the well-known property of the intensity moments for fully developed speckle patterns) and falls to zero in case of the non-coherent illumination lc0.

Let us consider stochastic interference of non-polarized partial waves; in this case we introduce the statistically independent local random values of the intensity I and III for orthogonally polarized components of the propagating light, which are characterized by equal statistical properties. Thus, the normalized second- and third-order moments of the total intensity I=I+III can be expressed as

I=2I,II;I2=(I+III)2=I2+2IIII+III2;I3=(I+III)3=I3+3I2III+3IIII2+III3;M2=I2I2=2I,II2+2I,II24I,II2=12I,II2I,II2+12;M3=I3I3=2I,II3+6I,II2I,II8I,II3=14I,II3I,II3+34I,II2I,II2.
The second- and third-order moments of intensity fluctuations for the non-polarized light with the pathlength difference probability density ρ(Δs,Δs) can be expressed as
{M2=I2I21+120|g(Δs)|2ρ(Δs,Δs)d(Δs);M3=I3I31+320|g(Δs)|2ρ(Δs,Δs)d(Δs)++12(0|g(Δs)|ρ(Δs,Δs)d(Δs))3.
Thus, the upper limits of I2/I2 and I3/I3 (the case of coherent non-polarized light) are equal to 1.5 and 3.0, respectively. This is the consequence from the moment rule for non-polarized speckle fields with the Rayleigh statistics of the intensity In/In=(n+1)!/2n.

3. Universality of the ratios of pathlength-dependent functionals

The numerical analysis of the functionals

J1(lc,Δs)=0|sin(πΔs/lc)πΔs/lc|2ρ(Δs,Δs)d(Δs);J2(lc,Δs)=(0|sin(πΔs/lc)πΔs/lc|ρ(Δs,Δs)d(Δs))3
with the boundary conditions
Δslc=0J1(lc,Δs)=J2(lc,Δs)=1;Δslc=J1(lc,Δs)=J2(lc,Δs)=0
allowed us to ascertain with appropriate accuracy the universality of J1(lc,Δs)/J2(lc,Δs) for the various forms of the probability density distributions ρ(Δs,Δs) and coherence function given by Eq. (3). Figure 2 illustrates this universality occurring in the ranges 0.15<J1(lc,Δs)1.0;0.015<J2(lc,Δs)1.0 for various numerically simulated cases

 figure: Fig. 2

Fig. 2 The plot of values J1(lc,Δs) and J2(lc,Δs) for the various shapes of ρ(Δs,Δs); a – open triangles; b – closed circles; c – open circles. The Δs/lcratios are equal to: i – 0; ii – 0.318; iii – 0.637; iv – 0.955; v – 1.273; vi – 1.592; vii – 2.228; viii – 3.183. Dashed line – the power-law approximation of the relationship between J1(lc,Δs) and J2(lc,Δs).

Download Full Size | PDF

  • (a) – the “triangle” probability density distributions
    ρ(Δs,Δs)={(23Δs)(1Δs3Δs),0Δs3Δs;0,Δs<0;Δs>3Δs;
  • (b) – the negative exponential distributions ρ(Δs,Δs)=(1/Δs)exp(Δs/Δs);
  • (c) – the Gaussian distributions ρ(Δs,Δs)={(2πΔs)exp({Δs}2π{Δs}2),Δs>0;0,Δs<0..

Universality of the ratios J1(lc,Δs)/J2(lc,Δs) can be described with adequate accuracy by the power-law approximation

J2(lc,Δs){J1(lc,Δs)}γ
with the exponent γ equal to 2.26. Note that the relationship between J1(lc,Δs) and J2(lc,Δs) deviates in the case (b) from the power-law approximation (3) in the region of large Δs/lc ratios (J1(lc,Δs)<0.15;J2(lc,Δs)0.015). This is caused by sufficiently slower decay of ρ(Δs,Δs) for the negative exponential distribution in comparison with the cases (a) and (c).

The established relationships between M2,M3 (Eq. (13) for the non-polarized light or Eqs. (9) and (10) in the case of polarization discrimination of the detected multiply scattered light) and the integrands J1(lc,Δs),J2(lc,Δs) (Eqs. (14) and (16)) allows for quantification of the general tendencies relating suppression of the stochastic interference modulation of the quasi-monochromatic light in the random multiply scattering media. The suppression can be caused by broadening the emission spectrum of the probe light or/and broadening of the path length distribution for the light propagating in the medium. These relationships can be used for rapid recovery of the characteristic scales of light propagation in the medium (Δs and s) using the measured values M2 and M2 depending on the experimental conditions (e.g., the used wavelength). This approach is discussed below in sections 4 and 5.

