Sunteți pe pagina 1din 12

Properties of Glass Materials

M Hasanuzzaman, Norwegian University of Science and Technology (NTNU), Norway


A Rafferty, Trinity College Dublin, Ireland
M Sajjia, University of Limerick, Ireland
A-G Olabi, University of the West of Scotland, Paisley, UK
r 2016 Elsevier Inc. All rights reserved.

1 Definition of Glass 1
2 The Glass Transition 2
3 Structure of Glass 2
3.1 Random Network Model 3
4 Amorphous Phase Separation 3
4.1 Spinodal Decomposition 4
4.2 Nucleation and Growth 4
5 Chemical Durability of Glasses 5
5.1 Chemical Reaction Mechanism 5
5.2 Crystallization Effect on Chemical Durability 5
6 Mechanical and Thermal Properties 6
6.1 Viscosity 6
6.2 Strength 6
6.2.1 Strengthening of glass 6
6.3 Thermal Expansion Behavior 7
7 Electrical Properties 7
7.1 Resistivity 7
7.2 Dielectric Properties 8
8 Optical Properties 8
8.1 Refraction of Light 8
8.1.1 Effect of composition of glasses on refractive index 8
8.2 Absorption 8
9 Common Glass Systems 9
9.1 Soda-Lime Glass or Commercial Glass 9
9.2 Lead Glass 9
9.3 Aluminosilicate Glass 9
9.4 Borosilicate Glass 9
10 Effect of Composition on Glass Properties 10
10.1 Silicon Dioxide (SiO2) 10
10.2 Boron trioxide (B2O3) 10
10.3 Alkali Metal Oxides 10
10.4 Alkali Earth Metal Oxides 10
10.5 Other Property Modifying Oxides 10
10.6 Refining Component 11
10.7 Coloring Components 11
References 11
Further Reading 12

1 Definition of Glass

In everyday language the term glass designates a transparent substance, possessing the properties of hardness, rigidity, and
brittleness, and, apart from transparency, these are the typical properties one normally associates with a solid. Glass also possesses
a number of properties which are characteristic of the liquid state, and classification of glass as a liquid of very high viscosity rather
than a solid would be in accordance with modern views (McMillan, 1979). Unlike crystals, glass does not have a sharp melting
point, but like crystalline solids, glasses show elasticity (Paul, 1990).
Due to the complexity of the structure of glass, it is not altogether surprising that an exact, all-encompassing definition for
glass remains elusive, and instead a number of definitions have been suggested over the years. A more generally accepted definition

Reference Module in Materials Science and Materials Engineering doi:10.1016/B978-0-12-803581-8.03998-9 1


2 Properties of Glass Materials

is that offered by ASTM (C162) which states “a glass is an inorganic product of fusion which has cooled to a rigid condition
without crystallization.” However, the ASTM definition limits the definition of glass to inorganic constituents, which fails to
explain organic and molecular glasses that now represent a rapidly growing area of study (Varshneya and Mauro, 2010; Doremus,
1994). Newly developed solution based sol–gel synthesis of oxide materials occurs at much lower temperature than traditional
solid-state fusion processes and also allows powderless non-fusion processing of glasses. X-ray and electronic diffraction studies
have shown that glass lacks long-range periodic atomic arrangement, and every type of glass exhibits time-dependent glass
transformation behavior (Paul, 1990; Shelby and Lopes, 2005). Pointing this out, Varshneya and Mauro (2010) suggest a scientific
definition of glass as “a solid having a non-crystalline structure, which continuously converts to a liquid upon heating.” Zarzycki
(1991) favors a more simplified definition: “a glass is a non-crystalline solid exhibiting the phenomenon of glass transition.”
Thereby this definition also conveniently separates non-crystalline materials into the categories of glass and amorphous materials.

2 The Glass Transition

Glass is usually formed on solidification from the melting stage. The cooling is so rapid that crystallization does not have the time
to occur. With a further decrease in temperature, viscosity continues to increase, resulting in a progressive freezing of the liquid to
its final solidification. The relationship between crystal, liquid, and glass can be explained by means of specific volume as a
function of temperature (see Figure 1).

Figure 1 Schematic specific volume–temperature relationships for crystallization and glass formation. 1: liquid, 2: supercooled liquid, 3: glass on
fast cooling, 4: glass on slow cooling, 5: crystal; Ts: melting temperature, Tg: glass transition temperature.

