Encyclopedia of Glass Science, Technology, History, and Culture. Группа авторов

Чтение книги онлайн.

Читать онлайн книгу Encyclopedia of Glass Science, Technology, History, and Culture - Группа авторов страница 126

Encyclopedia of Glass Science, Technology, History, and Culture - Группа авторов

Скачать книгу

the distribution of interatomic distances in the measured sample, each of its peaks corresponding to a commonly occurring interatomic distance. For v‐SiO2, the first two peaks in the correlation function arise from the Si─O bond length and the O─O distance within SiO4/2 tetrahedra, respectively. Since these peaks are as sharp as for α‐quartz, both distances are as well defined in the glass as in the crystal. Within the limits of what is achievable experimentally, the Si─O and O─O coordination numbers determined from the areas of these peaks are essentially four and six, as expected for a fully connected CRN of SiO4/2 tetrahedra (Figure 3).

Schematic illustration of the ball-and-stick model constructed by Bell and Dean for SiO2 glass. Upper part: complete 614-atom model; lower part: small portion with higher magnification. Small dark spheres: silicons; large light spheres: oxygens. Graphs depict the diffraction results for α-quartz and v-SiO2. (a) Neutron diffraction measurements of their distinct scattering compared with the X-ray diffraction data for v-SiO2. Position of the first sharp diffraction peak indicated at Q1 ≈ 1.52 Å-1. (b) Neutron correlation function for α-quartz and v-SiO2, and X-ray correlation function for v-SiO2. Approximate positions of the short distances for a pair of connected tetrahedra indicated as (Si-O)1, and so on. Graph depicts the comparison between the Si-Ô-Si bond angle distributions, V(θ), in SiO2 glass, obtained from an analysis of the X-ray correlation function and in a recent NMR study.

      A nice example of the structural information that can now be drawn from advanced forms of microscopy is provided by v‐SiO2. For example, the amorphous region of a 2‐D layer of SiO2 on a graphene support has recently been imaged at the atomic scale (Figure 5b in Chapter 2.5). A bi‐tetrahedral layer is visible in the image where the nodes are the locations of Si atoms. Hence, the view is that of a layer of faces of tetrahedra whose similarity with the 2‐D representation of a random network shown in Figure 2 is striking.

      Early in its study, the structure of silica glass was described in terms of the so‐called microcrystalline model [13], and hence it is useful to mention it briefly here. Its starting point is that, although crystals have diffraction peaks that are much narrower than for glasses (Figure 5a), significant broadening is observed if crystallites are very small, in accordance with the Scherrer equation, ΔQ = 2πK/L, which relates the width of a Bragg peak ∆Q to the crystallite dimension L and to a shape factor K (~1).

      To account for its broad diffraction peaks, one might thus describe a glass in terms of very small crystallites. However, a problem with the model is that to explain the large widths of the observed glass diffraction peaks, the crystallite size should typically be on order of 5 Å, a value similar to unit‐cell dimensions. Philosophically, it makes no sense to consider crystallites as ordered entities if they contain only one unit cell, since there would be no translational symmetry. Furthermore, with such small crystallites, a crystalline powder would be composed almost entirely of grain boundary material, which by definition differs structurally from the bulk. Hence, microcrystalline models cannot provide a description of the structure of most of the material in the glass. For these two reasons, they need not be considered further.

      5.1 The Role of Network Modifiers

      If a second oxide with weaker, ionic bonding is added, then it can lead to some depolymerization of the network, as shown schematically for Na2O by the following reaction.

Chemical equation of the modifiers and bridging oxygens.

Скачать книгу