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

Photospectroscopically observed pore-space correlations of a wetting fluid during the drying process in nanoporous Vycor glass

Open Access Open Access

Abstract

We use light scattering to study spatial correlations in the pore space of Vycor glass upon draining a wetting fluid. We analyze the transmission spectrum of forward-scattered light on the basis of the theory of dielectric constant fluctuation, whereas conventional light scattering analyzes the scattered light at small angles of monochromatic incident light. Assuming that the drained pores, which are surrounded by filled pores, exhibit long-range correlations of a fractal dimension of 2.5, we analytically derive the corresponding turbidity. The slight deviation from the λ4 Rayleigh wavelength dependence directly provides the correlation length of the interconnected network of drained pores. The estimated length, ranging from 0.5 to 18 nm at most, is almost the same order as that indirectly estimated from our previous simple effective Rayleigh scatterer model.

© 2015 Optical Society of America

1. INTRODUCTION

Porous Vycor glass with nanopores is transparent in the visible region and is used in colorimetric chemical sensing when impregnated with selectively reacting regents [13]. The pores in Vycor have an average radius 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. Optical methods [4] 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 changes in the humidity of ambient air strongly affect its transmission. Although completely dried or wetted nanoporous Vycor glass is transparent for visible light even though there exist innumerable pores within it, Vycor during the drying process after it has been filled with a wetting fluid such as water vapor exhibits a transitory whitish turbidity phenomenon. The same phenomenon has been observed by several authors [59]. Page and co-workers [5,6] have observed that Vycor samples partially filled with hexane prepared by adsorption are transparent, but for samples prepared by desorption, the transmission is lower by several orders of magnitude. They attributed this lowering of transmission to “pore-space correlations” and explained the phenomenon qualitatively by the existence of the long-range correlations of empty pores in the pore space of draining Vycor with a quite long correlation length of about 10 μm. Furthermore, this turbidity phenomenon has more general character in nanoporous media, and as pointed out recently by Gruener et al. [10], the turbidity problem also appears at the interface during an imbibition process, not restricted to the drying process.

In our previous studies [11,12], we analyzed the wavelength dependence of the transitory whitish turbidity of nanoporous Vycor on draining water vapor, and found that the turbidity is dependent on the inverse fourth power of the wavelength (1/λ4), and that, from this dependence, the phenomenon can be consistently interpreted and quantitatively analyzed by a simple Rayleigh scattering model [11,12], which is based on the particle-scattering approach [13]. From the turbidity as a linear function of 1/λ4, we obtained the effective radius of a Rayleigh scatterer, which is regarded as a measure of the extent of inhomogeneities that cause the scattering [11,12]. The effective radius was found to be about 10 nm at most when the maximum of scattering occurs. This value is 1000 times smaller than that estimated by Page and co-workers [5,6]. Furthermore, the slight deviation of the observed turbidity from theoretical linear dependence on 1/λ4 cannot be explained at all by the previous particle-scattering viewpoint.

In general, the problem of light scattering can also be approached from Einstein’s fluctuation viewpoint, apart from the above particle-scattering viewpoint. In the fluctuation approach, the inhomogeneity present in a condensed medium such as porous glasses 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)ηAηBAV/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 of relating them to the scattered intensity has been treated by Debye and co-workers [14,15]. Their approach has been utilized both for small angle x-ray scattering and conventional light scattering. Their mathematical scheme is summarized as follows: for an unpolarized incident beam of light, the turbidity (τ) takes the form [14,15]

τ=η2AV·π2λ04·4πdΩ1+cos2θ2·ω(q),
where the scattering intensity has been summed over all solid angles dΩ with all scattering angles θ (the angle between the incident and scattered beams), λ0 is the wavelength in air, and ω(q) is the correlation volume, which is a function both of γ(r) and θ, defined as
ω(q)=04πr2γ(r)sin(qr)qrdr,
with the scattering vector q=2k0sin(θ/2) and the wave number k02π/λ0. A simple physical interpretation of Eq. (1) 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. A measure of the extent of 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) from the sample with a monochromatic incident light beam (or the dependence of the scattered intensity on wave vector q, i.e., I(q)) is measured to obtain the correlation function.