4. Experimental data

The empirical data used in the further analysis were obtained using experimental techniques and model samples similar to those previously described in [28, 29]. Arrangement of the experiment is shown in Fig. 3. The fluorescent coarse-grained layers were pumped by the CW laser radiation at 532 nm. The layers were composed by randomly distributed coarse silica grains of irregular shape. The average size of grains was equal to 156 μm and their volume fraction in the layers was approximately equal to 0.35. Thickness of the layers L was equal to 2.0 mm. Fluorescence of the probed layers was provided by their saturation with the high-concentration Rhodamine 6G water solution (the weight fraction of Rhodamine 6G in the solution was equal to 0.005). The laser beam was expanded using the concave lens to cover the whole surface of the sample; the estimated pumping fluence rate at the sample surface was about 80 mW/cm2.

 figure: Fig. 3

Fig. 3 Arrangement of the experiment. 1 – the CW pumping laser (λ1=532 nm, the output power is 50 mW); 2 – the concave lens with the focal length of −200 mm; 3 – the sample under study; 4 – the confocal system; 5 – the pinhole diaphragm; 6 – the monochromator; 7 – the processing unit (PC). The fluorescence radiation with the wavelength λ2 is detected while scanning along the xdirection; Δz is the scan depth. Units 4 −7 are parts of the Horiba Jobin Yvon LabRam HR800 assembly.

Download Full Size | PDF

The fluorescence radiation was detected from the various points of the probed layer and from the various depths ranging from 50 μm to 200 μm with respect to the layer surface using the confocal detection scheme with the narrow-band spectral selection of the captured light. The Horiba Jobin Yvon LabRam HR800 assembly was used as the detection system; the width of the spectral window was equal to 0.052 nm and the signals were detected in the spectral range from 560 nm to 700 nm. The transverse size of the sampling volume in the confocal scheme was not larger than 2 μm. The choice of the detection zones within the layers was provided using transversal scanning of the probed samples with respect to the axis of the detection system with the scan step equal to 10 μm; the length of the scan traces at various depths was equal to 1000 μm.

The intensity acquisition time T for the chosen wavelength λ2 at the fixed detection position x was equal 0.1 s; this value significantly exceeds the parameter 2π/Δω~10−11 s, which is the measure of quasi-monochromaticity of the detected radiation. Note that the sequences of the intensity values acquired under the conditions λ2=const, x=const for the non-overlapping time intervals T exhibit the appropriately high stability in time (the normalized standard deviation σI/I of the intensity values in these sequences does not exceed 0.02).

As an example, the inset in Fig. 4 displays typical distribution of the fluorescence intensity along the arbitrarily chosen scan trace located at the depth of 150 μm below the layer surface. After collecting the raw data sets Ii,j,h,λ (i denotes the sampling point position at the scan trace j for the given scan depth h; 1i100;1j10; h=1 (50 μm), 2 (100 μm), 3 (150 μm), 4 (200 μm); λ denotes the chosen vavelength) the first-, second-, and third-order moments of spectrally-selected fluorescence intensity were calculated using the averaging procedure over the scan traces and scan depths:

Iλ=14103h=14j=110i=1100Ii,j,h,λ;Iλ2=14103h=14j=110i=1100Ii,j,h,λ2;Iλ3=14103h=14j=110i=1100Ii,j,h,λ3.
Averaging over the various scan traces (j) and scan depths (h) was applied after checking the ergodicity and statistical inhomogeneity of Ii,j,h,λ=const samples for the various used wavelengths.

 figure: Fig. 4

Fig. 4 The values of second- (M2) and third-order (M3) moments of fluorescence intensity versus the wavelength; 1 - M2; 2 - M3. Inset: the distribution of fluorescence intensity at λ= 575 nm along the arbitrarily chosen scan trace; the scan depth is 150 μm.

Download Full Size | PDF

Figure 4 displays the obtained dependencies of the normalized second- and third order moments M2(λ)=Iλ2/Iλ2, M3(λ)=Iλ3/Iλ3 on the wavelength; the error bars correspond to the confidence level 0.9. The remarkable feature is that the spectral positions of M2(λ) and M3(λ) minima are significantly red-shifted with respect to the fluorescence maxima positions varied from one detection point to another in the range from 580 nm to 610 nm. Additionally, the degree of the linear polarization PL=(IIII)/(III+I) of the outgoing fluorescence radiation was evaluated for the used detection geometry. The value PL is insufficient and does not exceed 0.05 in the whole range of the examined wavelengths (from 560 nm to 700 nm). Therefore the fluorescence light can be considered as the totally depolarized and the expressions (10) are appropriate for the follow-up analysis of intensity moments.