On cooling from an elevated temperature (Ts), at the point of solidification (or freezing), two phenomena may occur. There is
either a discontinuous change in volume at the melting point if the liquid crystallizes, or crystallization is avoided and liquid
passes to a supercooled state (between Ts and Tg) and the volume of the liquid decreases steadily until there is a decrease in the
expansion coefficient at a range in temperature called the glass transformation range. Below this temperature range the glass
structure does not relax and the expansion coefficient is usually same as that for the crystalline solid. The intersection between the
curve of the glassy state and that for the supercooled liquid is known as the glass transition temperature, Tg (Kingery and Uhlmann
et al., 1976). This Tg varies with cooling rate (see Figure 1) and thus it is more accurate to call it a transformation range rather than a
fixed point (Lancry and Régnier et al., 2012). The glass transition temperature increases with increasing cooling rate. The specific
volume of the formed glass follows this same trend, increasing with increased cooling rate. With a slower cooling rate the time
available for the structure to relax increases and the supercooled liquid persists to a lower temperature resulting in a higher-
density glass.

3 Structure of Glass

A number of models exist to describe the structure of glass, including a random network model, a crystallite model, as well as other
structural models. Random network theory, which is originally based on the ideas of Zachariasen, is the most common model
Properties of Glass Materials 3

used for explaining glass structure. Here only the random network model is briefly discussed and readers are referred to the
following literature for details of other models (Zarzycki, 1991; Kingery et al., 1976; Uhlmann and Kreidl, 1983; Doremus, 1994).

3.1 Random Network Model


The random network model depicts glass as a three-dimensional network, lacking symmetry and periodicity, in which no unit of
the structure is repeated at regular intervals. For oxide glass, these networks are composed of oxygen polyhedra. Assuming that a
glass should have an energy content similar to that of the corresponding crystal, Zachariasen (1932) suggested four rules for the
formation of an oxide glass:

i. An oxygen atom should be linked to not more than two glass forming atoms.
ii. The coordination number of glass-forming atoms is low, 4 or less.
iii. Oxygen polyhedra share corners, not edges or faces.
iv. For 3-D networks, at least three corners of each polyhedron should be shared.

The bonding force responsible for the formation of the network structure in crystalline silica (SiO2) forms similar networks in
glass (see Figure 2(a)) and is considered as network formers. Alkali silicates form glass easily and when alkali oxides such as Na2O
are added to silica glass, the sodium ions become part of the structure and occupy random positions distributed throughout the
structure to provide charge neutrality (see Figure 2(b)).

Figure 2 Schematic two-dimensional glass structures: (a) amorphous SiO2 network; (b) SiO2 network modified through addition of Na2O.

These alkali oxides (e.g., Na2O, K2O, Li2O, CaO, MgO, and PbO) provide additional oxygen ions that modify the network
structure. For this reason, they are known as network modifiers. With an increase in the amount of modifiers, the ratio of oxygen to
silicon also increases and breaks up the three-dimensional network with the formation of singly-bonded oxygens that do not
participate in the network, thus reducing the length of the chain. The principal effect of these modifiers is a reduction in melting
temperature and working temperature by decreasing the viscosity. An excess of modifiers can make melting sufficiently simple and
mobile that crystallization occurs in preference during the formation of glass. Cations of higher valency and lower coordination
number than the alkalis and alkali earth metals may act as both network formers and modifiers and are referred to as inter-
mediates; for example, ZrO2, TiO2, Al2O3, ZnO.

4 Amorphous Phase Separation

The phenomenon of amorphous phase separation (APS) in glasses has, for decades now, been an important topic in glass research.
It has been the subject of many investigations and there is an entire area of glass science devoted to this phenomenon in different
glass-forming systems.
A homogeneous single phase separates into two or more phases of different compositions, if the free energy of the system with
two or more distinct phases is lower than that of the system with one single homogeneous phase (Paul, 1990). Phase separation
4 Properties of Glass Materials

that occurs above the glasses melting point (the liquidus line) is known as stable immiscibility, whereas phase separation at
temperatures below the liquidus is known as metastable immiscibility. There are two routes to the formation of discrete phases by
metastable phase separation; these are either by spinodal decomposition or by nucleation and growth process. The route taken
depends on the free energy of the mixing of the system. Figure 3 shows an immiscibility diagram for the binary sodium silicate
system when heat treated in different regions of the phase diagram. The dark phase in the figure represents a sodium rich
composition and light phase denotes a silica rich composition. Phase separation in type A and type C (see Figure 3) are in a
binodal region and undergo APS by nucleation and growth process, whereas type B is within the spinodal region and undergoes
APS by spinodal decomposition.

Figure 3 Metastable immiscibility diagram for the sodium silicate system showing typical morphologies (Wheaton and Clare, 2007).