Several scattering determinations of the inhomogeneities of Vycor during drainage of a wetting fluid [5,6,16] can be found in the literature. By means of light scattering, which is a definitive probe of the long-range spatial correlations of the empty pores, Page and co-workers [5,6] reported that the draining pores, which are surrounded by filled pores, exhibit long-range correlations of a fractal dimension of about 2.6, and their length is larger than 10 μm. However, they did not report the correlation length as a function of the amount of wetting fluid in the pore space.

In this paper, we reexamine our results for the transitory whitish turbidity of nanoporous Vycor during the drying process 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 and co-workers [14,15]. The slight deviation from the λ4 Rayleigh wavelength dependence directly provides the correlation length of an interconnected network of drained pores.

2. THEORY—TURBIDITY ANALYSIS TO OBTAIN THE CORRELATION LENGTH

Below we outline some equations we have developed for analyzing scattering data in terms of fractal scattering and the corresponding turbidity. As mentioned above [5,6,16], when porous Vycor glass is partially saturated with a wetting fluid, a fractal-like percolation network may be formed, although the pores themselves are not distributed with fractal geometry. As the draining proceeds, the transmission becomes small due to strong scattering of light. This scattering may be characterized by the fractal-like percolation network with a fractal dimension of 2.6 [5,6] using the following fractal correlation function [16]:

γ(r)=Aexp(r/ξ)r3D.

Here, the parameter ξ, the correlation length, characterizes the average distance between fluctuations in dielectric constant. Inserting the above correlation function into Eq. (2), with analytically treatable fractal dimension (D) of 2.5, the correlation volume is given as ω(q)=ω0·f5/2(q), with a constant ω0(D)=4πξDΓ(D)=4πξ52(3π/4) and the dissymmetry factor

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

Integrating the correlation volume over all angles, the turbidity is given as

τ=η2AVk0416π2·ω0·Ω5/2,
where
Ω5/2=2π0πdθsinθ1+cos2θ2f5/2(2k0sinθ2)=22π3b3{h(1+4k02ξ2,b)h(1,b)}
with b=2k02ξ2 and
h(z,b)=2z{z37z25z11z12b(z3+1+1z1)2b2z1}.

Equation (5) implies that the wavelength dependence of the turbidity τ is determined mainly by 1/λ04 (which comes from the factor k04) and the correlation volume, ω0Ω5/2, which is a function of ξ only and determines the slight deviation from the above λ4 Rayleigh wavelength dependence.

3. RESULTS AND DISCUSSION

We have already reported the time-dependent change in the transmission spectrum of a porous glass chip in the UV–vis–near-IR region (300–2500 nm) during the drying process (see Fig. 1 in Ref. [11]). Here, we will replot the same data in a different way, since the original transmission of the porous Vycor glass showed several inherent characteristics, such as an ultraviolet absorption edge, absorption due to adsorbed organic contaminants, and transmission loss due to reflection at interfaces between the glass slab and air. The wavelength dependence of the turbidity derived from the original transmission reveals not only the λ04 dependence but also another type of wavelength dependence, characterized by the convex-downward curve of the turbidity versus 1/λ04. This convex-downward dependence can be explained by the spinodal decomposition process that takes place in Vycor as part of the manufacturing process, which will be discussed elsewhere. Here, in order to focus on the change in turbidity upon draining a wetting fluid, we separate the transmission change due to the drainage only from the inherent properties of the porous glass matrix by dividing all transmission spectra measured during the drying process by the final transmission spectrum measured at low relative humidity condition. That is, we analyzed the change in the relative transmission, defined as Tr(λ0)T(λ0,t)/T(λ0,te). Here, as a reference transmission, we used the final transmission data, T(λ0,te), which were measured at the last moment, te=165min after the removal from the immersion chamber. In this case, the turbidity change in the ratio, Tr(λ0), is considered to be purely caused by the drainage of a wetting fluid from pores deep inside the porous glass to the surface.

 figure: Fig. 1.