5. Discussion of the results

The probability density distributions ρ(Δs,Δs) of fluorescence radiation in the analyzed systems were simulated using the Monte-Carlo procedure similar to the described in [26–29]. The surface source of fluorescence photons was considered because of high absorbance of the pumping laser light by the saturating dye solution. Thus, we have the shallow CW source of fluorescence radiation in the medium; spectrally selected light waves emitted by this source interfere into the detection zone after passing through the medium along the various random paths si. We used the principle of weighted photon packets for the simulation procedure; in concordance with this principle, the single photon wandering in the probed layer is considered as the photon packet with the initial weight 1, which propagates along the same trajectory in the model layer. The model layer has the same thickness, effective refractive index, and scattering properties (the scattering coefficient μs and the reduced scattering coefficient μs) as the simulated layer, but is characterized by zero absorbance. The influence of the non-zero absorption coefficient can be taken into account at the stage of ρ(s) evaluation via summation of the unit weights of photon packets with the propagation paths within s and s+Δs(s>>Δs) with the follow-up renormalization of the sum by the Bougier factor exp(μas), where μa is the absorption coefficient of the probed system.

We used the following assumptions for the Monte-Carlo based numerical analysis of the shapes of path length distributions for detected fluorescence radiation in the examined samples:

  • - absorbance of the layers in the used spectral range in the absence of laser pumping is insufficient except the narrow spectral interval around 560 nm (the effect of residual absorption of the doping Rhodamine 6G at the long-wavelength edge of the absorption spectrum);
  • - the scattering anisotropy parameter is very large and varies between 0.90 and 0.98 (it follows from the large size of silica grains applied in the examined layers); we used the value 0.91 in the simulation;
  • - the spectrally selected fluorescence radiation is detected in the backward mode with respect to the pumping broad beam; it is taken from the depths significantly exceeding the average penetration depth of pumping radiation in the probed layers;
  • - the effective refractive index of the probed layers was taken as equal to ≈1.4; the reason is that this value is between the expected refractive index of silica grains and the index of saturating dye solution.

The shapes of the simulated path length probability density distributions ρ(s) (see inset in Fig. 5) significantly depend on the factor Lμs; in the case of Lμs>>1 the distributions exhibit the power-law behavior ρ(s)~sα in the region of small s values and exponential decay in the case of large propagation paths. The exponent α is close to 1.5 and crossover between the power-law and exponential regimes of the probability density decay occurs around the values s5 ÷ 10 mm. In general, this behavior corresponds to predictions of diffusion approximation for the temporal response of highly scattering slabs probed by short light pulses in the backscattering mode (see, e.g., [32]). The cases of “optically thin” layers with Lμs1 are featured by the expressed probability density peaks associated with fluorescence waves back-reflected from the bottom boundary of the probed layer. A strongly suppressed rudimentary peak marked by the arrow is caused by the double reflection from the bottom boundary; the peaks associated with higher-order reflections (triple, quadruple, etc.) are blurred due to path length randomness and relatively small reflectivity of the boundaries. Appearance of the single- and double-reflection features on the probability density distributions is caused by specific conditions of the fluorescence light propagation (the large scattering anisotropy and small values of Lμs) applied in the simulation procedure.

 figure: Fig. 5

Fig. 5 Distributions of the probability density of path length difference ρ(Δs); 1 - Lμs = 12; 2 - Lμs = 1.6; 3 - Lμs = 0.4; 4 - Lμs = 0.16. Inset: distributions of the path length probability density ρ(s); 1 - Lμs = 12; 2 - Lμs = 0.16.

Download Full Size | PDF

Here, we also see the exponentially decaying long-path tail related to the components of the fluorescence field undergoing multiple reflections from the layer boundaries and propagating inside the layer over distances significantly exceeding the layer thickness. Note that this feature is characteristic for the case of backward detection of the diffusing components propagating in the optically thin random layers [33].

Figure 5 displays the retrieved probability density distributions ρ(Δs) for the case of “optically thick” layer (Lμs>>1; 1) and cases of “optically thin” layers (Lμs1; 2, 3, 4). The latter cases are characterized by the exponential decay of ρ(Δs) in the wide range of Δs values. This behavior results from the relationship between ρ(s) and ρ(Δs): ρ(Δs)=20ρ(s)ρ(s+Δs)ds; in the case of the exponentially decaying path length probability density ρ(s)~exp(Ks), where K is the normalization constant related to the average path length density, and the probability density of the path length difference can be written as ρ(Δs)~exp(KΔs)0exp(2Ks)ds.

The simulation results allow us to interpret the above mentioned experimentally observed ergodicity and statistical homogeneity of intensity distributions obtained from the various scan depths in the range from 50 μm to 200 μm. These statistical properties (ergodicity and statistical homogeneity) manifest themselves in the close values M2(λ) and M3(λ) for the intensity samplings captured at the fixed wavelength λ from the different depths. Indeed, the expected average values Δs for the probed system and the used illumination and detection conditions significantly exceed the used scan depths.