4.1 Spinodal Decomposition


Because phase separation into the binodal and spinodal regions are achieved by means of different thermodynamic processes, the
morphology of the separated phases in each region are distinctly different. Phase separation in the spinodal region refers to a
continuous type of phase transformation in which the change begins as small fluctuations in composition, and the compositional
differences grow with time resulting in two continuous interpenetrating phases. Since these changes occur spontaneously and as
apparently no energy barrier is present to hinder the separation, the region is considered unstable. The system will lower its free
energy by continuously changing the composition of two phases until the equilibrium is reached. At this equilibrium, the free
energy is in the lowest state and compositions remain constant. Afterwards the phase-separated regions will grow in size through
diffusion. The boundary between phases is initially diffuse but sharpens with time. The phases formed in the spinodal region show
a high degree of connectivity (Wheaton and Clare, 2007).

4.2 Nucleation and Growth


In the binodal region, a large change in composition must occur to lower the free energy of the system, and thus form nuclei. This
region is referred as the metastable region, as the system is unstable to small fluctuations in composition but becomes stable at
Properties of Glass Materials 5

larger changes in composition. Once nucleated, the new phase will grow in size through diffusion. During the growth stage, the
composition of the nucleated phase is unchanged with time, for a constant temperature. The phase separation resulting from
nucleation and growth is characterized by spherical droplets, with low connectivity, of random size and distribution with a sharp
boundary existing between the two phases. Figure 4 shows etched surfaces of sodium borosilicate glass where the structures are
formed by (a) nucleation and growth and (b) spinodal decomposition.

Figure 4 Sodium borosilicate glass showing (a) a droplet microstructure and (b) a connective structure (Hasanuzzaman et al., 2013, 2014).

5 Chemical Durability of Glasses

Chemical durability is a property of a material that is measured by its ability to resist the corrosive action of water, aqueous
solution of acids, alkalis and salts. The chemical durability of glasses is superior to most metals and polymers. However, non-
silicate glasses are susceptible to dissolution in aqueous solution, especially with low pH solutions. Almost all kind of glasses
readily dissolve in hydrofluoric acid (HF) which attacks the network bonds. The glasses, which possess excellent chemical
durability, are also attacked significantly if exposed to very high or low pH, – particularly at elevated temperatures.

5.1 Chemical Reaction Mechanism


Chemical stability and the resistance of a glass to any chemical attack depend on factors like the solution type, exposure time and
temperature, and condition of the glass surface. There are several processes involved during chemical attack by an aqueous
solution on glass. However, every chemical attack involves dissociation of water into H þ or OH ions. Ion exchange occurs if the
glass contains alkali or other highly mobile ions, where the solution attacks the network bonds to keep the concentration ratio of
the components in glass and solution identical. This ion exchange process between highly mobile ions from glass and protonic
species from the liquid solution is known as congruent dissolution and requires the presence of a liquid or water vapor. Most of the
commercial glasses are silicate based glasses that contain alkali ions, and thus their dissolution process usually occurs by congruent
dissolution. The reaction products form a layer on the glass surface and influence the further dissolution rate. On the contrary,
non-aqueous solutions (i.e., organic) do not react with glass.
At extreme pH (o1 or 49), silicate based glasses dissolve at a higher rate where the ion exchange process and leached thin
silica gel layer is no longer the determining factor for the dissolution. The chemical attack occurs directly on the Si–O bonds at
these extreme pH conditions. For phase separated alkali borosilicate glasses, where two interconnecting phases: a silica-rich phase,
and an alkali-rich borate phase formed up on heat-treatment. Leaching is then used on the heat-treated glass to selectively remove
alkali borate phase which has very low durability even in weak acids, while silica phase is resistant to this weak acids. A post-leach
alkali wash step is needed to remove trapped silica gel and leave behind a silica-rich porous skeleton. The heat-treatment step and
leaching conditions can be adjusted to achieve the desired pore size, pore volume, and surface area. Zircon or zirconia can be used
to improve the durability of a porous glass (Kiefer, 1989; Marques, 2008; Kiefer, 1989; okubu et al., 1987; Paul, 1990; Hasa-
nuzzaman et al., 2014).

5.2 Crystallization Effect on Chemical Durability


The chemical durability of glasses can be influenced by crystallization if the crystalline phase has properties different to those of
the base glass. During crystallization, the mobile ions (i.e., Na þ or Li þ ) of the glass moved in to very durable crystals. In general,
6 Properties of Glass Materials

the durability of crystalline glass-ceramic is superior to that of amorphous glass. The durability depends on the nature of the
crystalline phases formed, the nature of the glassy matrix, and the overall microstructure of the resulting glass-ceramic.

6 Mechanical and Thermal Properties

Eventhough most glasses possess high hardness values compared to other materials, their brittle nature makes them vulnerable to
many engineering applications. Plenty of work has been done to improve the fracture toughness of glass. A number of mechanical
and thermal properties of glasses are discussed below which are determining factors for using glass in specific engineering
applications.