Fig. 1. (a) Change in UV–vis light transmission spectra of a porous Vycor glass chip relative to that in the dry state at 165 min, after drying for 0, 15, 30, 45, 60, 75, and 90 min immediately after removal from ultrapure water immersion for 2 h at room temperature. (b) Changes in the turbidity (estimated from the logarithm of the relative transmission) as a function of the inverse fourth power of wavelength in air (1/λ04). The slight deviation of the turbidity from the λ4 Rayleigh wavelength dependence is well fitted by the one-parameter theoretical curves based on the fractal scattering with fractal dimension of 2.5. In both (a) and (b), the previous results (Fig. 1 in Ref. [11]) are reexamined on the basis of the theory of dielectric constant fluctuation, and a comparison between the measured and fitted data is shown by dotted and solid lines, respectively.

Download Full Size | PDF

The time-dependent change in the relative transmission spectrum of a porous glass chip in the UV–vis region (300–850 nm) during the drying process is shown in Fig. 1(a). Measured data are shown by dotted lines and fitted ones by solid lines. The transmission curves were measured every 15 min immediately after the sample removal from the immersion from time 0 to 165 min. In the figure, only the first seven (0–90 min) curves are depicted. As shown in Fig. 1(a), the relative transmission of Vycor approaches the maximum of 1.0 at long wavelength because the transmission loss due to the reflection at the interfaces inherent to the original transmission data is canceled out when each transmission at time t is divided by the reference transmission at te=165min. However, only the initial relative transmission measured immediately after the removal from the immersion container (i.e., t=0min, dotted line 1) deviates considerably from the ideal transmission. This might be due to the light scattering by surface roughening caused by intense evaporation of a wetting fluid for the surface.

In the absence of absorption and reflection at interfaces, from Eqs. (1) and (5), the turbidity τ can be expressed as

τ(λ0)=1dln(1Tr(λ0))Δη2AV·π2λ04·ω0·Ω5/2,
where Tr(λ0) is the above relative transmission and d is the sample thickness in μm, and τ is in μm1.

To determine the parameter ξ by fitting the experimentally observed turbidity to the theoretical curve, we need to estimate the proportional coefficient η2AV. A dry Vycor glass can be considered to consist of two homogeneous phases. One of these may be pore regions of varying and undetermined shape distributed within a homogeneous matrix. Only two dielectric constants need to be considered: εs in the matrix and εp in the pores. If the pore is filled with air or water, then εp is equal to that of air or water (εair or εwater). The mean square variation of the dielectric constant is expressed as [17]

η2AV=(εpεs)2ϕ(1ϕ),
where ϕ is the porosity of the porous glass (for Vycor, porosity ϕ is about 0.3 [18]). Furthermore, the dielectric constant of the pores is assumed to be estimated on the basis of the effective medium expression:
εp=(1f)·εair+f·εwater,
where f is the pore-filling fraction, which is estimated from the absorbance peak as the ratio of α1900 to [α1900]max, i.e., fα1900/[α1900]max, as previously described in detail [11,12]. Since the reference transmission, T(λ0,te), is measured under the dry condition, i.e., f0, the change in the mean square variation of the dielectric constant Δη2AV can be approximated as
Δη2AV(εairεwater)2f(1f)·ϕ(1ϕ).
Here, f(1f)-type dependence on the filling fraction has already been proposed by Li et al. [16].

