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

Fractal and spinodal-decomposed turbidities of nanoporous glass: fluctuation picture in turbid and transparent Vycor

Open Access Open Access

Abstract

The light propagation and scattering in monolithic transparent nanoporous materials such as Vycor glasses exhibit two optical turbidities, both of which are slightly deviated from the λ4 Rayleigh wavelength dependence in the visible region: one is a transient white turbidity τf, characterized by the convex-upward dependence on the inverse fourth power of wavelength, and the other is turbidity τsp inherent to the structural inhomogeneity, characterized by the convex-downward dependence. The former is attributed to a fractal-like percolation network of imbibed or drained pores as a consequence of transient imbibition or drainage of wetting fluid into or from the pore space. The latter is attributed to the structural inhomogeneities inherent to the original dry porous glass, which are produced by spinodal decomposition. In this paper, we develop a general scheme to estimate the transmittance spectra of Vycor through the turbidities τf and τsp in the visible region on the basis of the theory of dielectric constant fluctuations. We show the applicability and its limitation of the turbidity analysis of the photospectroscopically measured data as a method to study the correlation functions that characterize the pore space and the structural features of isotropic transparent nanoporous media, on the presupposition that there exists no light attenuation other than the scattering.

© 2017 Optical Society of America

1. INTRODUCTION

Porous Vycor glass with nanopores is transparent in the visible region and is used in colorimetric chemical gas sensing when impregnated with selectively reacting regents [14]. The pores in Vycor have an average diameter of about 4.2 nm and fill about 0.3 of the total volume of the material. Vycor’s large internal surface area of about 200m2/g and its transparency are very attractive for gas sensing by adsorption. This is because the larger the area the higher the adsorption, resulting in higher sensitivity. In addition, optical methods [5] have an advantage over conventional ones in that they can distinguish gas species by chemically detecting the inherent optical absorption.

However, they have some disadvantages in sensing, since it turns out that changes in the humidity of ambient air strongly affect the transmittance of porous glass-based sensor elements [6,7]. In the transmittance measurement, we measure the attenuation or extinction of the light emerging from the sensor element impregnated with a proper chemical reagent. Unfortunately, this extinction is the result of both absorption and scattering. The cumulative absorbance changes at a specified wavelength in the visible region certainly originate from the absorption due to selective reactions of impregnated reagent with the target gas. But the scattering is originated from the optical inhomogeneities occurring in the pore space of the porous glass matrix [8]. The transitory turbidity is found to occur even for the elements without reagents and is a problem common to all the elements that use porous glass as a base material. In our previous studies [810], we have shown that this transitory whitish turbidity phenomenon can be attributed to Rayleigh scattering caused by the spatial fluctuation (inhomogeneity) of constituent materials within the glass, i.e., the distribution of wetting fluid such as water vapor in the pore space.

Given this situation, there is considerable value in modeling and understanding the cause of turbidity in inhomogeneous nanoporous media such as Vycor. The study of the optical properties of porous glasses, in particular light attenuation in scattering and absorption, is important for correcting the unnecessary contribution from the transitory white turbidity from the porous matrix. A numerical model to account for the turbidity is required to precisely estimate the concentrations from the cumulative absorbance change in the porous glass-based sensor element.

For a long time, porous Vycor glass had been used an intermediate product for obtaining high-silica glass (96–98%) [11], which implies that its silica skeleton is highly pure in composition. Since silica glass is a dielectric insulator, it is transparent in the near ultraviolet and the visible regions because the electrons of the atomic shells are bound to the individual atom. Its optical properties are determined by its complex refractive index, n˜(ω)=n(ω)ik(ω), where n is the refractive index and k is the extinction coefficient. For wavelengths in the range 3.70μm>λ>0.21μm, the refractive index n of silicon dioxide (SiO2) (glass) is almost constant around the value of 1.45, and no k values are given in this wavelength range [12], which implies that no absorption is observed for pure SiO2 glass [12,13]. Its electronic structure [14] indicates that the energy gap of about 8.9 eV between the conduction and valence bands suppresses the electronic excitations by absorbing visible light photons. Thus, the classical electromagnetic theory is sufficient for our purpose.

In general, attenuation (extinction) or power loss, as light travels a distance dopt through any optical materials, can be measured in terms of turbidity τext=dopt1ln[I(0)/I(dopt)] with the dimension of reciprocal length (usually μm1), where I(0) is the incident light intensity and I(dopt) is the intensity of the light beam emerging from the sample. As the sources of attenuation in optically transparent materials, loss mechanisms fall into two categories, one involving the slight absorption of radiation τabs (e.g., by reagent molecules in the listed sensor elements with light energy causing electronic transitions in the visible region), and the other mainly concerned with scattering τsca, in which radiation is diverted (with or without change of frequency) from its primary propagating path. The law of conservation of energy then requires that τext=τabs+τsca.

The sources of optical loss in glass are of three types: (i) intrinsic absorption in the ultraviolet region and molecular vibrations in the infrared region [15], (ii) impurities such as transition elements which causes colors [15], and (iii) scattering caused by bubbles, phase separation, and/or density fluctuations [16,17]. Among these, the optical properties of porous glasses in the visible region are mainly determined by scattering [18,19] because they are structurally inhomogeneous due to the presence of the innumerable pores and surrounding silica skeleton, and compositionally inhomogeneous due to the nonuniform distribution of the imbibed or drained wetting fluid such as water in the pore space during imbibition or drainage of fluid.

In this paper, we will treat the optical properties of porous Vycor glass in the view of scattering caused by structural and compositional inhomogeneities, both of which might scatter light strongly.

Structural studies by means of small-angle x-ray and neutron scattering (SAXS and SANS) on Vycor [2036] have shown that nanoporous Vycor glass has a pronounced inhomogeneous structure. The inhomogeneities of the refractive index observed in nanoporous Vycor glass can be conveniently divided into two classes: Inhomogeneities of type I and those of type II. The former are due to the inhomogeneity of the original leached borosilicate glass structure itself and arise in the process of glass manufacturing—the so-called spinodal phase separation and decomposition process. Several SAXS and SANS measurements [2123,25,2636] have clearly shown that the structure of the pore space of Vycor is reminiscent of spinodal decomposition, with a broad peak centered around q=0.25nm1 (where q is the scattering vector) in the structure factor corresponding to a length of about 20–30 nm. This kind of inhomogeneity should be called “physical inhomogeneity,” as distinguished from the type-II inhomogeneity.

Inhomogeneities of type II are due to local deviations from the mean glass composition with long-range correlation in the pore space [810,37,38], that is, “chemical inhomogeneity.” The porous matrix medium itself is static and rigid, but the distribution of wetting fluid imbibed into pores randomly varies in the pore space, so that the amplitude and phase of the light waves traveling through the whole medium might fluctuate randomly in space, which, subsequently, would change the transparency of the whole medium. Several SANS studies [2729] have shown that, upon the filling or draining of pores, incomplete, partially saturated Vycor can exhibit fractal properties via the formation of a fractal-like percolation network of imbibed or drained pores in the pore space. In this sense, to clarify the turbidity problem in nanocomposite materials, it is very important to characterize properties of soft materials in hard confinement, being rather general in the characterization of the nanomaterials.

In our previous studies [810], we have shown that the transitory whitish turbidity phenomenon occurs in nanoporous Vycor glass that was partially saturated with a wetting fluid such as water, that is, the light scattering caused by the inhomogeneities of type II. The phenomenon can be consistently interpreted and quantitatively analyzed by a simple Rayleigh scattering model [810], which is based on the particle-scattering approach [39]. In our subsequent study [40], we attributed the phenomenon to the fractal-like percolation network of drained pores, on the basis of the theory of dielectric constant fluctuations [41,42] with the previously reported experimental observations [2729,37,38]. In the study, it was very important to explain the slight deviation of the observed turbidity from the λ4 Rayleigh wavelength dependence, which cannot be explained at all from the previous particle-scattering viewpoint [39].

This paper is organized as follows. The next section briefly describes the theoretical scheme to analyze the transmittance spectrum in terms of turbidities on the basis of the theory of dielectric constant fluctuations. In Section 2.A, we develop mathematical expressions for a correlation volume with experimentally observed value of the fractal dimension D=5/2 (for details for values other than 5/2, see Appendix B). In Section 2.B, the same fluctuation analysis scheme is applied to the case for the structural inhomogeneity caused by the spinodal decomposition. For both cases, we derived the deviation from the λ4 wavelength dependence of corresponding turbidities explicitly. In the subsequent sections, these derived expressions are compared with the previously obtained experimental results, and then we offer a possible explanation of the transmittance spectra of dried and completely saturated Vycor glass and that of incompletely saturated Vycor glass in terms of the previously mentioned spinodal-decomposed and fractal turbidities, respectively. Finally, Section 5 concludes the paper. The detailed deviation of the basic equations for the turbidity analysis based on the classical electromagnetic theory is available in Appendix A, and the detailed mathematical expressions for correlation volumes with half-integer and integral values of the fractal dimension D are available in Appendix B.

2. THEORETICAL EXPRESSIONS

As is well known, the problem of light scattering can be approached from Einstein’s fluctuation viewpoint [41,42], apart from the previously mentioned particle-scattering one. This approach can treat the problem more systematically and comprehensively by considering the correlation function that characterizes both classes of inhomogeneities. That is, in the fluctuation approach, the inhomogeneity present in a condensed medium is generally characterized by introducing the concept of a spatial correlation function, in which the inhomogeneity of the medium is considered to be due to a continuous variation of the dielectric constant or refractive index rather than to the presence of discrete particles. The spatial correlation function is defined as γ(r)ηAηB/η2AV, where ηA and ηB are the local fluctuations in dielectric constant from some average value εAV at points A and B a distance r apart, and η2AV is the average value of η2. The problem of characterizing the inhomogeneities in solids and relating them to the scattered intensity has been treated by Debye et al. [43,44]. Their approach has been utilized both for SAXS and conventional light scattering.

Their mathematical scheme is summarized as follows: for an unpolarized incident beam of light, the scattered intensity for an isotropic system is simply proportional to ω(q), that is, the Fourier transform of γ(r), namely,

I(q)η2AVπ2λ04Vω(q),
where V is the scattering volume and λ0 is the wavelength in vacuo. [For the detailed derivation of Eq. (1), including the proportional constant, see Appendix A, Eq. (A5)]. Here, the quantity ω(q) is called the correlation volume, which is a function of both γ(r) and q and is defined as
ω(q)=04πr2γ(r)sin(qr)qrdr,
with the scattering vector q=2ksin(θ/2), where θ is the scattering angle, defined as the angle between the incident and scattered beams; k2πn/λ0 is the wave number; and n is the refractive index of the medium. The corresponding turbidity τ takes the form [43,44]
τ=η2AVπ2λ044πdΩ1+cos2θ2ω(q),
where the scattering intensity, Eq. (1), has been summed over all solid angles dΩ(θ,φ) with all scattering angles θ.

In all scattering determinations of the inhomogeneities of Vycor [2036], Eq. (1) implies that, to obtain the correlation function via Eq. (2), the angular dependence of the scattered intensity [which is proportional to the correlation volume ω(q)] from the sample with a monochromatic incident radiation beam is measured as a function of the scattering vector q.