6.1 Viscosity
Viscosity is an important property of glass. Most processing techniques require the viscosity to be in a certain range (103–
107 dPa s) at a certain temperature. This is known as the working point. At temperature when viscosity of glass reaches 107.6 dPa  s is
defined as the softening point. At this point glass, deforms rapidly and is used for glassblowing or sintering. In the glass trans-
formation range (see Figure 1), most glasses show viscosity in the range of 1012–1013.5 dPa s, and the upper limit 1013 dPa s is
called the annealing point. Annealing is very important for relaxation of stresses generated during hot forming of glasses (Krause,
2005). At 30–40 1C below the annealing point, where the viscosity is 1014.5 dPa s, the relaxation of stress requires more time
(43 h) compare to the annealing point (o1 h). This is known as the strain point. The optimum stress relaxation temperature to use
should be chosen according to the application and the requirements of dimensional accuracy.
Below the glass transformation temperature (Tg), the viscosity of glass increases many orders of magnitude and the structure
becomes frozen-in (quasi-solid). The rate of cooling influences the density of the glass and also the formation of crystals. In
sodium silicate glasses, the most common glass system, increasing the alkali content breaks up the three-dimensional network
with the formation of singly-bonded oxygens that do not participates in the network, thus reducing the length of the chain. The
principal effect of increased alkali content is to reduce the melting temperature by decreasing the viscosity, which would indicate a
lower Tg. However, variations are observed for glass compositions containing B2O3, which is known as the boric oxide anomaly.
Increasing alkali content in sodium borate glass compositions converts BO3 triangles to BO4 tetrahedra, thereby increasing the
network connectivity. Therefore, the glass transition temperature increases. Non-bridging oxygen (NBO) sites do not become
apparent until the alkali content is increased to approximately 30 mol% (Vogel et al., 1985). In other studies with borosilicate
glass, it was mentioned that NBO’s are not generated in glasses with a Na2O/B2O3 ratio o0.5 (Saiki et al., 2010; Chen et al., 2004).
Replacement of alkali oxide (Na2O) by alkaline earth oxide (CaO), while maintaining a constant silica concentration, strengthens
the network by replacement of the low field strength sodium ion by the higher field strength divalent alkaline earth ions and thus
increases the Tg (Shelby and Lopes, 2005).

6.2 Strength
The high theoretical structural strength of glasses (ranges from 1 to 100 GPa), which depends on glass composition, is an attractive
property. However, this high strength of glasses cannot be utilized in practical applications due to presence of flaws/defects on its
surface which reduce the actual fracture strength by many orders of magnitude. These flaws, in the form of tiny chips or cracks, act
as stress concentrators. At the tips of these flaws, stress concentration may be introduced by a load which exceeds the theoretical
strength and cause fracture/breakage of the glass. Contact with other hard objects often cause flaws on glass surface. Chemical
attacks and thermal stresses introduced during rapid cooling may also generate flaws on glass surface. Annealing can also be a
reason for the reduction of strength as it may introduce small crystals to the glass surface. The thermal expansion mismatch
between the matrix glass and these crystals during cooling lead to flaws. In general, it is very challenging to keep glass surfaces free
from flaw formation during production process or during annealing. Lubrication or coatings are often applied to the fresh glass
surfaces to help prevent the formation of flaws. Reducing the flaw length below the Griffith critical crack length for crack growth,
by etching or polishing is another way to improve the glass strength.

6.2.1 Strengthening of glass


As discussed earlier the strength of glasses can be improved by preventing or removing the flaws from the surface which causes the
reduction of glass strength. The other way is to prevent the crack growth. Tensile stress concentrates on crack tip while mechanical
load is applied to glass and crack propagation occurs when it overcomes the residual compressive stress. Compressive strength of
glass which is very high can be exploited for strengthening. Internal stresses can be introduced to the glass by thermal tempering
(quenching of softened glass) or ion exchange. By air-quenching glass, compressive stresses can be formed on the surface while the
interior is held in tension. In this way, the strength of glass can be increased typically 4–6 times. Such glass is commonly used in
bus stops and automobile applications and it has the added advantage that the glass crumbles when broken, as opposed to
Properties of Glass Materials 7

forming dangerous shards. Ion-exchange or ‘chemical toughening’ works by replacing sodium ions, for example, with potassium
ions (which are 30% larger), by immersion of the glass into a bath of molten potassium nitrate. Chemical toughening can be
applied to glass objects of complex shapes. Oftentimes, polymeric materials (polycarbonates) are used in combination with glass
to form layered composite materials with enhanced properties, as is the case with bullet-proof glass, for example.