On the basis of the above consideration, the transmission profile between 300 and 850 nm is replotted as a function of 1/λ04 in Fig. 1(b), which shows the convex-upward curves as a function of the inverse fourth power of the wavelength in air. These convex-upward curves are predicted by the dielectric fluctuation theory, since the correlation volume, ω0Ω5/2, behaves as a convex-upward function of 1/λ04.

With a specified value of the pore-filling fraction f, each turbidity curve is fitted by the theoretical curve, which contains only one fitting parameter, ξ. Here, the theoretical curve is determined by combining Eqs. (8) and (11). In Figs. 1(a) and 1(b), a comparison between the measured and fitted data is shown by dotted and solid lines, respectively. The agreement between the theory and experiment in both figures is excellent.

Figure 2 shows the obtained correlation length ξ as a function of the pore-filling fraction f. The longest correlation length is estimated to be 18 nm at most, which is still one-tenth of the wavelength of the incident light.

 figure: Fig. 2.

Fig. 2. Correlation length ξ as a function of pore-filling fraction f, extracted from the peak absorbance at around 1900 nm, normalized by the initial maximum value measured immediately after removal from the immersion container [11]. The maximum correlation length reaches a value of about 18 nm at most at about f=0.6. The large error bar at f=0.97 is due to the deviation of measured curve 1 from the corresponding theoretical curve, as shown in Figs. 1(a) and 1(b). For comparison, the scatterer’s effective radius rsca, based on the previous Rayleigh scattering viewpoint [11,12], as a function of f is also included in the figure.

Download Full Size | PDF

Some authors [5,6,16] reported that the correlation length might extend to the order of several micrometers, especially Page and co-workers [5,6], who estimated that the correlation length of the fractal region was about ξ10μm, which was determined from the crossover at the scattering vector q0.1μm1 of the I(q) curve. This value seems to be long compared to the optical path length and the incident monochromatic wavelength (514.5 nm). All of these authors guessed such a large value of ξ on the basis of the consideration of evacuating a wetting fluid deep inside the pore space of Vycor to the surface. However, from the optical point of view, such a large extent of inhomogeneity leads to no light transmission from the sample, which is not the case. Thus, the correlation length on the order of micrometers might be an overestimation. Furthermore, if the correlation length becomes more than the incident light wavelength or an order of micrometers, the wavelength dependence of the scattering will behave as Mie scattering. However, this is also not the case. The experimental results [dotted lines in Fig. 1(b)] show that the deviation from the λ4 Rayleigh scattering dependence is quite small, which implies that the possibility of Mie scattering is negligibly small.

Figure 2 also shows that the maximum of the correlation length is observed at the filling fraction of about 0.6. In this f range approximately the capillary condensation takes place, which is just a crossover state between the entire filling (f=1) and adsorbed layer film (f0.4) regimes. The maximum ξ observed at about f0.6 corresponds to the appearance of long-range capillary bridges inside the pores.

Our analysis of the wavelength dependence of turbidity of forward-scattered light neglects the short-range atomic structure in Vycor and then provides the long-range correlations in pore space, whereas the scattering at a large angle is determined primarily by the short-range order correlations, which permit resolution of structures with atomic dimensions.

4. CONCLUSIONS

From the slight deviation from the λ4 wavelength dependence of the photospectroscopically observed turbidity of nanoporous Vycor glass, we estimated the correlation length, ξ, of the fractal-like network of drained pores on the basis of the theory of dielectric constant fluctuation. The correlation length is considered to be the extent of optical inhomogeneity that causes scattering. The correlation length ranged from 0.5 to 18 nm at most, which is almost the same order as that indirectly estimated from our previous Rayleigh scatterer’s model. Both pictures, i.e., the particle-scattering viewpoint and the fluctuation viewpoint, can reasonably explain the transient white turbidity phenomenon during the drying process in nanoporous Vycor.