On the other hand, in this paper, we re-examine our transmittance spectrum of nanoporous Vycor glass from the fluctuation viewpoint and show that the photospectroscopically observed forward-scattered intensity from Vycor glass provides the same information as that obtained from the scattering intensity of monochromatic incident light at small angle θ, for which the theoretical scheme was originally proposed by Debye et al. [43,44]. To discuss the slight deviation of the observed turbidity from the λ4 Rayleigh wavelength dependence, we derive the explicit formula for the turbidity as a function of the incident wavelength corresponding to the assumed correlation functions. In the following section we outline some equations we have developed for analyzing scattering data in terms of fractal scattering and the corresponding turbidity.

A. Fractal Correlation Function and Corresponding Turbidity

As mentioned previously, when porous Vycor glass is partially saturated with a wetting fluid, a fractal-like percolation network might form [10,2729,37,38], although the pores themselves are not distributed with fractal geometry. As the draining proceeds, the transmittance becomes small due to strong scattering of light. Taking into account both the finite size and the fractal-like percolation network structure of drained pores, Sinha, Freltoft, and Kjems [4547] proposed the fractal correlation function γf(r) of the form

γf(r)=CfrdDexp(rξ),
where Cf is a proportional constant and D is the dimensional number of the fractal dimensionality smaller than the corresponding Euclidean dimensionality, d. Here, the parameter ξ, the correlation length, characterizes the average distance between fluctuations in the dielectric constant. Substituting the expression into Eq. (2) and carrying out the Fourier transform for d=3 yields the correlation volume due to the fractal correlation of the form [48]
ωf(q)=Cf4πξDΓ(D1)[1+q2ξ2]D/2[1+q2ξ2(qξ)2]1/2sin[(D1)tan1(qξ)],
where Γ(α) is the gamma function of nonintegral argument α. With this expression of the correlation volume ωf(q), we can easily obtain the scattering angular dependence of the scattered light intensity I(q) as I(q)η2AVV·ωf(q), which is sufficient for analyzing the detailed structural data from the laser light scattering, SAXS, and SANS measurements with monochromatic probing radiation.

On the other hand, in the case of a spectrophotometer with a varying wavelength of the incident light, we have to analyze the turbidity of the forward-scattered light to derive any structural information about the wavelength dependence of the transmittance spectra [40]. Unfortunately, there is no closed formula for the integral of the correlation volume with any noninteger value of D in Eq. (5) over the entire solid angle dΩ. As a consequence, we have to integrate the correlation volume in Eq. (5) only for specified values of D, for which the analytical expression of the corresponding turbidity is obtainable.

Several scattering determinations of the inhomogeneities of Vycor during drainage of a wetting fluid [37,38] can be found in the literature. By means of light scattering, which is a definite probe of the long-range spatial correlations of the empty pores, Page et al. [37,38] reported that the draining pores, which are surrounded by filled pores, exhibits long-range correlations of a fractal dimension of about 2.6.

Inserting the correlation function of Eq. (4) for d=3 with analytically treatable fractal dimension (D) of 2.5 into Eq. (2), the correlation volume of Eq. (5) is given as ωf(q)=ω0(5/2)·f5/2(q), with a constant ω0(5/2)=4πξ5/2Γ(5/2)=3π3/2ξ5/2 and the dissymmetry factor

f5/2(q)=23[1+q2ξ21]1/2[1+q2ξ2+2]qξ[1+q2ξ2]3/2.

Substituting this expression into Eq. (3) and integrating over all angles, the corresponding turbidity τf(D=5/2) is given as

τf=ηf2AVk4(2π)2εAV2ω0(5/2)·Ω5/2,
where
Ω5/2=2π0πdθsinθ1+cos2θ2f5/2(2ksin(θ2))=162π3b3{h(1+1+b,b)h(2,b)},
with b4k2ξ2 and
h(z,b)=2z{z37z25z3(2+b)(z3+1)1+b+b2/2z1}.

Equation (7) implies that the wavelength dependence of the turbidity τf due to the fractal-like percolation network of drained pores is determined mainly by 1/λ04 [which comes from the factor k4/(16π2εAV2)] and the integrated correlation volume ω0(5/2)·Ω5/2, which is a function of ξ only and determines the slight deviation from the λ4 Rayleigh wavelength dependence. We hereafer refer to τf as “fractal turbidity.”

Figure 1(a) shows the dependence of the anisotropic correlation volume ωf(q) on the scattering wave vector q, illustrating the fractal correlations of half-integer and integer dimensions D. For the sake of comparison with the experimental data represented by open circles, which are reproduced from [37,38], the correlation length ξ is temporally set equal to 10 μm. It turns out that the curve with the fractal dimension of 5/2 (solid line) well approximates the slope of the reported data. Figure 1(b) shows that the corresponding integrated dissymmetry factors ΩD of the fractal dimension of D are convex-upward functions of 1/λ04. Here, the correlation length ξ is set to about 18.5 nm. This value of ξ is chosen to fit our observed data, as discussed in the next section. This λ4 wavelength dependence makes the turbidity τf due to a fractal-like percolation network of drained pores to also be a convex-upward function of 1/λ04, since τf is proportional to the integrated factor ΩD as in Eq. (7).

 figure: Fig. 1.

Fig. 1. (a) Dependence of anisotropic correlation volume ωf(q) on scattering wave vector q, illustrating the fractal correlations of half and integral numbers of fractal dimension D. The curve with D=5/2 well approximates the slope of experimental data depicted by open circles, which represent scattered light intensity as a function of q with a monochromatic incident light from a laser source (514.2 nm) [37,38]. This comparison is justified by considering the fact that ωf(q) is proportional to the scattered intensity. For this comparison with data in [37,38], we have to intentionally set the correlation length (ξ) equal to 10 μm, which is about 103 times longer than our estimated value [40]. (b) Dependence of integrated dissymmetry factors (ΩD) for various values of D, where ΩD is proportional to the fractal turbidity τf on 1/λ04, indicating both τf and ΩD are convex-upward functions of 1/λ04.

Download Full Size | PDF

A wetting fluid such as water in the pores of Vycor glass is reported to have a fractal structure with a fractal dimension of 2.6 [37,38], which will be well approximated by the given expression with D=5/2, while the matrix skeleton of silica glass does not have a fractal structure [21,22]. Next, we will discuss the turbidity inherent to the structural inhomogeneities, such as spindal decomposition, in the following subsection.

B. Spinodal-Decomposition Correlation Function and Corresponding Turbidity

Leached porous glass such as Vycor is produced by a spinodal phase separation and leaching process [11] in which sodium borosilicate glass is heat-treated below the liquidus temperature to induce separation into continuous phases. The Na2O-B2O3-rich phase is then leached out with hydrochloric acid. Nanoporous Vycor glass has a narrow pore-size distribution, an average nominal pore radius of 2.0 nm, a porosity φ of 0.28–0.3, and an internal surface area of 200m2/g. Several SAXS and SANS measurements have shown that the structure of this pore space is reminiscent of spinodal decomposition, with a broad peak centered around 0.3 [21], 0.23 [22,36], 0.22 [25], 0.20 [23,32], and 0.25 [2629,30,31,3335] nm1 in the structure factor, corresponding to a periodicity of 20–30 nm.

The reported pair correlation function inherent to the spinodal decomposition is [28]

g(r)1=A·sin(βmr)βmrexp(r/ζ),
which determines the microscopic structure of porous glass. Here βm is the maximum wave number at the spinodal decomposition, which determines the fine structure of this porous material. The SAXS and SANS structure analysis tells us that the value of βm is 0.25nm1 [2629,30,31,3335]. On the other hand, ζ is the characteristic length of the exponential decay of the correlation, that is, the correlation length, which is determined from model fitting to be ζ=10.2nm [28]. Both parameters have physical significance in that parameter βm is associated with the position of the peak in the structure factor and the correlation length ζ determined by the width at half-height.

Now, we assume that the local dielectric fluctuation is driven by a local density fluctuation and that the dielectric correlation function η(r1)η(r2) is proportional to the density–density correlation function g(r)1 [16,17]. Therefore, we can express the correlation function, which characterizes the dielectric fluctuation due to the spinodal decomposition, in the form

γsp(r)=Cspsin(βr)βrexp(r/ζ),
where Csp is a proportional constant, and β and ζ are the same parameters inherent to Eq. (10). Substituting Eq. (11) into Eq. (2) with Csp=1, the correlation volume ωsp(q) (i.e., the factor expressing the change of scattered light intensity as a function of the scattering vector q) is given by the Fourier transform of γsp(r) as follows:
ωsp(q)=8πζ3q4ξ42(1β2ζ2)q2ζ2+(1+β2ζ2)2,
from which the corresponding turbidity, τsp, is obtained as
τsp=ηsp2AVk4(4π)2εAV2·Ωsp,
where
Ωsp=2π0πdθsinθ1+cos2θ2ωsp(2ksin(θ2))=(8πb)2ζ3[1(12+1ab)ln{1+b2+2b(1a)(1+a)2}+b8a{1+(1+2(1b)b)216ab2}{tan1(1a+b2a)tan1(1a2a)}],
where a=β2ζ2 and b=4k2ζ2. We hereafter refer to τsp as “spinodal-decomposed turbidity,” or briefly as “spinodal turbidity.”

By fitting our observed data with the given model, a number of parameters can be obtained from Eq. (13). Figure 2(a) shows the dependence of the integrated spinodal correlation volume Ωsp on 1/λ04, which indicates that Ωsp is a convex-downward function of 1/λ04, with the parameter β=0.25nm1 [2629,30,31,3335] and the correlation length ζ=19.95nm. The value of β—the widely accepted one in the literature [2629,30,31,3335]—corresponds to the density fluctuation wavelength λm=25nm (where λm=2π/β), while the value of ζ is determined to fit best our measured turbidity for the dry sample, as will be discussed later (see Fig. 4). Its value is about twice as large as that in the literature [28]. Figure 2(b) shows the dependence of spinodal-decomposed turbidity τsp on 1/λ04, which indicates that τsp is also a convex-downward function of 1/λ04.

 figure: Fig. 2.

Fig. 2. (a) Dependence of integrated spinodal correlation volume Ωsp on 1/λ04, indicating Ωsp is a convex-downward function of 1/λ04, with the parameter β=0.251nm1 [2729] and the correlation length ζ=19.95nm, which was determined to fit best our data described in the text. (b) Dependence of corresponding turbidity τsp on 1/λ04, indicating τsp is also a convex-downward function of 1/λ04.

Download Full Size | PDF

The parameters for our dry sample fitted using this model provide both the estimated spinodal correlation function γsp(r) and the corresponding estimated correlation volume ωsp(q), as shown in Figs. 3(a) and 3(b), respectively. In Fig. 3(a), below 60 nm a clear oscillation can be observed with a period of around 25 nm. Figure 3(b) shows a strong correlation peak around 0.25nm1, as expected.

 figure: Fig. 3.

Fig. 3. (a) Dependence of spinodal correlation function γsp on the radius r in nanometers. (b) Dependence of the corresponding correlation volume ωsp(q) on the scattering vector q. Both curves are estimated from the respective models, namely, Eq. (11) for (a) and Eq. (12) for (b), with parameters β=0.251nm1 and ζ=19.95nm, both of which were determined to fit best the measured τsp in the range of 350 to 850 nm.