6.3 Thermal Expansion Behavior


The length and volume of a glass usually increases with increasing temperature. The thermal expansion curve measured by
dilatometry gives three important properties – the thermal expansion coefficient (a), glass transformation temperature (Tg), and
dilatometric softening temperature (Td). Coefficient of thermal expansion is a measure of the rate of expansion with temperature.
For glasses, the measured expansion coefficient is one directional and measured value is linear thermal expansion coefficient (aL). The
linear thermal expansion coefficient value of commercial glasses is measured over a specified temperature range and most cases
from 0 to 300 1C, 20 to 300 1C, or 25 to 300 1C (Shelby and Lopes, 2005). Figure 5 shows the linear thermal expansion curve of a
bulk sample of an alkali borosilicate glass (Hasanuzzaman et al., 2013). The glass expands until 305 1C, above which the glass
contracts suddenly up to about 425 1C, before expanding again. The contraction above 305 1C is believed due to bending of bonds
which absorbs the lattice expansion by changing position of atoms. The filling of interstices inhibits further contraction and causes
expansion again. Shelby and Lopes (2005) and Kingery et al. (1976) revealed a dramatic distinction in thermal expansion behavior
between quenched glasses and glasses which had been annealed prior to dilatometric thermal behavior characterization. Quen-
ched glasses showed very prominent contraction–expansion characteristics (see Figure 5). For annealed glasses the expansion–
contraction was minimal and to a large extent ‘washed out’ as a result of annealing the glass above Tg. The sudden expansion in the
temperature range 430 to 520 1C is referred to as the transformation range and indicates the onset of viscoelastic behavior, where
bond breaking is predominant. The onset temperature of the transformation range at 430 1C is attributed to Tg. The temperature of
maximum expansion is identified at 550 1C and can be referred to as the dilatometric softening temperature (Td). A sharp
contraction occurs at about 560 1C after the dilatometric softening temperature which corresponds to viscous flow of the sample
under stresses imposed by the dilatometric push rod. The linear thermal expansion coefficient of a glass is an essential property for
thermally induced stress formation. This property (linear thermal expansion coefficient) is utilized in tempering processes for
generating compressive surface layers in order to prevent flaw propagation, hence increasing strength.

Figure 5 Linear thermal expansion curve of alkali borosilicate glass obtained at 10 1C min  1 (Hasanuzzaman et al., 2013).

7 Electrical Properties

Glasses are well known for their excellent insulating properties. Most oxide glasses are ionic conductor, but very poor electrical
conductor due to low numbers of free monovalent ions. Due to its high insulating properties, glasses are used in the area of
electrical and electronics engineering for the production of seals, high-voltage insulators, microelectronic packaging, high-vacuum
tubes, lamps, etc.

7.1 Resistivity
Electrical conductivity in glasses depends on mobility of ions – especially alkali ions. Therefore, the conductivity of the alkali
silicate glasses increases with the increasing alkali oxide content. However, for the glasses containing two or more alkali oxides, the
8 Properties of Glass Materials

mixed-alkali effect occurs where the conductivity decreases. Addition of alkaline earth oxide to any alkali containing glass also
decreases the electrical conductivity. The immobile divalent alkali earth ions block and occupy the interstitial sites and thus reduce
the mobility of the alkali ions in a glass. The conductivity of a glass also very much depend on temperature. At room temperature,
the mobility of ions is so small that volume resistivity almost always lies beyond the range of measurement. The ionic mobility
increases with increasing temperature and as a result resistivity also decreases. Crystallization also influences the conductivity. If
the formed crystal removes alkali ions from the residual glass, the conductivity usually decreases. However, conductivity of the
glass-ceramic increases if the crystal phase is free of alkali ions.

7.2 Dielectric Properties


The dielectric constant (er) of a glass can be described as the relative increase of capacitances by introducing polarizable dielectric
into a condenser which is previously separated in vacuum. The dielectric constant of technical glasses usually lies between 4.5 and
8 which is within the range of other electrically insulating material (Krause, 2005). The dielectric constant of glasses primarily
depends on the electronic and ionic polarization and increases with temperature. However it decreases very slowly over a wide
range of frequency. Content of heavy metals in glass, such as lead and barium containing glasses, gives high values of dielectric
constant. If a dielectric material is situated in an alternative electromagnetic field, where reversing the polarities and shifting the
dipoles occurs, an inherent dissipation of electromagnetic energy causes heating and the loss quantifies as dielectric loss. The small
phase shift (δ) of practical performance to ideal loss free performance is parameterized as loss tangent (tan δ). Dielectric loss
commonly depends on relaxation and becomes high around the relaxation or resonance frequencies of the polarization. Dielectric
loss in the glasses correlates with its content, frequency, and temperature.