Figure 6 displays the relationship between the second- and third-order statistical moments of fluorescence intensity for the examined system compared to the theoretical dependence M3(λ)=f{M2(λ)}for exponentially decaying probability densities of the path length differences and the rectangle spectral window with the fixed width. The latter dependence was retrieved using Eqs. (13) and the values J1 and J2 from the data set (b) in Fig. 2. A good agreement between the empirical and simulation data is obvious. Figure 7 shows the dependence of the average propagation path s of fluorescence radiation in the examined layers on the wavelength λ, which was recovered using the above mentioned intensity moments analysis. The theoretical values M3=f(M2), which are nearest to the average empirical values marked by the open circles in Fig. 6, were used to retrieve Δs/lc and, correspondingly, s magnitudes.

 figure: Fig. 6

Fig. 6 The values of third-order moments of fluorescence intensity versus the values of second-order moment; 1 – empirical data presented in Fig. 3; 2 – theoretical dependence for the exponential form of ρ(Δs). The ratios Δs/lc are equal to: i – 0; ii – 0.318; iii – 0.637; iv – 1.273; v – 2.546.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 The recovered values of the average path length propagation of fluorescence radiation in the examined layers versus the wavelength. 1 – the recovery using the considered intensity moments analysis; 2 – the recovery using the numerical inversion of the functional J1(lc,Δs) (in accordance with [28, 29]).

Download Full Size | PDF

It should be noted that such recovery procedure based on the a-priori knowledge of the shapes of ρ(Δs) distributions is sufficiently simpler and faster than the numerical inversion of the integral transform 0|g(Δs,lc)|2ρ(Δs)d(Δs) previously used in [27–29] and based on the inverse Monte-Carlo procedure. For reference, some s values obtained by means of this numerical inversion technique, are also plotted in this figure.

The error bars for the data set (2) correspond to the confidence level 0.9 and are related to uncertainties in the empirical data M2 and in the simulated path length distributions used for the recovery procedure. In our opinion, consistency between the two recovered data sets is reasonable; some discrepancies can be caused by a slightly inappropriate choice of the input parameters (the mean transport free path, the scattering anisotropy parameter, and the effective refractive index) in the numerical (Monte-Carlo) inversion of 0|g(Δs)|2ρ(Δs,Δs)d(Δs) in the second case.

The dotted part of the curve (1) in Fig. 7 corresponds to the propagation mode with L~s, when the shape of the path length probability distribution significantly deviates from the exponential form in the region of small s values (see inset in Fig. 5).

A remarkable feature is the sufficiently non-monotonic behavior of the recovered spectrum of s and extremely large values of the average propagation path in the spectral range from 620 nm to 680 nm. Such expressed spectral selectivity cannot be caused by the wavelength dependencies of optical properties of the non-pumped examined system (the reduced scattering coefficient, the absorption coefficient, and the scattering anisotropy). In particular, selective absorbance of the doping substance (the Rhodamine 6G solution) occurs at much shorter wavelengths).

Moreover, the abrupt increase in s in a relatively narrow spectral region under the absence of laser pumping must manifest itself as a spectrally selective optical clearing of the examined layers with expressed red coloring in the transmittance mode. Such coloring does not occur in the examined systems. A reasonable interpretation of this pump-induced spectrally selective effect is related to amplification of the stochastic fluorescence emission in the examined system under the condition of high dye concentrations. This amplification can be described by introducing the additional Bougier factor exp(μ˜as) with the negative absorption coefficient μ˜a to the path length probability density ρ(s). This factor will distort the probability density distributions due to increasing statistical weights of the long-path components. Note also that the above mentioned redshift of the minimal values of M2(λ) and M3(λ) with respect to the spectral maximum of fluorescence emission is rather caused by competition of the assumed negative absorption and residual “positive” absorption of the dye solution associated with the long-wavelength tail of the Rhodamine 6G absorption spectrum.

In our opinion, the obtained results can be useful not only for the path length interferometry of random media, but also in the analysis of speckle noise reduction in the course of quasi-monochromatic light propagation in the multiple scattering substances. Particularly, the third moment of the light intensity is associated with skewness of the probability distribution of randomly varying intensity, and its value characterizes efficiency of suppression of rare high-level intensity fluctuations in the detected optical signal.