Download Full Size | PDF

In the given developments both for fractal and spinodal-decomposed trubidities, we assumed the dielectric correlation function η(0)η(r) can be expressed as in Eq. (4) or (11). However, strictly speaking, the expressions in Eqs. (4) and (11) are those for the density–density correlation, which originates from the angular dependence of the scattered intensities observed by SAXS or SANS. The dielectric constant is mainly a function of local density, and then the dielectric constant fluctuation can be related to the density fluctuation around its thermodynamic equilibrium value. Therefore, the correlation function for η is related to that for ρ in the following manner [16,17]:

η(0)η(r)=(εAVρ)S2ρ(0)ρ(r).

Hence, the constants Cf in Eq. (4) and Csp in Eq. (11), both of which include the coefficient (εAV/ρ)2, have been assumed to be constant in this derivation.

3. EXPERIMENTAL VERIFICATION

We have reported the time-dependent change in the transmittance spectrum of a porous glass chip in the UV-Vis-near-IR region (300–2500 nm) during the drying process (see Fig. 1 in [8]). Subsequently, we have also reported the effect of water desorption on the transparency change of the nanoporous Vycor glass from the fluctuation viewpoint [40]. In that report, the effect of water desorption could be clearly separated from the intrinsic optical properties of the Vycor glass itself, such as ultraviolet absorption, structural inhomogeneities due to spinodal decomposition, and so on, by dividing the observed transmission by the transmission observed in a very low humidity state. Here, let us consider how this treatment is justified.

The turbidities derived in the previous section, τf and τsp, are considered to be additive. Thus, for nonabsorbing glasses, the observed turbidity τext can be expressed as τext=τsca=τf+τsp. Since the transient white turbidity is the scattering caused by drainage of a wetting fluid from the pore space of porous Vycor glass in the process of forming a fractal-like percolation network of drained pores, no fractal turbidity τf is expected to be observed in samples for which drainage has been completed. For such a dry sample, only the turbidity inherent to structural inhomogeneity, which was formed during the process of spinodal decomposition, τsp, might be observed. Thus, the sample during the drying process will exhibit both the turbidity due to the fractal-like percolation network of drained pores and that due to the structural inhomogeneity caused by the spinodal decomposition, whereas the dried sample will exhibit only that due to the spinodal decomposition.

In terms of the incident light intensity (Iin) and the transmitted intensity through the scattering medium (Iout), the observed transmittance is expressed as

T=IoutIin=(1r)2exp(τext·dopt),
where τext is the total turbidity and r is the reflection of a single interface between air and porous glass, which is given for normal incidence by
r=(nairnpGnair+npG)2,
where nair and npG are the respective refractive indices of air and porous glass.

Using the additive property of the turbidities, we can factorize the transmittance as follows:

T=exp(τf·dopt)·(1r)2exp(τsp·dopt).

For the time-dependent change in the transmittance of a porous glass during the drying process, if we assign the time-dependent and final transmittances, T(λ0,t) and T(λ0,te), to the term involving the factor exp(τf·dopt) and that involving the factor (1r)2exp(τsp·dopt), respectively, then the transmittance due to the fractal-like percolation network of drained pores can be obtained from the relative transmittance, defined as

Tr(λ0)T(λ0,t)/T(λ0,te).

In the absence of absorption and reflection at interfaces, the fractal turbidity τf can be simply expressed as

τf(λ0)=1doptln{1Tr(λ0)}=ηf2AVπ2λ04ω0Ω5/2,
where Tr(λ0) is the relative transmittance and dopt is the sample thickness (or optical path length) in μm, and τf is in μm1. This expression is exactly the same as that used in the previous report [40]. Therefore, it is possible to extract the effect on a fractal-like percolation network of drained pores from the observed turbidity by dividing the intermediate transmittance data by the transmittance of dry Vycor.

On the other hand, the spinodal-decomposed turbidity τsp is expressed in a slightly complicated form as

τsp(λ0)=1doptln{1T(λ0,te)}+1doptln(1r),
where the reflectance r should be calculated with the refractive index of dry porous Vyrcor glass. In this treatment, the effect of water desorption is related to the turbidity due to the fractal-like percolation network of drained pores, τf. Therefore, the remaining turbidity at the dry state is considered to be mainly due to the inhomogeneities caused by spinodal decomposition.

To determine parameters ξ or ζ by fitting the experimentally observed turbidity to the respective theoretical curve, we need to estimate the respective proportional coefficient ηf2AV or ηsp2AV.

A completely dry or completely water-saturated Vycor glass can be considered to consist of two homogeneous phases. One of them is a pore region of varying and undetermined shape distributed within a homogeneous matrix. The other is a homogeneous matrix in which pores are formed. Only two dielectric constants need to be considered: εp in the pores and εs in the matrix. If the pores are filled with air or water, then εp is equal to that of air or water (εair or εwater). In both the completely dry or completely water-saturated cases, it should be considered that there remain only the inhomogeneities of type I, that is, those reminiscent of spinodal decomposition. Therefore, the mean square variation of the dielectric constant ηsp2AV is expressed as [49]

ηsp2AV=(εpεs)2φ(1φ),
where φ is the porosity of the porous glass (for Vycor, porosity φ is about 0.3 [11]).

On the other hand, for the partially water-saturated case, we have to consider the change in the dielectric constant of porous glass, εpG, as a function of the amount of water in the pores (precisely speaking, in terms of the pore-filling fraction, as explained subsequently) because of the imbibition/drainage of water into/from pores. In this case, the dielectric constant of porous glass can be estimated by using effective-medium models in which a pore is regarded as a superposition of its components. The most frequently used model is that of Bruggeman [50]. It is assumed in this model that the following formula is valid for pore clusters, which are cavities that connect with each other:

0=f·εwaterεp(f)εwater+2εp(f)+(1f)·εairεp(f)εair+2εp(f),
where f is the pore-filling function, which is experimentally estimated from the absorbance peak at around 1900 nm converted from the transmittance observed in the near-infrared (NIR) region, as previously described in detail [810]. The Maxwell–Garnett approximation is commonly used to determine the effective dielectric constant of dispersive media in which the size of inhomogeneities and the distances between them are much smaller than the wavelength of incident light. In this approximation, a dispersive phase is characterized by its relative volume fraction, and inhomogeneities are assumed to be negligibly small in size. In this case, the Lorentz–Lorenz formula [51] is used to calculate the effective dielectric constant of porous Vycor glass, εpG(f), which gives
εpG(f)=εSiO2[εp(f)+2εSiO2+2φ{εp(f)εSiO2}εp(f)+2εSiO2φ{εp(f)εSiO2}],
where εSiO2 is the dielectric constant of the silica skeleton. From Eq. (24), the refractive index of porous glass is obtained as npG(f)=εpG(f).

During the drying process, as the pore-filling fraction is reduced, the turbidity due to the fractal-like percolation network of drained pores shows characteristic transient behavior: it is initially small, rapidly increases to its maximum value, and then decreases and approaches the saturated value. This time dependence exactly corresponds to the time sequence of the phenomenological appearance of the sample, that is, initially transparent, then transitory whitish when it scatters light, and finally transparent again. This behavior parallels that of a binary alloy, where the intensity of the diffuse scattering varies as c(1c), c being the proportion of one of the components, such as the empty pore concentration. Instead of c, we will use the pore-filling fraction f to specify the transient mean square variation of the dielectric constant as follows:

ηf2AV=(εairεwater)2f(1f)·φ(1φ).

Here, f(1f)-type dependence of ηf2AV on the filling fraction f has already been proposed by Li et al. [2729].

On the basis of this consideration, we replot the turbidities derived from the measured transmittance profiles between 350 and 850 nm as a function of 1/λ04 in Fig. 4(a), when the sample exhibits white turbidity (2) and when the completely saturated or dry sample exhibits its transparency (1,3). The figure shows the convex-upward curve for 2 or convex-downward curves for 1 and 3, respectively, as a function of the inverse fourth power of the wavelength in air. Convex-upward curve 3 is predicted by the fractal turbidity model, namely, ω0Ω5/2, whereas convex-downward curves 1 and 3 are predicted by the spinodal-decomposed turbidity model, namely, Ωsp, both of which are derived by the dielectric fluctuation theory. Correspondingly, Fig. 4(b) shows the respective transmittance between 350 and 850 as a function of wavelength.

 figure: Fig. 4.

Fig. 4. (a) Changes in the turbidity (estimated from the logarithm of the observed transmittance) as a function of the inverse fourth power of wavelength in air (1/λ04). The slight deviations of the turbidities for 1 and 3 from the λ4 Rayleigh wavelength dependence are well fitted by the spinodal-decomposed turbidity curves, while that for 2 is well fitted by the one-parameter theoretical curve based on the fractal scattering with fractal dimension of 2.5. (b) Corresponding change in UV-Vis-light transmittance spectra of a porous Vycor glass after drying for 0, 45, and 165 min immediately after removal from ultrapure water immersion for 2 h at room temperature. In both (a) and (b), the previous results (Fig. 1 in [8]) are re-examined on the basis of the theory of dielectric constant fluctuations, and a comparison between the measured and fitted data is shown by solid and dotted lines, respectively.

Download Full Size | PDF

In particular, with a specified value of the pore-filling fraction f, the measured fractal turbidity curve 2 is fitted by the theoretical curve, which contains only one fitting parameter, ξ. In Figs. 4(a) and 4(b), a comparison between the measured and fitted data is shown by solid and dotted lines, respectively. The agreement between the theory and experiment in both figures is excellent. From this fitting, the correlation length ξ can be obtained as a function of the pore-filling fraction f, whose functional form has been already reported in [40].

4. DISCUSSION

So far we have shown that how the microscopic structures described by the correlation functions—which were experimentally observed by light scattering, SAXS, and SANS measurements—affect the optical properties of nanoporous Vycor glass in the visible range via scattering. In our analysis, the key assumption is based on the form of dielectric functions, which are taken to be equivalent to the correlation functions. We will start to discuss this point first.

Our approach might be justified by the following reasons: (i) the optical properties of isotropic transparent nanoporous materials are reported to be mainly determined by their microscopic structure consisting of innumerable nanopores and almost pure silica skeleton [18,19]; (ii) their microscopic structure has been extensively studied by the scattering techniques such as SAXS and SANS measurements [2036] and has been shown to have a pronounced inhomogeneous structure which is well characterized by the relevant correlation function; (iii) optical inhomogeneities are known to scatter light strongly [16,17]; nevertheless (iv) porous Vycor glasses are transparent in the visible range and predominantly exhibit insignificant light attenuation, determined by the relatively low level of scattering by inhomogeneities, even they are structurally inhomogeneous due to the presence of the innumerable nanopores. This implies that inhomogeneities are almost uniform within the glass and much smaller than the wavelength of the visible light scale. In this sense, a scheme is desirable to account for two kinds of turbidities, that arising due to fractal percolation of pores with fluid inside and that arising due to the structural inhomogeneity in terms of the correlation function, which contains a parameter called the correlation length; when the correlation length becomes comparable to the incident wavelength, strong scattering will be expected.

In applications of scattering techniques, interaction of radiation with matter is used for studying structural properties of the system of interest. In that case, the probing radiation has been chosen to be weakly coupled to the system and to match the condition for a required special resolution, which is inversely proportional to the momentum transfer, q. Measurements of the time-averaged scattered intensity Is(q) can be related to structural inhomogeneity, from which the spatial correlation functions such as the pair correlation function g(r), that is, Eq. (10), are obtained. In this sense, our approach is just the reverse of the scattering measurements; that is, from the assumed correlation functions, we have shown what optical responses are expected in the visible region on the basis of the fluctuation theory of dielectric constants.