8 Optical Properties

The optical properties of glasses are of utmost importance and can be subdivided into three categories – refraction, absorption, and
transmission of light. Glass is among only few other solids that transmit light in the visible region of the spectrum. Therefore, the
development of optical glasses with appropriate refractive index and dispersion characteristics plays an important role in modern
scientific evolution, especially in nano-science, medical science, astronomy, and biology.

8.1 Refraction of Light


The refractive index (nl) of a material is defined as the ratio of velocity of light in a vacuum to that in a defined material. The
refractive index of a glass is not a constant, but depends on the wavelength of the incident light, known as dispersion. Dispersion, a
decisive factor for optical applications, explains the entire refractive index curve with the wave length of the light over the desired
range. Dispersion causes the separation of white or compound light into its constituent colors in prisms, and chromatic aberration
in optical lenses. Mean dispersion is the difference between the refractive indices of hydrogen, nF–nC (lF ¼ 486.1 nm,
lC ¼ 656.3 nm), measured at the F and C emission lines (Shelby and Lopes, 2005; Krause, 2005). For technical glasses, this
dispersion value lies between 0.007 and 0.013 (Krause, 2005).
The refractive indices of technical glasses are usually designated as nD at the yellow emission line of sodium (589.3 nm) or nd
for at yellow emission line of helium (587.6 nm). Due to very close proximity of these wavelengths, there is very little variance of
the refractive indices and they generally lie within the range 1.47–1.57. Only exception is for lead based glasses where PbO content
is 435% (nd ¼ 1.7) (Krause, 2005; Newton and Davison, 1989).

8.1.1 Effect of composition of glasses on refractive index


Increasing electron density or polarizability also increases the refractive index of a glass. Glasses containing low atomic number
ions, which have low electron density and low polarizabilities, also have low refractive indices. Non-bridging oxygens are more
polarizable than bridging oxygens. As a result, formation of non-bridging oxygens in a glass system due the compositional change
increases the refractive index of that glass and vice versa. For commonly used alkali silicate glasses, refractive indices increases with
increasing alkali oxide concentration. Density and refractive index are directly correlated, such that increasing the density of glass
also increases its refractive index.

8.2 Absorption
Ultraviolet (UV) light with wavelength (100–400 nm) cannot transmit through glass even if it is transparent. The electrons
attached to molecules in typical glasses can absorb radiation at UV wavelengths which are beyond their ultraviolet edge. Con-
version of a glass network from bridging state to non-bridging state shifts the ultraviolet edge to lower frequencies and toward the
visible region of the spectrum. A very intense absorption band can be achieved in glasses by adding small concentration of
impurities (i.e., iron) which result in a very intense absorption band due to transfer of electron from cations to neighboring
Properties of Glass Materials 9

anions. This absorption band is called charge transfer band. Absorption of light in the visible region of spectrum is known as color.
Coloring of glass can be created by a number of mechanisms, like (1) addition of either 3d transition metal ions of 4 f rare earth
(lanthanide) ions, (2) formation or precipitation of colloidal particles, (3) presence of charge transfer bands in the visible region of
spectrum, (4) light scattering due to phase separation or other optical defects.

9 Common Glass Systems

The primary glass formers in commercial oxide glasses are silica (SiO2), boron oxide (B2O3), and phosphorus pentoxide (P2O5),
all of which readily form single component glasses (Shelby and Lopes, 2005). Of these, other than silica, only boron oxide has
some commercial importance and only when mixed with silica. Silica is the most important glass former and silicate glasses
represent more than 95% of industrial glass production (Zarzycki, 1991). Silica-based glass is technically important for its excellent
chemical resistance (except HF and alkali) and small expansion coefficient which makes it a very good candidate for thermal shock
resistance (Zarzycki, 1991). Glass can be classified in different groups according to their intended usage or by their chemical
composition. The following sections describe the most common types of glass according to their chemical composition.

9.1 Soda-Lime Glass or Commercial Glass


Soda-lime glass is the most common commercial glass. It is comparatively inexpensive and amenable to recycling. A typical
composition of this glass is 70–75 wt% SiO2, 12–16 wt% of Na2O, and 10–15 wt% CaO (Bauccio, 1994; Pfaender, 1996). A small
percentage of other reagents can be added for specific properties and application requirements. The principal addition in this type
of glass, other than silica (SiO2), is sodium oxide or soda (Na2O). Even though sodium oxide contains oxygen atoms, it is held
together by ionic rather than covalent bonds. The sodium atoms in the mixture donate electrons to the oxygen atom, producing a
mixture of negatively charged oxygen ions and positively charged sodium ions. The oxygen atom with an extra electron binds to
one silicon atom and does not form a bridge between pairs of silicon atoms. Therefore, the melting temperature of the mixture is
considerably reduced (Bloomfield, 2001). Relatively high amount of alkali content in the glass also causes an increase of the
thermal expansion coefficient by about 20 times (Pfaender, 1996). Since sodium ions are so soluble in aqueous solution, calcium
oxide (CaO) is added to the mixture to improve its insolubility. Soda-lime glass is produced on a large scale and used for bottles,
drinking glasses, and windows. Its light transmission properties, as well as low melting temperature, make it suitable for use as
window glass. Its smooth and non-reactive surface makes it excellent as containers for food and drinks. Nowadays recycled glass,
also known as cullet, is used to make green glass, which helps to save energy and reduce emissions.