6. Conclusions

Thus, the considered method of reference-free path length spectroscopy with the intensity moments analysis in combination with the a-priori knowledge of the expected distribution shape of path length probability allows rapid recovery of the average path length spectra for the probed random medium. In addition to high performance, the obvious advantage of this method, compared to the previously applied numerical inversion of the functional J1(lc,Δs)=0|g(Δs)|2ρ(Δs,Δs)d(Δs), is that only a minimal set of the input parameters (the second- and/or third-order moment of intensity and the coherence length of probe radiation) are necessary for the recovery of the average path length of radiation propagation in the medium. In contrast, the J1(lc,Δs) inversion using sequential numerical simulations of path length distributions in the probed medium requires a reasonable choice of optical transport parameters of the medium (the mean transport free path, the scattering anisotropy parameter, and the effective refractive index).

The considered technique can be useful for various practical applications in material science, biomedicine, laser physics, etc. It should be noted that it can be applied not only to the “frozen” random media with the suppressed scatter dynamics but also to non-stationary scattering systems. The only limitation is that the correlation time of intensity fluctuations caused by the scatter motions should be much larger than the intensity sampling time at a given sample position.

Funding

Russian Science Foundation (RSF) (16-19-10455); Ministry of Education and Science of the Russian Federation (3.7567.2017).

Acknowledgments

The authors are grateful to I.A. Asharchuk and A.P. Sviridov for collaboration in obtaining the raw experimental data. The authors are also grateful to RSF for support of the experimental part of this research. D.A. Zimnyakov thanks the Ministry of Education and Science of the Russian Federation for the support of this work in the part of theoretical modeling.

References and links

1. A. Ishimaru, Wave Propagation and Scattering in Random Media (Academic, 1978).

2. A. Ishimaru and Y. Kuga, “Attenuation constant of a coherent field in a dense distribution of particles,” J. Opt. Soc. Am. 72(10), 1317–1320 (1982). [CrossRef]  

3. F. C. MacKintosh and S. John, “Diffusing-wave spectroscopy and multiple scattering of light in correlated random media,” Phys. Rev. B Condens. Matter 40(4), 2383–2406 (1989). [CrossRef]   [PubMed]  

4. M. P. V. Albada, A. Lagendijk, and A. Lagendijk, “Observation of weak localization of light in a random medium,” Phys. Rev. Lett. 55(24), 2692–2695 (1985). [CrossRef]   [PubMed]  

5. M. C. W. van Rossum and Th. M. Nieuwenhuizen, “Multiple scattering of classical waves: microscopy, mesoscopy, and diffusion,” Rev. Mod. Phys. 71(1), 313–371 (1999). [CrossRef]  

6. D. S. Wiersma, M. P. V. Albada, A. Lagendijk, B. A. van Tiggelen, and A. Lagendijk, “Experimental evidence for recurrent multiple scattering events of light in disordered media,” Phys. Rev. Lett. 74(21), 4193–4196 (1995). [CrossRef]   [PubMed]  

7. A. A. Middleton and D. S. Fisher, “Discrete scatterers and autocorrelations of multiply scattered light,” Phys. Rev. B Condens. Matter 43(7), 5934–5938 (1991). [CrossRef]   [PubMed]  

8. G. Maret and P. E. Wolf, “Multiple light scattering from disordered media. The effect of brownian motion of scatterers,” Zeitschrift für Physik B 65(4), 409–413 (1987). [CrossRef]  

9. D. J. Pine, D. A. Weitz, P. M. Chaikin, and E. Herbolzheimer, “Diffusing wave spectroscopy,” Phys. Rev. Lett. 60(12), 1134–1137 (1988). [CrossRef]   [PubMed]  

10. G. Dietsche, M. Ninck, C. Ortolf, J. Li, F. Jaillon, and T. Gisler, “Fiber-based multispeckle detection for time-resolved diffusing-wave spectroscopy: characterization and application to blood flow detection in deep tissue,” Appl. Opt. 46(35), 8506–8514 (2007). [CrossRef]   [PubMed]  

11. D. Wang, A. B. Parthasarathy, W. B. Baker, K. Gannon, V. Kavuri, T. Ko, S. Schenkel, Z. Li, Z. Li, M. T. Mullen, J. A. Detre, and A. G. Yodh, “Fast blood flow monitoring in deep tissues with real-time software correlators,” Biomed. Opt. Express 7(3), 776–797 (2016). [CrossRef]   [PubMed]  

12. Y. Shang, Y. Zhao, R. Cheng, L. Dong, D. Irwin, and G. Yu, “Portable optical tissue flow oximeter based on diffuse correlation spectroscopy,” Opt. Lett. 34(22), 3556–3558 (2009). [CrossRef]   [PubMed]  

13. Y. Shang, T. B. Symons, T. Durduran, A. G. Yodh, and G. Yu, “Effects of muscle fiber motion on diffuse correlation spectroscopy blood flow measurements during exercise,” Biomed. Opt. Express 1(2), 500–511 (2010). [CrossRef]   [PubMed]  

14. L. Brunel, A. Brun, P. Snabre, and L. Cipelletti, “Adaptive Speckle Imaging Interferometry: a new technique for the analysis of microstructure dynamics, drying processes and coating formation,” Opt. Express 15(23), 15250–15259 (2007). [CrossRef]   [PubMed]  

15. F. Scheffold, S. Romer, F. Cardinaux, H. Bissig, A. Stradner, L. F. Rojas-Ochoa, V. Trappe, C. Urban, S. E. Skipetrov, L. Cipelletti, and P. Schurtenberger, “New trends in optical microrheology of complex fluids and gels,” Prog. Colloid Polym. Sci. 123, 141–146 (2004).