All transparent matter scatters light due to thermal fluctuations which, in turn, generate fluctuations in dielectric constants or refractive index [16,17]. Glass differs in that these fluctuations are frozen in when the material is cooled through the annealing range of temperature [16]. As mentioned in the last part of Section 2, Eq. (15) directly relates the dielectric function to the density-density correlation function g(r)1, observed by small-angle scattering (SAS) measurements. In the liquid state, single component systems are physically inhomogeneous, the average density ρ0 being broken up by thermal fluctuations of mean square density (Δρ)2AV given by [52]

V(Δρ)2AV/ρ02=kBTAβT,
where βT is the isothermal compressibility evaluated at the fictive temperature TA related to thermal treatments, kB is Boltzmann’s constant, V is the volume of the fluctuating element, and the average dielectric constant εAV is related to the refractive index n as n=εAV. Given the presence of density fluctuations, (Δρ)2AV, in a single component material, light or X-rays will be scattered in proportion to (Δρ)2AV. Statistical mechanics (see [52]) gives the proportional constant in Eq. (15) as [53]
(εAVρ)S2(Δρ)2AV=εAV4p·kBTAβT,
where p is the photoelastic coefficient. For annealed samples of borosilicate silica, the fictive temperature is estimated as 1127 K (900°C), the high-temperature isothermal compressibility is about 6.8×1012cm2/dyn, and the photoelastic coefficient is about 3.65, nD=1.33 (for Vycor) [11,54,55].

As shown in Section 2, the structural information such as the fractal-like percolation network or spinodal-decomposed structure is incorporated into the functional form of the correlation function, Eq. (4) or Eq. (11), respectively. As discussed in [28], the functional form of the density–density correlation function, Eq. (10), is derived on the basis of Cahn’s linear theory for spinodal decomposition [56]. On the other hand, the compositional information is incorporated into the mean square variation of the dielectric constant, ηsp2AV or ηf2AV, that is, Eq. (22) or Eq. (25) [49]. The form given by Eq. (23) to Eq. (24) is based on the effective medium model [50,51,57,58].

The functional form of the dielectric function due to spinodal decomposition, Eq. (12), contains two adjustable parameters, β and ζ, while its proportional constant ηsp2AV, Eq. (22), is uniquely determined by material parameters such as the porosity and dielectric constants of constituent materials.

To see the effect of these adjustable parameters in the correlation function on the optical properties of Vycor in the visible range, a series of estimations with various combinations of β and ζ is demonstrated in Fig. 5. The effects of parameter changes of β and ζ on the correlation volume ωsp(q) and on the transmittance are demonstrated in Figs. 5(a) and 5(b), respectively. As depicted by curves denoted as E, A, and D, the value of β becomes smaller from 0.3 to 0.2nm1 while the correlation length ζ is fixed to the value of 19.95 nm. As shown in Fig. 5(a), the value of β determines the position of the so-called Vycor peak while the value of ζ determines the width of half-height of the correlation volume. The corresponding transmittance curves in Fig. 5(b) demonstrate that the transmittance in the short wavelength range is remarkably improved as the value of β becomes short. Because of the relation β=2π/λm, the smaller value of β implies the longer quasi-periodicity λm of the structural inhomogeneity existing even in the inhomogeneous porous structure, and subsequently the scattering due to inhomogeneities in the short wavelength range becomes less. As depicted by curves denoted by B, A, and C, the correlation length ζ becomes long from 10.2 to 29.0 nm, while the value of β is fixed to 0.251nm1, the most frequently reported value. Correspondingly, the transmittances in the short wavelength range are improved as the correlation length ζ becomes longer. This tendency of transparency improvement is the same as that of the effect of smaller β, but in this case it is less effective. In this case, the correlation length ζ determines the extent of the local quasi-periodic range caused by the spinodal decomposition. Thus, the longer the correlation length ζ, the more transparent the porous glass in the short wavelength range. However, this effect is considered to be local as compared to that caused by the small value of β. That is to say, the value of β seems to determine the quasi-periodicity of the entire structural inhomogeneity.

 figure: Fig. 5.

Fig. 5. (a) Effects of changes in parameters (β,ζ) on the correlation volume ωsp(q) as a function of q. Each ωsp(q) curve corresponds to the parameter pair denoted by A: (β [nm1], ζ [nm]) = (0.251, 19.95), B: (0.251, 10.2), C: (0.251, 29.0), D: (0.200, 19.95), and E: (0.300, 19.95), respectively. (b) Corresponding changes in the estimated UV-Vis-light transmittance spectra of a porous Vycor glass.

Download Full Size | PDF

Next, let us discuss the SAS data on Vycor 7930 in literatures. There are numerous contradictions and discussions in the literature regarding the fractal structures in Vycor glasses [2138] (see Table 1). The fractal dimensions of solids are mainly determined by SAXS and SANS. For example, an analysis of the SAS pattern from dry Vycor [23,25,32] suggests that the glass possesses a rough surface with a fractal dimension D2.5 and that the carriers of the fractal surface properties are interconnected units (cluster, drained pores) with an average diameter of 35 nm for Vycor glass [22]. However, several other investigations have shown no evidence of surface fractality in H2O-saturated samples [26,30].

Tables Icon

Table 1. Structural Information on Vycor 7930 Obtained by SAS Measurementsa

Mildner and Hall [20] provided a unified scheme to interpret SAS data in terms of the concept of fractal structure. Its essence is that from the power law dependence of the scattered intensity Is(Q)Qn at the high-Q region within the SAS regime, the fractal dimensionality Ds is defined as n=6Ds in three-dimensional objects. However, in the case of Vycor, the so-called Vycor peak around 0.25nm1 had prevented the accurate determination of the gradient of Is(Q). If the slope of 4 at large Q is observed, it is the signature of scattering from a smooth surface, that is, nonfractal.

In subsequent investigations [2933], it has been shown that a fractal surface can be defractalized upon deposition of an adsorbed film of water [24]. On the other hand, a fractal geometry was indicated in Vycor glass on a length scale larger than 100 nm with the percolation network formed by the empty pores and the value of the fractal dimension of 1.75 by SANS studies [2629]. This geometry was also indicated on a length scale larger than 10 μm with a pore-space correlation of the fractal dimension of 2.6 by laser light scattering (LLS) measurements [37,38]. Contrary to these results, it has been noted [23] that the SANS and SAXS data leave little room for the idea that Vycor glass has a fractal pore network on a scale above 4 nm.

We will discuss this issue in terms of the turbidities derived in the previous section. The slight deviation from the λ4-wavelength dependence of the observed turbidity, which is quantitatively expressed as in Eq. (7) or (13), is expected to shed light on this problem.

As shown in Fig. 4(a), the observed turbidities for the completely water-saturated Vycor (1) and dry Vycor (3) exhibit the convex-downward curves, while the turbidity observed for the partially saturated Vycor (2) exhibits the convex-upward curve. The convex-downward curve can be well reproduced by our derived spinodal-decomposed turbidity τsp, while the convex-upward curve can be reproduced by our derived fractal turbidity τf. Since we have adopted the f(1f)-type dependence of ηf2 on the pore-filling fraction f in Eq. (25) as the coefficient for Eq. (7), it is natural to observe that there is no fractal scattering for the completely water-saturated Vycor and dry Vycor, as previously reported in [21,22,26,30]. However, this functional choice for ηf2 conflicts with the observations in [23,36].

Our simple interpretation of these contradictions is that porous Vycor glasses are not fractal in themselves (although they are very porous), but the filling of their pores, if incomplete, can exhibit fractal properties. Our observations [810] on the transitory whitish turbidity phenomenon strongly support this interpretation that the transient phenomenon is due to a fractal-like percolation network of drained/imbibed pores in the pore space.

Our third issue concerns the effective optical constants for porous dielectric structures proposed by Vereshchagin et al. [57]. In their derivation of the effective refractive index for a porous nonabsorbing layer, they started from the following expression:

neff=n[1i2πNk3S(0)],
where N is the number of pores in a unit volume, n is the refractive index of the material of the matrix in which pores are formed, k=2πn/λ0 is the wave number, λ0 is the wavelength in vacuo, and S(0) is the amplitude scattering function for a separate pore. Using the dipole approximation to specify the amplitude scattering function and taking into account the back action of radiation on dipoles, they finally derived the following formula:
neff=n[1i2πNk3(ik3α+23k6α2)],
where α is the polarizability of a nonuniformity.

Their approach seems to be based on the particle-scattering viewpoint with absorption, originally from the famous text on light scattering by Van de Hulst [59]. However, within the given framework, this formulation will lead to a simple λ4-wavelength dependence, because the field amplitude of a wave traveling in the z-direction in the medium specified by Eq. (29) is proportional to

exp{ik0(ctneffz)}=exp(iωt)exp{ik(1+2πNα)}exp(4π3Nk4α2·z),
where c=ω/k0 is the speed of light, and then the intensity of the wave decreases as exp{(8π/3)Nk4α2·z}, which provides the k4-dependence of the extinction. Consequently, it does not explain any deviation from the λ4 Rayleigh wavelength dependence.

In contrast, our approach based on the fluctuation viewpoint has the following advantage: by introducing two types of turbidity—one caused by structural inhomogeneity such as that reminiscent of spinodal decompition and the other by compositional inhomogeneity such as inhomogeneous distribution of wetting fluid in the pore space—we can describe exactly the transmittance spectrum of porous Vycor glass in the visible region without introducing the complex refractive indices, namely, n˜=n+iκ, as had been done by Kuchinskii et al. [51] and other groups [18,19,50,57].

Next, we will briefly comment on the absorption peak at around 1900 nm in the NIR region, although this is not the main subject of this paper. As described in the previous section, we used the absorbance peak at around 1900 nm to estimate the pore-filling fraction, which is related to the amount of water within the pore space. This treatment is based on the result reported by Wood et al. [60], who performed infrared spectroscopic studies on dried alkoxide silica gels. They found that the broad absorption at 5292cm1 (1890nm) is attributed to hydrogen-bonded H2O molecules only and that its isolation from other absorption bands makes it possible to study this molecular species independently of other OH-containing groups. Therefore, a convenient quantitative determination of H2O content in the pore space can be made from the intensity of the 5292cm1 infrared absorption band by using the appropriate value of the extinction coefficient. Subsequent infrared studies on porous Vycor glasses [61,62] verified their result, and now the absorption band near 5260cm1 (1900nm) is used for the characterization of molecular water within the pore space.

Now, we will discuss the difference between our fitted and measured transmittances in the shorter range of less than 350 nm. As shown in Fig. 6(a), the measured transmittance for the dry Vycor drops off more steeply than the estimated transmittance, which was best fitted in the range between 350 and 850 nm. If the transmittance theoretically estimated by the spinodal-decomposed turbidity is considered to provide a theoretical limit of optical transmittance in the ultraviolet region, this observed steep drop of the transmittance implies that, for the following two reasons, there must exist some absorbing species within this energy range: First, as shown in Fig. 4(b), both observed transmittances for completely water-saturated and dry Vycor glass start to drop off steeply at about 297 nm, where the two curves intersect and below which both curves drop off along the same line. This strongly indicates the existence of an absorbing species around that wavelength. Second, since the optical bandgap and the Urbach tail for vitreous silica are reported to be about 9 eV [13] and 7.4–8.8 eV [63,64], which are much larger than our predicted limit of 155 nm (= 8 eV), the porous structure of Vycor with innumerable pores, which is characterized by the spinodal correlation function, Eq. (10), does not cause the observed steep drop of transmittance.

 figure: Fig. 6.