9.2 Lead Glass


Lead glass is similar to soda-lime glass where lime is replaced by a larger part of lead oxide (PbO). Lead glass typically contains 55–
65 wt% SiO2, 18–38 wt% of PbO, and 13–15 wt% Na2O or K2O (Bauccio, 1994; Pfaender, 1996). Lead glass is usually used for
decorative glassware. It is also included in special optical glasses for their high refractive index. The networks in lead glass are more
complete than those in soda-lime glass and thus they are stronger and have less internal friction (Bloomfield, 2001). Lead oxide
also makes the glass dense, hard, and X-ray absorbing, and therefore suitable for use in radiation shielding.

9.3 Aluminosilicate Glass


Aluminosilicate glasses are usually prepared from a ternary system with a typical composition 52–58 wt% SiO2, 15–25 wt% of
Al2O3, and 4–18 wt% CaO (Bauccio, 1994). With low thermal expansion and high softening temperature, this glass can tolerate
high temperature better than soda-lime glass and is used in thermometers, combustion tubes, cookware, halogen lamps, furnaces,
and fiberglass insulation.

9.4 Borosilicate Glass


Borosilicate glass contains substantial amounts of silica (SiO2) and boron oxide (B2O348%) as glass network formers, and are
typically composed of 70–80 wt% SiO2, 7–13 wt% of B2O3 4–8 wt% Na2O or K2O, and 2–8 wt% of Al2O3 (Bauccio, 1994;
Pfaender, 1996). Glass containing 7–13 wt% of B2O3 is known as low-borate borosilicate glass, and is mainly used to produce
chemical apparatus, lamps, and tube envelopes. Glasses containing 15–25% B2O3, is known as high-borate borosilicate glass.
High-borate borosilicate glass is also known as leachable alkali-borosilicate glass with an optimum composition of 62.7 wt% SiO2,
26.9 wt% of B2O3, 6.6 wt% Na2O, and 3.5 wt% of Al2O3 (Elmer, 1992). This glass can be further processed to produce Controlled
Pore Glass (CPG) which is widely used as a stationary media in chromatography, or alternatively, the pores can be closed up to
yield a clear impervious glass known as Vycor 96% silica glass, commonly used in cookware. The increase of B2O3 content,
10 Properties of Glass Materials

coupled with a very fine-scale secondary phase separation within the silica phase increases the chemical resistance, and in this
aspect high-borate borosilicate glass differs greatly from low-borate.

10 Effect of Composition on Glass Properties

The major compounds in borosilicate glass are silicon dioxide and boron trioxide. Alkali metal oxides and alkali earth metal
oxides are added up to a certain level in order to optimize the properties. Some other oxides, for example, Al2O3, TiO2, ZrO2, are
added to achieve specific properties depending on end-user requirements. In some cases coloring and refining components are
also added.

10.1 Silicon Dioxide (SiO2)


Silicon dioxide is the glass former in borosilicate glass composition. Higher levels of silica increases the melting temperature as
well as the working point, and reduces the coefficient of thermal expansion. With lower levels of silica the resistance to acids
deteriorates (Peuchert et al., 2004).

10.2 Boron trioxide (B2O3)


Boron trioxide reduces melting and working temperatures and improves hydrolytic stability when used below 13% by weight in
the composition (Reben and Li, 2011). Higher boron trioxide contents have an adverse effect on acid resistance and increases
highly volatile alkali metal borates. On the other hand, lower borate contents increase the melting point of the glass by creating
secure bonds with alkali metal ions. This helps to reduce the susceptibility to crystallization (Peuchert et al., 2004). Borate also play
a major role in reducing the dielectric constant of glass (Reben and Li, 2011).