ACKNOWLEDGMENTS

The authors are sincerely grateful to Dr. Yasuyuki Sugiyama of NTT for his 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. Y. Y. Maruo, “Measurement of ambient ozone using newly developed porous glass sensor,” Sens. Actuators B 126, 485–491 (2007). [CrossRef]  

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

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

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

6. 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]  

7. V. P. Soprunyuk, D. Wallacher, P. Huber, and K. Knorr, “Freezing and melting of Ar in mesopores studied by optical transmission,” Phys. Rev. B 67, 114105 (2003). [CrossRef]  

8. V. P. Soprunyuk, D. Wallacher, P. Huber, R. Ackermann, K. Knorr, and A. V. Kityk, “Optical transmission measurements of phase transitions of O2 and CO in mesoporous glass,” J. Low Temp. Phys. 134, 1043–1053 (2004). [CrossRef]  

9. D. Wallacher, V. P. Soprunyuk, A. V. Kityk, P. Huber, and K. Knorr, “Capillary sublimation of Ar in mesoporous glass,” Phys. Rev. B 71, 052101 (2005). [CrossRef]  

10. S. Gruener, Z. Sadjadi, H. E. Hermes, A. V. Kityk, K. Knorr, S. U. Egelhaaf, H. Rieger, and P. Huber, “Anomalous front broadening during spontaneous imbibition in a matrix with elongated pores,” Proc. Natl. Acad. Sci. USA 109, 10245–10250 (2012). [CrossRef]  

11. 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]  

12. 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]  

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

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

15. 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]  

16. 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]  

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

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

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 (2)

Fig. 1.
Fig. 1. (a) Change in UV–vis light transmission spectra of a porous Vycor glass chip relative to that in the dry state at 165 min, after drying for 0, 15, 30, 45, 60, 75, and 90 min immediately after removal from ultrapure water immersion for 2 h at room temperature. (b) Changes in the turbidity (estimated from the logarithm of the relative transmission) as a function of the inverse fourth power of wavelength in air (1/λ04). The slight deviation of the turbidity from the λ4 Rayleigh wavelength dependence is well fitted by the one-parameter theoretical curves based on the fractal scattering with fractal dimension of 2.5. In both (a) and (b), the previous results (Fig. 1 in Ref. [11]) are reexamined on the basis of the theory of dielectric constant fluctuation, and a comparison between the measured and fitted data is shown by dotted and solid lines, respectively.
Fig. 2.
Fig. 2. Correlation length ξ as a function of pore-filling fraction f, extracted from the peak absorbance at around 1900 nm, normalized by the initial maximum value measured immediately after removal from the immersion container [11]. The maximum correlation length reaches a value of about 18 nm at most at about f=0.6. The large error bar at f=0.97 is due to the deviation of measured curve 1 from the corresponding theoretical curve, as shown in Figs. 1(a) and 1(b). For comparison, the scatterer’s effective radius rsca, based on the previous Rayleigh scattering viewpoint [11,12], as a function of f is also included in the figure.

Equations (11)

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

τ=η2AV·π2λ04·4πdΩ1+cos2θ2·ω(q),
ω(q)=04πr2γ(r)sin(qr)qrdr,
γ(r)=Aexp(r/ξ)r3D.
f5/2(q)=23[1+q2ξ21]1/2[1+q2ξ2+2]qξ[1+q2ξ2]3/2.
τ=η2AVk0416π2·ω0·Ω5/2,
Ω5/2=2π0πdθsinθ1+cos2θ2f5/2(2k0sinθ2)=22π3b3{h(1+4k02ξ2,b)h(1,b)}
h(z,b)=2z{z37z25z11z12b(z3+1+1z1)2b2z1}.
τ(λ0)=1dln(1Tr(λ0))Δη2AV·π2λ04·ω0·Ω5/2,
η2AV=(εpεs)2ϕ(1ϕ),
εp=(1f)·εair+f·εwater,
Δη2AV(εairεwater)2f(1f)·ϕ(1ϕ).
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.