Fig. 6. (a) Measured and fitted UV-Vis light transmittance spectra of a porous Vycor in the dry state at 165 min, on the basis of the theory of dielectric constant fluctuations. The fitted curve is drawn with the parameters derived from the best fitting of τsp in the range of 350 to 850 nm. (b) Difference between measured and fitted UV-Vis light absorbances, estimated from Abs(λ0)=log10{Tfit(λ0)/Tmeas(λ0)}, as a function of the photon energy. The observed optical absorption spectrum is best fitted by three Gaussian curves, denoted as A, B, and C.

Download Full Size | PDF

Two possible candidates for the absorption in this region are as follows: one is the absorption caused by some organic or inorganic contaminants, which are unintentionally adsorbed on the pore walls; the other might be the optical absorption due to structural defects inherent to the amorphous network of vitreous silica itself, such as peroxy linkage (SiOOSi), E centers (Si·), nonbridging oxygen hole centers (NBOHCs, SiO·), and oxygen-deficiency centers (SiSi) [65].

In order to identify these absorption peaks, as shown in Fig. 6(b), we have tried to separate them by fitting the absorbance versus photon energy plot with three Gaussians of the form

Abs(E)=μ=A,B,CIμexp{(EEμ)22σμ2},
where Eμ, σμ, and Iμ are their peak positions, half-widths (full width at half-maximum), and intensities, respectively, assuming Gaussian shapes. The separated results are tabulated in Table 2.

Tables Icon

Table 2. Optical Absorption Bands Separated by Three-Gaussian Fitting

E centers around 5.8 eV [6567] and Si-Si bonds (oxygen vacancy) around 7.6 eV [65,66,68,69] are the major reported optical absorption bands of defects in vitreous silica. SiSi bonds seem to be easy to exclude because of their higher peak positions. The observed largest peak, denoted as “C,” might be from E centers because our fitting with the peak energy of 5.29 eV has some arbitrariness of peak positioning. On the other hand, three oxygen-excess defects—NBOHC [65,67], peroxy radicals (PORs) [67], and interstitial O3 (ozone)—are not so easy to exclude from our consideration because those defects are observed around 4.8 eV. For example, the POR absorption band is found at 4.8 eV in bulk samples [67]. Furthermore, even the SiSi oxygen vacancy defect has also been attributed to the optical absorption band near 5.0 eV [68]. Unfortunately, the most controversial region in the literature is also that around 4.8 eV. This fact makes the assignment of these peaks to oxygen-excess defects very difficult. Thus, the attribution of the observed peak to one of these defects is beyond the scope of our consideration in the present circumstances.

Another possible reason for the strong absorption peak at around 5.29 eV might be related to the extrinsic charge transfer and sp absorption bands due to trace impurities of metal ions such as Fe3+/Fe2+ [70]. In the course of an investigation focused on the transparency improvement in the vacuum-ultraviolet region of high-purity silica glass, reducing the trace impurities of transition metals turned out to be the most effective way to improve the ultraviolet transmittance [71]. In glasses melted under normal atmospheric conditions, El-Batal et al. [70] reported that the iron impurity exists mainly as Fe3+ species with a UV band at about 230–250 nm. This absorption band seems to be coincidental with our observed absorbance peaks. However, we cannot draw any definitive conclusion about this issue because of a lack of related evidence.

Finally, we will discuss the applicability of our scheme to other transparent nanoporous materials. In this paper, we are mainly concerned with isotropic transparent nanoporous materials such as Vycor glasses, which are characterized by a three-dimensional random pore network. However, there are other types of nanoporous materials, which particularly are characterized by the network of parallel cylindrical nanochannels, that is, considerably different from the three-dimensional pore network of Vycor. It is reported that such materials also exhibit turbidities and that the light scattering in such materials drastically rises in the capillary condensate regime, quite similar to Vycor glass [72,73]. At this moment, it seems quite difficult to account for such an anisotropy in our scheme because the integrated correlation volume is obtained by assuming the fundamental spatial isotropy so that the expressions such as Eqs. (8) and (14) are possible. If the correlation function is not isotropic but possesses rotational symmetry around a unique axis, as claimed by these experimental observations, one would obtain a scattering law that has an elliptically symmetric dependence on the azimuthal orientation of the scattering vector q. This elliptical dependence necessarily might be removed by averaging the scattering law in terms of a reduced scattering vector to obtain the integrated correlation volume Ω in Eqs. (8) and (14). In this sense, our analysis will encounter difficulty at this point. Further work is required to address this point.

5. CONCLUSION

We have studied the applicability of the turbidity analysis of the photospectroscopically measured data as a method to study the correlation functions that characterize the pore space and the structural features of isotropic transparent porous media, on the presupposition that there exists no light attenuation other than the scattering. This kind of structural analysis has been conventionally carried out by measuring the small-angle scattered intensity with a monochromatic probing radiation as a function of scattering vector.

In this paper, we have shown that the light propagation and scattering in monolithic nanoporous Vycor glass exhibit two optical turbidities in the visible (Vis) region: one is a transient white turbidity, characterized by the convex-upward dependence to the inverse fourth power of wavelength (1/λ04), and the other is the turbidity inherent to the structural inhomogeneity, characterized by the convex-downward dependence. The former is assumed to be caused by a fractal-like percolation network of imbibed/drained pores as a consequence of imbibition/drainage of wetting fluid into/from pore space. The latter is considered to be caused by the structural inhomogeneities inherent to the Vycor glass itself, which are produced by spinodal decomposition.

To explain and verify the slight deviation from the λ4 wavelength dependence of photospectroscopically observed turbidities, we have derived respective analytical expressions for them on the basis of the theory of dielectric constant fluctuation. For the fractal-like percolation network, we obtained the general expressions for the turbidity due to a fractal-like percolation network, τf, with half and integer values of fractal dimensions from 1/2 to 3, and verified that all of them are the convex-upward functions of 1/λ04. On the other hand, the derived expression for the spinodal-decomposed turbidity τsp turned out to be a convex-downward function of 1/λ04.

In summary, we have developed a general scheme to estimate the transmittance spectra of isotropically transparent nanoporous glass such as Vycor through the turbidity in the visible region. The fluctuation approach can more systematically and comprehensively treat the problem of concern with a slight deviation from the λ4 Rayleigh wavelength dependence of the observed turbidity.

APPENDIX A: DERIVATION OF THE BASIC EQUATIONS

Scattering theory based on the Maxwell equations for a nonconducting, nonmagnetic medium can be used to derive Eqs. (1) to (3), the basic equations for the turbidity analysis. We follow the treatment of Landau et al. [41] and Chu [42].

In the fluctuation approach, the inhomogeneity present in a condensed medium such as porous glass is generally characterized by introducing the concept of a spatial correlation function, in which the inhomogeneity of the medium is considered to be due to a continuous variation of the dielectric constant or refractive index rather than to the presence of discrete particles. Let ε(r) denote a local dielectric constant at position r and εAV be its average (which is related to the refractive index n=εAV). The function η(r)=ε(r)εAV represents the local fluctuation of the dielectric constant, and

γ(r1,r2)η(r1)η(r2)/η2AV
defines a dimensionless spatial correlation function, where η2AV is the average value of η2. Here represents an ensemble average. For a plane-wave incident electric field,
Ein(R,t)=n^iE0exp[i(ki·Rωit)],
where Ein is assumed to be a simple sinusoidal traveling wave of frequency ωi in the direction ki with phase velocity cm(=ωi/|ki|). Here, ki is the wave vector in the propagating direction with magnitude of |ki|=2πn/λ0 (λ0 is the wavelength in vacuo), n^i is a unit vector in the direction of polarization of the incident field, and E0 is the field amplitude. According to the classical theory of electromagnetic waves, the component of the scattered electric field at a large distance R (=|R|) from the scattering volume V with the dielectric fluctuation η(r) can be expressed in vector form as [41,42]
Es(R,t)=ks×(ks×n^i)E0εAVexp[i(ksRωit)]4πR×Vd3rη(r)exp[i(kiks)·r],
where the wave vector ks of the scattered field is in the same direction as R and the subscript V indicates that the integral is over the scattering volume V.

Assuming that the scattering is weak and the Born approximation is valid, the time-averaged scattered intensity is expressed as [41,42]

Is(kiks)=12εAVμ|E0|2|ks×(ks×n^i)|2(4πR)2εAV2×vd3r1d3r2exp[i(kiks)·(r1r2)]η(r1)η(r2),
where μ is the permeability, which is unity for a nonmagnetic dielectric medium. Now, we introduce the vector defined by q=kiks as the scattering vector and the angle between ki and ks as the scattering angle θ. For an elastic scattering such as Rayleigh scattering, the scattered wave has essentially the same wavelength as that of the incident wave: |ks||ki|k=2πn/λ0. Thus it follows that q=2ksin(θ/2).

For an isotropic medium, the spatial correlation function γ(r1,r2) depends only on the relative distance r=|r1r2| and not on the direction in space. Then γ(r) becomes a spherically symmetric function that has a value of unity at r=0 and decays to zero as r. The steepness with which this function drops off from unity to zero is a measure of the average extension of the inhomogeneities. If the dielectric medium is isotropic, the double integral in Eq. (A4) can be replaced by the product of the scattering volume V and the integral over d3r whose angular integration can be performed, yielding

Is(q)=Iinks4sin2χ(4πR)2·η2AVεAV2V04πr2γ(r)sin(qr)qrdr,
where we use the facts that the incident intensity of the form Iin=12εAVμ|E0|2 is the flow of power per unit area of the incoming light beam and that the magnitude of the vector cross product ks×(ks×ni) is equal to ks2sinχ, where χ is the angle measured from the scattering direction to the dipole. The sin2χ angular dependence in Eq. (A5) is characteristic of the dipole radiation, so its dependence can also be expressed in terms of the scattering angle θ as (1+cos2θ)/2 with respect to the scattering plane.

The differential cross section is defined as

dPdΩ=12Re[R2R^·Es×Hs*]=Is(q)R2IinV,
where V is the scattering volume and R^ is the unit vector in the same direction as R, that is, R^R/|R|, and * indicates the complex conjugate. The differential cross section has the dimensions of reciprocal length. The total scattering cross section, known as the turbidity τ, is related to the differential cross section by integrating over the whole solid angle, namely, τ=4πdΩ(θ,φ)·dPdΩ. Substitution of Eqs. (A5) and (A6) into this expression gives
τ=η2AVks4(4π)2εAV24πdΩ1+cos2θ2ω(q)=η2AVπ2λ044πdΩ1+cos2θ2ω(q),
with the correlation volume, which is a function both γ(r) and the scattering angle θ, defined as
ω(q)=04πr2γ(r)sin(qr)qrdr.

A simple physical interpretation of Eq. (A7) is that the turbidity is proportional to the product of the mean square fluctuation of the dielectric constant, η2AV, and an associated correlation volume ω(q).