16. N. Menon and D. J. Durian, “Diffusing-wave spectroscopy of dynamics in a three-dimensional granular flow,” Science 275(5308), 1920–1922 (1997). [CrossRef]   [PubMed]  

17. V. Viasnoff, F. Lequeux, and D. J. Pine, “Multispeckle diffusing-wave spectroscopy: A tool to study slow relaxation and time-dependent dynamics,” Rev. Sci. Instrum. 73(6), 2336–2344 (2002). [CrossRef]  

18. P. Zakharov, F. Cardinaux, and F. Scheffold, “Multispeckle diffusing-wave spectroscopy with a single-mode detection scheme,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 73(1), 011413 (2006). [CrossRef]   [PubMed]  

19. J. Y. Lee, J. W. Hwang, H. W. Jung, S. H. Kim, S. J. Lee, K. Yoon, and D. A. Weitz, “Fast dynamics and relaxation of colloidal drops during the drying process using multispeckle diffusing wave spectroscopy,” Langmuir 29(3), 861–866 (2013). [CrossRef]   [PubMed]  

20. O. V. Angelsky and P. P. Maksimyak, “The investigation of the transformation phenomenon of the longitudinal correlation function of the field propagating in the light scattering medium,” Opt. Spectrosc. 60(2), 331–336 (1986).

21. O. V. Angelsky, A. P. Maksimyak, P. P. Maksimyak, and S. G. Hanson, “Optical correlation diagnostics of rough surfaces with large surface inhomogeneities,” Opt. Express 14(16), 7299–7311 (2006). [CrossRef]   [PubMed]  

22. J. W. Goodman, Statistical Optics (Wiley Classics Library Edition, 2000).

23. T. M. Nieuwenhuizen and M. C. W. van Rossum, “Intensity distributions of waves transmitted through a multiple scattering medium,” Phys. Rev. Lett. 74(14), 2674–2677 (1995). [CrossRef]   [PubMed]  

24. E. Kogan and M. Kaveh, “Random-matrix-theory approach to the intensity distributions of waves propagating in a random medium,” Phys. Rev. B Condens. Matter 52(6), R3813–R3815 (1995). [CrossRef]   [PubMed]  

25. C. A. Thompson, K. J. Webb, and A. M. Weiner, “Imaging in scattering media by use of laser speckle,” J. Opt. Soc. Am. A 14(9), 2269–2277 (1997). [CrossRef]  

26. D. A. Zimnyakov, J.-T. Oh, Y. P. Sinichkin, V. A. Trifonov, and E. V. Gurianov, “Polarization-sensitive speckle spectroscopy of random media beyond the diffusion limit,” J. Opt. Soc. Am. A 21(1), 59–70 (2004). [CrossRef]  

27. D. A. Zimnyakov and M. A. Vilensky, “Blink speckle spectroscopy of scattering media,” Opt. Lett. 31(4), 429–431 (2006). [CrossRef]   [PubMed]  

28. D. A. Zimnyakov, I. A. Asharchuk, S. A. Yuvchenko, and A. P. Sviridov, “Stochastic interference of fluorescence radiation in random media with large inhomogeneities,” Opt. Commun. 387, 121–127 (2017). [CrossRef]  

29. D. A. Zimnyakov, I. A. Asharchuk, S. A. Yuvchenko, and A. P. Sviridov, “Speckle spectroscopy of fluorescent randomly inhomogeneous media,” Quantum Electron. 46(11), 1047–1054 (2016). [CrossRef]  

30. D. A. Zimnyakov, V. V. Tuchin, and A. G. Yodh, “Characteristic scales of optical field depolarization and decorrelation for multiple scattering media and tissues,” J. Biomed. Opt. 4(1), 157–163 (1999). [CrossRef]   [PubMed]  

31. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University Press, 1995).