A measure of the extent of the inhomogeneities is mathematically expressed by correlation length ξ, with the help of the correlation function involving the length scale in the form r/ξ as the exponential argument. Conventionally, the angular dependence of the scattered intensity [which is proportional to correlation volume ω(q)] from the sample with a monochromatic incident light beam [or the dependence of the scattered intensity on scattering vector q, i.e., I(q)] is measured to obtain the correlation function.

APPENDIX B: DERIVATION OF ΩD

Detailed mathematical expressions for correlation volumes and related quantities with half-integer and integral values of the fractal dimension D for the fractal correlation function of Eq. (4) with d=3 are summarized here. The integrated dissymmetry factor ΩD is required to obtain the turbidity τf due to a fractal-like percolation network of drained pores, as explained in Section 2.A.

For the integer and half-integer dimensions from 1/2 to 3, the explicit expressions for the anisotropic correlation volume ωf(q) are tabulated in Table 3 in terms of the dissymmetric factor of Dth order fD(q), defined as

fD(q)=ωf(q)ω0(D)=Γ(D1)Γ(D)[1+q2ξ2(qξ)2]1/2sin[(D1)tan1(qξ)][1+q2ξ2]D/2,
with the symmetrical correlation volume ω0(D), defined as
ω0(D)=4πξD·Γ(D).

Tables Icon

Table 3. Correlation Volumes and Dissymmetrical Factors of Fractal Correlation Functions

Thus, the correlation volume ωf(q) is provided by the product of Eqs. (B1) and (B2).

As shown in Table 3, Eqs. (4) and (5) for d=3 provide two particular cases: D=3, γf(r)exp(r/ξ), which gives the Debye law [43], ωf(q)ξ3/(1+q2ξ2)2, and D=2, γf(r)r1exp(r/ξ), which gives the Ornstein–Zernike law, ωf(q)ξ2/(1+q2ξ2).

The corresponding turbidity [τf(D)] is given by integrating the correlation volume over the entire solid angle dΩ, as follows:

τf(D)=ηf2AVk4(4π)2εAV2ω0(D)·ΩD,
where ΩD is the integrated dissymmetrical factor, defined as
ΩD=2π0πdθsinθ1+cos2θ2fD(2ksin(θ2)).

The product of Eq. (B10) with the symmetrical correlation volume ω0(D) provides the integrated correlation volume for the fractal scattering, that is, the integrated expression in Eq. (3): 4πdΩ1+cos2θ2ω(q)=ω0(D)·ΩD.

Equation (B9) implies that the wavelength dependence of the turbidity τf due to the fractal-like percolation network of fractal dimensionality D is determined mainly by 1/λ04 [which comes from the factor k4/(16π2εAV2)] and the integrated correlation volume ω0(D)·ΩD, which is a function of ξ only and determines the slight deviation from the λ4 Rayleigh wavelength dependence.

For half-integer and integer values of D, the explicit expressions for ΩD are as follows:

For D=1/2, in the difference form with b4k2ξ2,

Ω1/2=162πb3{f(1+1+b,b)f(2,b)},
where
f(z,b)=2z{z5115z49+8z374z25b(z373z25+2z3)+b22(z31)}.

In the explicit form,

Ω1/2=162π3465b3[2{256+33b(16+35b)}1+1+b{128(1+1+b)+8b(31+231+b)+b2(18227951+b)}].

For D=3/2, in the difference form with b4k2ξ2,

Ω3/2=162πb3{g(1+1+b,b)g(2,b)},
where
g(z,b)=2z{z494z37+4z25b(z252z3)+b22}.

In the explicit form,

Ω3/2=162π315b3[2{256+21b(16+15b)}+1+1+b{128(1+1+b)+8b(19+111+b)+259b2}].

For D=5/2, in the difference form with b4k2ξ2,

Ω5/2=162π3b3{h(1+1+b,b)h(2,b)},
where
h(z,b)=2z{z37z25z3(2+b)(z3+1)1+b+b2/2z1}.

In the explicit form,

Ω5/2=162π315b3[2(768+560b+105b2)1+1+b1+b{384(1+1+b)+8b(53+291+b)+145b2}].

For D=1, with b=4k2ξ2,

Ω1=4π15[22btan1(b)+6+7bb2(6+10b+15b2)b3ln(1+b)]

For D=2, with b=4k2ξ2,

Ω2=4πb2{(2+b)+2+2b+b2bln(1+b)}.

For D=3, with b=4k2ξ2,

Ω3=4πb2{(2+b)21+b2(2+b)bln(1+b)},
which is the same as that of Eq. (6′′) of Debye [43].

Acknowledgment

The authors are sincerely grateful to Drs. Osamu Kagami and Akinori Furuya of NTT for their encouragement and support throughout this work.

REFERENCES

1. T. Tanaka, T. Ohyama, Y. Y. Maruo, and T. Hayashi, “Coloration reactions between NO2 and organic compounds in porous glass for cumulative gas sensor,” Sens. Actuators B 47, 65–69 (1998). [CrossRef]  

2. T. Tanaka, A. Guilleux, T. Ohyama, Y. Y. Maruo, and T. Hayashi, “A ppb-level NO2 gas sensor using coloration reactions in porous glass,” Sens. Actuators B 56, 247–253 (1999). [CrossRef]  

3. Y. Y. Maruo, “Measurement of ambient ozone using newly developed porous glass sensor,” Sens. Actuators B 126, 485–491 (2007). [CrossRef]  

4. Y. Y. Maruo, J. Nakamura, M. Uchiyama, M. Higuchi, and K. Izumi, “Development of formaldehyde sensing element using porous glass impregnated with Schiff’s reagent,” Sens. Actuators B 129, 544–550 (2008). [CrossRef]  

5. A. F. Novikov and V. I. Zemskii, “Glassy spectral gas sensors based on the immobilized indicators,” Proc. SPIE 2550, 119–129 (1995). [CrossRef]  

6. T. Ohyama, Y. Y. Maruo, T. Tanaka, and T. Hayashi, “A ppb-level NO2 detection system using coloration reactions in porous glass and its humidity dependence,” Sens. Actuators B 64, 142–146 (2000). [CrossRef]  

7. K. Izumi, M. Utiyama, and Y. Y. Maruo, “Evaluation of air quality with simple and easy chemical sensors: development of porous glass-based elements,” in International Conference on Control, Automation and Systems (IEEE, 2008), pp. 2756–2760.

8. S. Ogawa, “1/λ4 scattering of light during the drying process in porous Vycor glass with nano-sized pores,” J. Opt. Soc. Am. A 30, 154–159 (2013). [CrossRef]  

9. S. Ogawa and J. Nakamura, “Hysteretic characteristics of 1/λ4 scattering of light during adsorption and desorption of water in porous Vycor glass with nanopores,” J. Opt. Soc. Am. A 30, 2079–2089 (2013). [CrossRef]  

10. S. Ogawa and J. Nakamura, “Optically observed imbibition and drainage of wetting fluid in nanoporous Vycor glass,” J. Opt. Soc. Am. A 32, 2397–2406 (2015). [CrossRef]  

11. P. R. Wakeling, “What is Vycor glass?” Appl. Opt. 18, 3208–3210, 3248 (1979). [CrossRef]  

12. H. R. Philipp, “Silicon dioxide (SiO2) (glass),” in Handbook of Optical Constants of Solids, E. D. Palik, ed. (Academic, 1985), pp. 749–766.

13. H. R. Philipp, “Optical properties of non-crystalline Si, SiO, SiOx and SiO2,” J. Phys. Chem. Solids 32, 1935–1945 (1971). [CrossRef]  

14. S. T. Pantelides and W. A. Harrison, “Electronic structure, spectra, and properties of 4:2-coordinated materials, I. crystalline and amorphous SiO2 and GeO2,” Phys. Rev. B 13, 2667–2691 (1976). [CrossRef]  

15. G. H. Sigel Jr., “Optical absorption of glasses,” in Treatise on Materials Science and Technology, M. Tomozawa and R. H. Doremus, eds. (Academic, 1977), Vol. 12, pp. 5–89.

16. J. Schroeder, “Light scattering in glass,” in Treatise on Materials Science and Technology, M. Tomozawa and R. H. Doremus, eds. (Academic, 1977), Vol. 12, pp. 158–222.

17. N. G. van Kampen, “Scattering of light by density fluctuations,” in Proceedings of the International School of Physics “Enrico Fermi,”, R. J. Glauber, ed. (Academic, 1969), Vol. 42, pp. 235–246.

18. A. A. Evstrapov, T. V. Antropova, I. A. Drozdova, and S. G. Yastrebov, “Optical properties and structure of porous glasses,” Opt. Appl. 33, 45–54 (2003).

19. A. A. Evstrapov, N. Esikova, and T. V. Antropova, “Spectral characteristics and structure of porous glasses,” Opt. Appl. 35, 753–759 (2005).

20. D. F. R. Mildner and P. L. Hall, “Small-angle scattering from porous solids with fractal geometry,” J. Phys. D 19, 1535–1545 (1986). [CrossRef]  

21. D. W. Schaefer, B. C. Bunker, and J. P. Wilcoxon, “Are leached porous glass fractal?” Phys. Rev. Lett. 58, 284 (1987). [CrossRef]  

22. P. Wiltzius, F. S. Bates, S. B. Dierker, and G. D. Wignall, “Structure of porous Vycor glass,” Phys. Rev. A 36, 2991–2994 (1987). [CrossRef]  

23. A. Hoehr, H.-B. Neumann, P. W. Schmidt, P. Pfeifer, and D. Avnir, “Fractal structure and cluster structure of controlled-pore glasses and Vycor porous glass as revealed by small-angle x-ray and neutron scattering,” Phys. Rev. B 38, 1462–1467 (1988). [CrossRef]  

24. E. Cheng, M. W. Cole, and P. Pfeifer, “Defractalization of films adsorbed on fractal surfaces,” Phys. Rev. B 39, 12962–12965 (1989). [CrossRef]  

25. P. Levitz, G. Ehret, S. K. Sinha, and J. M. Drake, “Porous Vycor glass: the microstructure as probed by electron microscopy, direct energy transfer, small-angle scattering, and molecular adsorption,” J. Chem. Phys. 95, 6151–6161 (1991). [CrossRef]  

26. M. J. Benham, J. C. Cook, J.-C. Li, D. K. Ross, P. L. Hall, and B. Sarkissian, “Small-angle neutron scattering study of adsorbed water in porous Vycor glass: supercooling phase transition and interfacial structure,” Phys. Rev. B 39, 633–636 (1989). [CrossRef]  

27. J.-C. Li, D. K. Ross, and M. J. Benham, “Small-angle neutron scattering studies of water and ice in porous Vycor glass,” J. Appl. Crystallogr. 24, 794–802 (1991). [CrossRef]  

28. J.-C. Li and D. K. Ross, “Dynamical scaling for spinodal decomposition—a small-angle scattering study of porous Vycor glass with fractal properties,” J. Phys. Condens. Matter 6, 351–362 (1994). [CrossRef]  

29. J.-C. Li, D. K. Ross, L. D. Howe, K. L. Stefanopoulos, J. P. A. Fairclough, R. Heenan, and K. Ibel, “Small-angle neutron-scattering studies of the fractal-like network formed during desorption and adsorption of water in porous materials,” Phys. Rev. B 49, 5911–5917 (1994). [CrossRef]  

30. A. Ch. Mitropoulos, J. M. Haynes, R. M. Richardson, and N. K. Kanellopoulos, “Characterization of porous glass by adsorption of dibromo-methane in conjunction with small-angle x-ray scattering,” Phys. Rev. B 52, 10035–10042 (1995). [CrossRef]  

31. A. Ch. Mitropoulos, P. K. Makri, N. K. Kanellopoulos, U. Keiderling, and A. Wiedenmann, “The surface geometry of Vycor,” J. Colloid Interface Sci. 193, 137–139 (1997). [CrossRef]  

32. M. Agamalian, J. M. Drake, S. K. Sinha, and J. D. Axe, “Neutron diffraction study of the pore surface layer of Vycor glass,” Phys. Rev. E 55, 3021–3027 (1997). [CrossRef]  

33. F. Katsaros, P. Makri, A. Mitropoulos, N. Kanellopoulos, U. Keiderling, and A. Wiedenmann, “On the morphology and surface geometry of Vycor,” Physica B 234–236, 402–404 (1997). [CrossRef]  

34. E. S. Kikkinides, M. E. Kainougiakis, K. L. Stefanopoulos, A. Ch. Mitropoulos, A. K. Stubos, and N. K. Kanellopoulos, “Combination of small angle scattering and three-dimensional stochastic reconstruction for the study of adsorption–desorption processes in Vycor porous glass,” J. Chem. Phys. 112, 9881–9887 (2000). [CrossRef]  

35. A. Ch. Mitropoulos, “Small-angle X-ray scattering studies of adsorption in Vycor glass,” J. Colloid Interface Sci. 336, 679–690 (2009). [CrossRef]  

36. M.-H. Kim and C. J. Glinka, “Ultra small angle neutron scattering study of the nanometer to micrometer structure of porous Vycor,” Microporous Mesoporous Mater. 91, 305–311 (2006). [CrossRef]  

37. J. H. Page, J. Liu, B. Abeles, E. Herbolzheimer, H. W. Deckman, and D. A. Weitz, “Pore–space correlations in capillary condensation in Vycor,” Phys. Rev. Lett. 71, 1216–1219 (1993). [CrossRef]  

38. J. H. Page, J. Liu, B. Abeles, E. Herbolzheimer, H. W. Deckman, and D. A. Weitz, “Adsorption and desorption of a wetting fluid in Vycor studied by acoustic and optical techniques,” Phys. Rev. E 52, 2763–2777 (1995). [CrossRef]  

39. M. Kerker, The Scattering of Light and Other Electromagnetic Radiation (Academic, 1969), pp. 31–39.

40. S. Ogawa and J. Nakamura, “Photospectroscopically observed pore-space correlations of a wetting fluid during the drying process in nanoporous Vycor glass,” J. Opt. Soc. Am. A32, 533–537 (2015). [CrossRef]  

41. L. D. Landau, E. M. Lifshitz, and L. P. Pitaevskii, Electrodynamics of Continuous Media, 2nd ed. (Pergamon, 1984).

42. B. Chu, Laser Light Scattering, 2nd ed. (Academic, 1991), pp. 21–28.

43. P. Debye and A. M. Bueche, “Scattering by an inhomogeneous solids,” J. Appl. Phys. 20, 518–525 (1949). [CrossRef]  

44. P. Debye, H. R. Anderson Jr., and H. Brumberger, “Scattering by an inhomogeneous solid, II, the correlation function and its application,” J. Appl. Phys. 28, 679–683 (1957). [CrossRef]  

45. S. K. Sinha, T. Freltoft, and J. Kjems, “Observation of power-law correlations in silica-particle aggregates by small-angle neutron scattering,” in International Conference on Kinetics of Aggregation and Gelation, Athens, 1984, pp. 87–90.

46. T. Freltoft, J. K. Kjems, and S. K. Sinha, “Power-law correlations and finite-size effects in silica particle aggregates studied by small-angle neutron scattering,” Phys. Rev. B 33, 269–275 (1986). [CrossRef]  

47. S. K. Sinha, “Scattering from fractal structures,” Physica D 38, 310–314 (1989). [CrossRef]  

48. I. S. Gradshteyn and I. M. Ryzhik, Tables of Integrals, Series and Products, A. Jeffrey, ed. (Academic, 1980).

49. M. Kerker, The Scattering of Light and Other Electromagnetic Radiation (Academic, 1969), p. 470.

50. A. A. Evstrapov and N. A. Esikova, “Study of porous glasses by the methods of optical spectroscopy,” J. Opt. Technol. 75, 266–270 (2008). [CrossRef]  

51. S. A. Kuchinskii, V. I. Sukhanov, M. V. Khazova, and A. V. Dotsenko, “Effective optical constants of porous glass,” Opt. Spectrosc. 70, 85–88 (1991).

52. L. D. Landau and E. M. Lifshitz, Statistical Physics (Addison-Wesley, 1969).

53. G. N. Greaves, Y. Vaills, S. Sen, and R. Winter, “Density fluctuations, phase separation and micro-segregation in silicate glasses,” J. Optoelectron. Adv. Mater. 2, 299–316 (2000).

54. M. E. Nordberg, “Properties of some Vycor-brand glasses,” J. Am. Ceram. Soc. 27, 299–305 (1944). [CrossRef]  

55. T. H. Elmer, “Porous and reconstructed glasses,” in Engineered Materials Handbook, Vol. 4 of Ceramics and Glasses (ASM International, 1992), pp. 427–431.

56. J. W. Cahn, “On spinodal decomposition,” Acta Metall. 9, 795–801 (1961). [CrossRef]  

57. V. G. Vereshchagin, R. A. Dynich, and A. N. Ponyavina, “Effective optical parameters of porous dielectric structures,” Opt. Spectrosc. 84, 427–431 (1998).

58. R. J. Gehr and R. W. Boyd, “Optical properties of nanostructured optical materials,” Chem. Mater. 8, 1807–1819 (1996). [CrossRef]  

59. H. C. van de Hulst, Light Scattering by Small Particles (Wiley, 1957), p. 66.

60. D. L. Wood and E. M. Rabinovich, “Infrared studies of alkoxide gels,” J. Non-Cryst. Solids 82, 171–176 (1986). [CrossRef]  

61. A. Burneau, J. Lepage, and G. Maurice, “Porous silica-water interactions, I. structural and dimensional changes induced by water adsorption,” J. Non-Cryst. Solids 217, 1–10 (1997). [CrossRef]  

62. A. Burneau and J.-P. Gallas, The Surface Properties of Silicas, A. P. Legrand, ed. (Wiley, 1998), pp. 147–234.

63. I. T. Godmanis, A. N. Trukhin, and K. Huebner, “Exciton-phonon interaction in crystalline and vitreous SiO2,” Phys. Status Solidi B 116, 279–287 (1983). [CrossRef]  

64. T. H. DiStefano and D. E. Eastman, “The band edge of amorphous SiO2 by photoinjection and photoconductivity measurements,” Solid State Commun. 9, 2259–2261 (1971). [CrossRef]  

65. D. L. Griscom, “Defect structure of glasses,” J. Non-Cryst. Solids 73, 51–77 (1985). [CrossRef]  

66. H. Imai, K. Arai, H. Imagawa, H. Hosono, and Y. Abe, “Two types of oxygen-deficient centers in synthetic silica glass,” Phys. Rev. B 38, 12772–12775 (1988). [CrossRef]  

67. L. Skuja, “Optically active oxygen-deficiency-related centers in amorphous silicon dioxide,” J. Non-Cryst. Solids 239, 16–48 (1998). [CrossRef]  

68. R. Tohmon, H. Mizuno, Y. Ohki, K. Sasagane, K. Nagasawa, and Y. Hama, “Correlation of the 5.0- and 7.6-eV absorption bands in SiO2 with oxygen vacancy,” Phys. Rev. B 39, 1337–1345 (1989). [CrossRef]  

69. H. Hosono, Y. Abe, H. Imagawa, H. Imai, and K. Arai, “Experimental evidence for the Si-Si bond model of the 7.6-eV band in SiO2 glass,” Phys. Rev. B 44, 12043–12045 (1991). [CrossRef]  

70. F. H. El-Batal, M. S. Selim, S. Y. Marzouk, and M. A. Azooz, “UV-vis absorption of the transition metal-doped SiO2-B2O-Na2O glasses,” Physica B 398, 126–134 (2007). [CrossRef]  

71. T. Akai and D. Chen, “Method for producing high silicate glass and high silicate glass,” U.S. patent 7,975,508B2 (July  12, 2011).

72. A. V. Kityk, K. Knorr, and P. Huber, “Liquid n-hexane condensed in silica nanochannels: a combined optical birefringence and vapor sorption isotherm study,” Phys. Rev. B 80, 035421 (2009). [CrossRef]  

73. P. Huber, M. Busch, S. Calus, and A. V. Kityk, “Thermotropic nematic order under nanocapillary filling,” Phys. Rev. E 87, 042502 (2013). [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 (6)

Fig. 1.
Fig. 1. (a) Dependence of anisotropic correlation volume ω f ( q ) on scattering wave vector q , illustrating the fractal correlations of half and integral numbers of fractal dimension D . The curve with D = 5 / 2 well approximates the slope of experimental data depicted by open circles, which represent scattered light intensity as a function of q with a monochromatic incident light from a laser source (514.2 nm) [37,38]. This comparison is justified by considering the fact that ω f ( q ) is proportional to the scattered intensity. For this comparison with data in [37,38], we have to intentionally set the correlation length ( ξ ) equal to 10 μm, which is about 10 3 times longer than our estimated value [40]. (b) Dependence of integrated dissymmetry factors ( Ω D ) for various values of D , where Ω D is proportional to the fractal turbidity τ f on 1 / λ 0 4 , indicating both τ f and Ω D are convex-upward functions of 1 / λ 0 4 .
Fig. 2.
Fig. 2. (a) Dependence of integrated spinodal correlation volume Ω s p on 1 / λ 0 4 , indicating Ω s p is a convex-downward function of 1 / λ 0 4 , with the parameter β = 0.251 nm 1 [2729] and the correlation length ζ = 19.95 nm , which was determined to fit best our data described in the text. (b) Dependence of corresponding turbidity τ s p on 1 / λ 0 4 , indicating τ s p is also a convex-downward function of 1 / λ 0 4 .
Fig. 3.
Fig. 3. (a) Dependence of spinodal correlation function γ s p on the radius r in nanometers. (b) Dependence of the corresponding correlation volume ω s p ( q ) on the scattering vector q . Both curves are estimated from the respective models, namely, Eq. (11) for (a) and Eq. (12) for (b), with parameters β = 0.251 nm 1 and ζ = 19.95 nm , both of which were determined to fit best the measured τ s p in the range of 350 to 850 nm.
Fig. 4.
Fig. 4. (a) Changes in the turbidity (estimated from the logarithm of the observed transmittance) as a function of the inverse fourth power of wavelength in air ( 1 / λ 0 4 ). The slight deviations of the turbidities for 1 and 3 from the λ 4 Rayleigh wavelength dependence are well fitted by the spinodal-decomposed turbidity curves, while that for 2 is well fitted by the one-parameter theoretical curve based on the fractal scattering with fractal dimension of 2.5. (b) Corresponding change in UV-Vis-light transmittance spectra of a porous Vycor glass after drying for 0, 45, and 165 min immediately after removal from ultrapure water immersion for 2 h at room temperature. In both (a) and (b), the previous results (Fig. 1 in [8]) are re-examined on the basis of the theory of dielectric constant fluctuations, and a comparison between the measured and fitted data is shown by solid and dotted lines, respectively.
Fig. 5.
Fig. 5. (a) Effects of changes in parameters ( β , ζ ) on the correlation volume ω s p ( q ) as a function of q . Each ω s p ( q ) curve corresponds to the parameter pair denoted by A: ( β [ nm 1 ], ζ [nm]) = (0.251, 19.95), B: (0.251, 10.2), C: (0.251, 29.0), D: (0.200, 19.95), and E: (0.300, 19.95), respectively. (b) Corresponding changes in the estimated UV-Vis-light transmittance spectra of a porous Vycor glass.
Fig. 6.
Fig. 6. (a) Measured and fitted UV-Vis light transmittance spectra of a porous Vycor in the dry state at 165 min, on the basis of the theory of dielectric constant fluctuations. The fitted curve is drawn with the parameters derived from the best fitting of τ s p in the range of 350 to 850 nm. (b) Difference between measured and fitted UV-Vis light absorbances, estimated from Abs ( λ 0 ) = log 10 { T fit ( λ 0 ) / T meas ( λ 0 ) } , as a function of the photon energy. The observed optical absorption spectrum is best fitted by three Gaussian curves, denoted as A, B, and C.