32. P. M. Johnson, A. Imhof, B. P. J. Bret, J. G. Rivas, and A. Lagendijk, “Time-resolved pulse propagation in a strongly scattering material,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 68(1), 016604 (2003). [CrossRef]   [PubMed]  

33. D. A. Zimnyakov, J. S. Sina, S. A. Yuvchenko, E. A. Isaeva, S. P. Chekmasov, and O. V. Ushakova, “Low-coherence interferometry as a method for assessing the transport parameters in randomly inhomogeneous media,” Quantum Electron. 44(1), 59–64 (2014). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 The simulated distributions ρ( I/ I ) in the cases of the deterministic value | E 0i |=1 (1, 3) and random uniform distribution of the phasor amplitudes (2, 4). The number of statistically independent phasors used in the simulation procedure is equal to 500; the sample size applied for the reconstruction of ρ( I/ I ) is equal to 10000. The cases (1, 2) correspond to coherent summation of the random phasors ( s i s m / l c = 1.0⋅10−3) and the cases (3, 4) correspond to the partially coherent summation ( s i s m / l c = 1.0).
Fig. 2
Fig. 2 The plot of values J 1 ( l c , Δs ) and J 2 ( l c , Δs ) for the various shapes of ρ( Δs, Δs ) ; a – open triangles; b – closed circles; c – open circles. The Δs / l c ratios are equal to: i – 0; ii – 0.318; iii – 0.637; iv – 0.955; v – 1.273; vi – 1.592; vii – 2.228; viii – 3.183. Dashed line – the power-law approximation of the relationship between J 1 ( l c , Δs ) and J 2 ( l c , Δs ) .
Fig. 3
Fig. 3 Arrangement of the experiment. 1 – the CW pumping laser ( λ 1 = 532 nm, the output power is 50 mW); 2 – the concave lens with the focal length of −200 mm; 3 – the sample under study; 4 – the confocal system; 5 – the pinhole diaphragm; 6 – the monochromator; 7 – the processing unit (PC). The fluorescence radiation with the wavelength λ 2 is detected while scanning along the xdirection; Δz is the scan depth. Units 4 −7 are parts of the Horiba Jobin Yvon LabRam HR800 assembly.
Fig. 4
Fig. 4 The values of second- ( M 2 ) and third-order ( M 3 ) moments of fluorescence intensity versus the wavelength; 1 - M 2 ; 2 - M 3 . Inset: the distribution of fluorescence intensity at λ= 575 nm along the arbitrarily chosen scan trace; the scan depth is 150 μm.
Fig. 5
Fig. 5 Distributions of the probability density of path length difference ρ( Δs ) ; 1 - L μ s = 12; 2 - L μ s = 1.6; 3 - L μ s = 0.4; 4 - L μ s = 0.16. Inset: distributions of the path length probability density ρ( s ) ; 1 - L μ s = 12; 2 - L μ s = 0.16.
Fig. 6
Fig. 6 The values of third-order moments of fluorescence intensity versus the values of second-order moment; 1 – empirical data presented in Fig. 3; 2 – theoretical dependence for the exponential form of ρ( Δs ) . The ratios Δs / l c are equal to: i – 0; ii – 0.318; iii – 0.637; iv – 1.273; v – 2.546.
Fig. 7
Fig. 7 The recovered values of the average path length propagation of fluorescence radiation in the examined layers versus the wavelength. 1 – the recovery using the considered intensity moments analysis; 2 – the recovery using the numerical inversion of the functional J 1 ( l c , Δs ) (in accordance with [28, 29]).

Equations (22)

Equations on this page are rendered with MathJax. Learn more.