Tables (3)

Tables Icon

Table 1. Structural Information on Vycor 7930 Obtained by SAS Measurements a

Tables Icon

Table 2. Optical Absorption Bands Separated by Three-Gaussian Fitting

Tables Icon

Table 3. Correlation Volumes and Dissymmetrical Factors of Fractal Correlation Functions

Equations (61)

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

I ( q ) η 2 A V π 2 λ 0 4 V ω ( q ) ,
ω ( q ) = 0 4 π r 2 γ ( r ) sin ( q r ) q r d r ,
τ = η 2 A V π 2 λ 0 4 4 π d Ω 1 + cos 2 θ 2 ω ( q ) ,
γ f ( r ) = C f r d D exp ( r ξ ) ,
ω f ( q ) = C f 4 π ξ D Γ ( D 1 ) [ 1 + q 2 ξ 2 ] D / 2 [ 1 + q 2 ξ 2 ( q ξ ) 2 ] 1 / 2 sin [ ( D 1 ) tan 1 ( q ξ ) ] ,
f 5 / 2 ( q ) = 2 3 [ 1 + q 2 ξ 2 1 ] 1 / 2 [ 1 + q 2 ξ 2 + 2 ] q ξ [ 1 + q 2 ξ 2 ] 3 / 2 .
τ f = η f 2 A V k 4 ( 2 π ) 2 ε A V 2 ω 0 ( 5 / 2 ) · Ω 5 / 2 ,
Ω 5 / 2 = 2 π 0 π d θ sin θ 1 + cos 2 θ 2 f 5 / 2 ( 2 k sin ( θ 2 ) ) = 16 2 π 3 b 3 { h ( 1 + 1 + b , b ) h ( 2 , b ) } ,
h ( z , b ) = 2 z { z 3 7 z 2 5 z 3 ( 2 + b ) ( z 3 + 1 ) 1 + b + b 2 / 2 z 1 } .
g ( r ) 1 = A · sin ( β m r ) β m r exp ( r / ζ ) ,
γ s p ( r ) = C s p sin ( β r ) β r exp ( r / ζ ) ,
ω s p ( q ) = 8 π ζ 3 q 4 ξ 4 2 ( 1 β 2 ζ 2 ) q 2 ζ 2 + ( 1 + β 2 ζ 2 ) 2 ,
τ s p = η s p 2 A V k 4 ( 4 π ) 2 ε A V 2 · Ω s p ,
Ω s p = 2 π 0 π d θ sin θ 1 + cos 2 θ 2 ω s p ( 2 k sin ( θ 2 ) ) = ( 8 π b ) 2 ζ 3 [ 1 ( 1 2 + 1 a b ) ln { 1 + b 2 + 2 b ( 1 a ) ( 1 + a ) 2 } + b 8 a { 1 + ( 1 + 2 ( 1 b ) b ) 2 16 a b 2 } { tan 1 ( 1 a + b 2 a ) tan 1 ( 1 a 2 a ) } ] ,
η ( 0 ) η ( r ) = ( ε A V ρ ) S 2 ρ ( 0 ) ρ ( r ) .
T = I out I in = ( 1 r ) 2 exp ( τ ext · d opt ) ,
r = ( n air n pG n air + n pG ) 2 ,
T = exp ( τ f · d opt ) · ( 1 r ) 2 exp ( τ s p · d opt ) .
T r ( λ 0 ) T ( λ 0 , t ) / T ( λ 0 , t e ) .
τ f ( λ 0 ) = 1 d opt ln { 1 T r ( λ 0 ) } = η f 2 A V π 2 λ 0 4 ω 0 Ω 5 / 2 ,
τ s p ( λ 0 ) = 1 d opt ln { 1 T ( λ 0 , t e ) } + 1 d opt ln ( 1 r ) ,
η s p 2 A V = ( ε p ε s ) 2 φ ( 1 φ ) ,
0 = f · ε water ε p ( f ) ε water + 2 ε p ( f ) + ( 1 f ) · ε air ε p ( f ) ε air + 2 ε p ( f ) ,
ε pG ( f ) = ε SiO 2 [ ε p ( f ) + 2 ε SiO 2 + 2 φ { ε p ( f ) ε SiO 2 } ε p ( f ) + 2 ε SiO 2 φ { ε p ( f ) ε SiO 2 } ] ,
η f 2 A V = ( ε air ε water ) 2 f ( 1 f ) · φ ( 1 φ ) .
V ( Δ ρ ) 2 A V / ρ 0 2 = k B T A β T ,
( ε A V ρ ) S 2 ( Δ ρ ) 2 A V = ε A V 4 p · k B T A β T ,
n eff = n [ 1 i 2 π N k 3 S ( 0 ) ] ,
n eff = n [ 1 i 2 π N k 3 ( i k 3 α + 2 3 k 6 α 2 ) ] ,
exp { i k 0 ( c t n eff z ) } = exp ( i ω t ) exp { i k ( 1 + 2 π N α ) } exp ( 4 π 3 N k 4 α 2 · z ) ,
Abs ( E ) = μ = A , B , C I μ exp { ( E E μ ) 2 2 σ μ 2 } ,
γ ( r 1 , r 2 ) η ( r 1 ) η ( r 2 ) / η 2 A V
E in ( R , t ) = n ^ i E 0 exp [ i ( k i · R ω i t ) ] ,
E s ( R , t ) = k s × ( k s × n ^ i ) E 0 ε A V exp [ i ( k s R ω i t ) ] 4 π R × V d 3 r η ( r ) exp [ i ( k i k s ) · r ] ,
I s ( k i k s ) = 1 2 ε A V μ | E 0 | 2 | k s × ( k s × n ^ i ) | 2 ( 4 π R ) 2 ε A V 2 × v d 3 r 1 d 3 r 2 exp [ i ( k i k s ) · ( r 1 r 2 ) ] η ( r 1 ) η ( r 2 ) ,
I s ( q ) = I in k s 4 sin 2 χ ( 4 π R ) 2 · η 2 A V ε A V 2 V 0 4 π r 2 γ ( r ) sin ( q r ) q r d r ,
d P d Ω = 1 2 Re [ R 2 R ^ · E s × H s * ] = I s ( q ) R 2 I in V ,
τ = η 2 A V k s 4 ( 4 π ) 2 ε A V 2 4 π d Ω 1 + cos 2 θ 2 ω ( q ) = η 2 A V π 2 λ 0 4 4 π d Ω 1 + cos 2 θ 2 ω ( q ) ,
ω ( q ) = 0 4 π r 2 γ ( r ) sin ( q r ) q r d r .
f D ( q ) = ω f ( q ) ω 0 ( D ) = Γ ( D 1 ) Γ ( D ) [ 1 + q 2 ξ 2 ( q ξ ) 2 ] 1 / 2 sin [ ( D 1 ) tan 1 ( q ξ ) ] [ 1 + q 2 ξ 2 ] D / 2 ,
ω 0 ( D ) = 4 π ξ D · Γ ( D ) .
2 [ 1 + q 2 ξ 2 1 ] 1 / 2 q ξ
tan 1 ( q ξ ) q ξ
2 q ξ [ 1 + q 2 ξ 2 1 1 + q 2 ξ 2 ] 1 / 2
1 1 + q 2 ξ 2
2 3 [ 1 + q 2 ξ 2 1 ] 1 / 2 [ 1 + q 2 ξ 2 + 2 ] q ξ [ 1 + q 2 ξ 2 ] 3 / 2
1 [ 1 + q 2 ξ 2 ] 2
τ f ( D ) = η f 2 A V k 4 ( 4 π ) 2 ε A V 2 ω 0 ( D ) · Ω D ,
Ω D = 2 π 0 π d θ sin θ 1 + cos 2 θ 2 f D ( 2 k sin ( θ 2 ) ) .
Ω 1 / 2 = 16 2 π b 3 { f ( 1 + 1 + b , b ) f ( 2 , b ) } ,
f ( z , b ) = 2 z { z 5 11 5 z 4 9 + 8 z 3 7 4 z 2 5 b ( z 3 7 3 z 2 5 + 2 z 3 ) + b 2 2 ( z 3 1 ) } .
Ω 1 / 2 = 16 2 π 3465 b 3 [ 2 { 256 + 33 b ( 16 + 35 b ) } 1 + 1 + b { 128 ( 1 + 1 + b ) + 8 b ( 31 + 23 1 + b ) + b 2 ( 1822 795 1 + b ) } ] .
Ω 3 / 2 = 16 2 π b 3 { g ( 1 + 1 + b , b ) g ( 2 , b ) } ,
g ( z , b ) = 2 z { z 4 9 4 z 3 7 + 4 z 2 5 b ( z 2 5 2 z 3 ) + b 2 2 } .
Ω 3 / 2 = 16 2 π 315 b 3 [ 2 { 256 + 21 b ( 16 + 15 b ) } + 1 + 1 + b { 128 ( 1 + 1 + b ) + 8 b ( 19 + 11 1 + b ) + 259 b 2 } ] .
Ω 5 / 2 = 16 2 π 3 b 3 { h ( 1 + 1 + b , b ) h ( 2 , b ) } ,
h ( z , b ) = 2 z { z 3 7 z 2 5 z 3 ( 2 + b ) ( z 3 + 1 ) 1 + b + b 2 / 2 z 1 } .
Ω 5 / 2 = 16 2 π 315 b 3 [ 2 ( 768 + 560 b + 105 b 2 ) 1 + 1 + b 1 + b { 384 ( 1 + 1 + b ) + 8 b ( 53 + 29 1 + b ) + 145 b 2 } ] .
Ω 1 = 4 π 15 [ 22 b tan 1 ( b ) + 6 + 7 b b 2 ( 6 + 10 b + 15 b 2 ) b 3 ln ( 1 + b ) ]
Ω 2 = 4 π b 2 { ( 2 + b ) + 2 + 2 b + b 2 b ln ( 1 + b ) } .
Ω 3 = 4 π b 2 { ( 2 + b ) 2 1 + b 2 ( 2 + b ) b ln ( 1 + b ) } ,
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.