I( t ) ¯ = E( t ) E * ( t ) ¯ = | E( t ) | 2 ¯ = | i N E i ( t )exp{ j φ i ( t ) } | 2 ¯ ,
I( t ) ¯ = i N i N E i ( t ) E i ( t )exp[ j{ φ i ( t ) φ i ( t ) } ] ¯ = = i N i N I i ( t ) I i ( t ) exp[ j{ φ i ( t ) φ i ( t ) } ] ¯ .
I( t ) ¯ = i N I i ( t ) ¯ + i N i i N I i ( t ) I i ( t ) cos[ k( s i s i )+Δϕ( t, s i s i ) ] ¯ ,
I i ( t ) I i ( t ) cos[ k( s i s i )+Δϕ( t, s i s i ) ] ¯ = = I i ( t ) ¯ I i ( t ) ¯ cos[ k( s i s i )+Δϕ( t, s i s i ) ] ¯ = I i ( t ) ¯ I i ( t ) ¯ cos[ Δϕ( t, s i s i ) ] ¯ cos[ k( s i s i ) ] I i ( t ) ¯ I i ( t ) ¯ sin[ Δϕ( t, s i s i ) ] ¯ sin[ k( s i s i ) ]= = I i ( t ) ¯ I i ( t ) ¯ Q 2 ( s i s i )+ S 2 ( s i s i ) cos[ k( s i s i )Ψ( s i s i ) ],
Q( s i s i )= 1 T 0 T cos[ Δϕ( t, s i s i ) ]dt; S( s i s i )= 1 T 0 T sin[ Δϕ( t, s i s i ) ]dt ; tg{ Ψ( s i s i ) }= S( s i s i ) / Q( s i s i ) ,
g( s i s i )= sin{ π( s i s i ) / l c } / { π( s i s i ) / l c } exp{ jk( s i s i ) },
I= i N E 0i 2 + i N i i N E 0i E 0 i cos{ k( s i s i ) }| g( s i s i ) | .
I 2 = ( i N E 0i 2 + i N i i N E 0i E 0 i cos{ k( s i s i ) }| g( s i s i ) | ) 2 ,
I 3 = ( i N E 0i 2 + i N i i N E 0i E 0 i cos{ k( s i s i ) }| g( s i s i ) | ) 3
I 2 = ( N+2 m N( N1 ) /2 cos( kΔ s m )| g( Δ s m ) | ) 2 ,
I 3 = ( N+2 m N( N1 ) /2 cos( kΔ s m )| g( Δ s m ) | ) 3 ,
I 2 = N 2 +4N m N( N1 ) /2 cos( kΔ s m )| g( Δ s m ) | + +4 ( m N( N1 ) /2 cos( kΔ s m )| g( Δ s m ) | ) 2 , I 3 = N 3 +6 N 2 m N( N1 ) /2 cos( kΔ s m )| g( Δ s m ) | + +12N ( m N( N1 ) /2 cos( kΔ s m )| g( Δ s m ) | ) 2 + +8 ( m N( N1 ) /2 cos( kΔ s m )| g( Δ s m ) | ) 3 .
M 2 = I 2 I 2 =1+ N1 N 0 | g( Δs ) | 2 ρ( Δs, Δs )d( Δs ) 1+ 0 | g( Δs ) | 2 ρ( Δs, Δs )d( Δs ) .
M 3 = I 3 I 3 =1+3 N1 N 0 | g( Δs ) | 2 ρ( Δs, Δs )d( Δs ) + +2 ( N1 ) 2 N 2 ( 0 | g( Δs ) |ρ( Δs, Δs )d( Δs ) ) 3 1+3 0 | g( Δs ) | 2 ρ( Δs, Δs )d( Δs ) +2 ( 0 | g( Δs ) |ρ( Δs, Δs )d( Δs ) ) 3 .
ρ( Δs, Δs )={ 2 0 ρ( s+Δs )ρ( s )ds ,Δs0; 0,Δs<0,
I =2 I ,II ; I 2 = ( I + I II ) 2 = I 2 +2 I I II + I II 2 ; I 3 = ( I + I II ) 3 = I 3 +3 I 2 I II +3 I I II 2 + I II 3 ; M 2 = I 2 I 2 = 2 I ,II 2 +2 I ,II 2 4 I ,II 2 = 1 2 I ,II 2 I ,II 2 + 1 2 ; M 3 = I 3 I 3 = 2 I ,II 3 +6 I ,II 2 I ,II 8 I ,II 3 = 1 4 I ,II 3 I ,II 3 + 3 4 I ,II 2 I ,II 2 .
{ M 2 = I 2 I 2 1+ 1 2 0 | g( Δs ) | 2 ρ( Δs, Δs )d( Δs ) ; M 3 = I 3 I 3 1+ 3 2 0 | g( Δs ) | 2 ρ( Δs, Δs )d( Δs ) + + 1 2 ( 0 | g( Δs ) |ρ( Δs, Δs )d( Δs ) ) 3 .
J 1 ( l c , Δs )= 0 | sin( πΔs / l c ) πΔs / l c | 2 ρ( Δs, Δs )d( Δs ) ; J 2 ( l c , Δs )= ( 0 | sin( πΔs / l c ) πΔs / l c |ρ( Δs, Δs )d( Δs ) ) 3
Δs l c =0 J 1 ( l c , Δs )= J 2 ( l c , Δs )=1; Δs l c = J 1 ( l c , Δs )= J 2 ( l c , Δs )=0
ρ( Δs, Δs )={ ( 2 3 Δs )( 1 Δs 3 Δs ),0Δs3 Δs ; 0,Δs<0;Δs>3 Δs ;
J 2 ( l c , Δs ) { J 1 ( l c , Δs ) } γ
I λ = 1 4 10 3 h=1 4 j=1 10 i=1 100 I i,j,h,λ ; I λ 2 = 1 4 10 3 h=1 4 j=1 10 i=1 100 I i,j,h,λ 2 ; I λ 3 = 1 4 10 3 h=1 4 j=1 10 i=1 100 I i,j,h,λ 3 .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.