10.3 Alkali Metal Oxides


Sodium oxide (Na2O) is widely used as a flux, especially in borosilicate glass composition, along with other alkali metal oxides
like potassium dioxide (K2O), lithium dioxide (Li2O), and lead oxide (PbO). Alkali metal oxide reduces the working temperature
and plays an important role in setting the thermal expansion. If the alkali metal oxides content is above a certain limit, glasses
exhibit a high coefficient of thermal expansion (Marques, 2008). A higher level of alkali oxide also causes an adverse effect on
hydrolytic stability (Peuchert et al., 2004). Peuchert et al. (2004) also discussed the role of alkali metal oxides on crystallization
and suggested using at least two alkali metal oxides, even in small amounts, in order to have a positive effect on resisting unwanted
crystallization. They also reported that beyond 1000 1C, potassium borates evaporate more easily than sodium borates, while
lithium borates are, by comparison, more stable in terms of evaporating when the glass is heated (Shelby and Lopes, 2005). Li þ
forms strong bonds with the glass network and increases the acid resistance of the glass (Marques, 2008). The use of PbO, which is
an excellent flux, is becoming limited these days due to concerns regarding lead toxicity.

10.4 Alkali Earth Metal Oxides


Calcium oxide (CaO) is most commonly used as a property modifier component in glass composition. Small amounts of
magnesium oxide (MgO), zinc oxide (ZnO), strontium oxide (SrO), and barium oxide (BaO) are also added separately based on
application requirements. Calcium oxide can help greatly to accelerate the phase separation of borosilicate glasses (Eguchi et al.,
1988; Kokubu et al., 1987; Yazawa et al., 1994). It also has a stabilizing effect on acid resistance (Watzke et al., 1998). It has been
found that limiting CaO to small amounts reduces the evaporation of highly volatile sodium and potassium borate compounds
during hot forming (Peuchert et al., 2004). If the amount of CaO exceeds certain limit, devitrification is likely to take place.
Moreover, heat resistance and alkali resistance also deteriorate with high contents of CaO (Kokubu et al., 1987). Eguchi et al.
(1988) has proposed using ZnO to promote ZrO2 retention in the SiO2 rich phase at a time of phase separation, thereby
maximizing ZrO2 content in the porous silica-rich skeleton for enhanced alkali-durability.

10.5 Other Property Modifying Oxides


Zirconium dioxide (ZrO2), aluminum trioxide (Al2O3), and titanium dioxide (TiO2) are other commonly used property-mod-
ifying oxides. It has been shown that ZrO2 greatly improves the alkali resistance of borosilicate glass without the glass suffering in
terms of the hydrolytic stability and resistance to acids (Kiefer, 1989; Marques, 2008; Kiefer, 1989; Kokubu et al., 1987; Paul,
1990). However, higher ZrO2 contents increase the working point as well as the risk of flaws forming in the glass. With excessively
high ZrO2 contents, crystallization is likely to occur (Eguchi et al., 1988). Although small amounts of ZrO2 suppress devitrification,
they reduce the growth rate of phase separation due to an increase of glass viscosity (Kukizaki, 2010). There is also a risk of passing
Properties of Glass Materials 11

un-melted ZrO2 into the product (Peuchert et al., 2004). Kord et al. (2009) claim that addition of ZrO2 (6 mol%) almost doubled
the bending strength. Alumina, most commonly used as a modifying oxide (Eguchi et al., 1988), has mixed effect on the property
of the borosilicate glass. It increases insolubility as well as resistance to devitrification of the borosilicate glass, while inhibiting
phase separation (Marques, 2008; Rose et al., 2011; Doremus, 1994). Fluidity of the glass decreases with the addition of Al2O3 in
these glass compositions (Marques, 2008). Alumina also acts as a balancing component to allow relatively high levels of CaO in
the borosilicate system. Alumina also acts to prevent cracks during leaching (porous glass production) and improve moldability by
minimizing the viscosity change relative to the change in temperature (Kokubu et al., 1987). Titania is used in small quantities for
special applications where UV absorption is required (Peuchert et al., 2004; Shorrock and Yale, 1993). It is also believed that the
presence of TiO2 suppresses crystallization (Mojumdar, 2004).

10.6 Refining Component


Borosilicate glasses can be refined with small amounts (r1% by weight) of conventional refining agents, such as chlorides, for
example, NaCl, and sulphates, Na2SO4 or BaSO4, fluorides, and bromides. Amongst other refining agents CeO2, As2O3, Sb2O3,
and CaF are noteworthy (Yazawa et al., 1999; Peuchert et al., 2004). These refining components are also used to remove
undesirable color tinges.

10.7 Coloring Components


Fe2O3, Cr2O3, and CoO are the most commonly used coloring components in glasses. Oxides of other transition elements
(copper, manganese, nickel, vanadium, titanium) and rare earths (mainly neodymium and praseodymium) are also used to obtain
different colored glasses, usually o3 wt% (Pfaender, 1996).

References

Bauccio, M.L. (Ed.), 1994. Engineered Materials Reference Book, second ed. USA: ASM International.
Bloomfield, L., 2001. How Things Work: The Physics of Everyday Life, second ed. New York: Wiley.
Chen, D., Miyoshi, H., Masui, H., Akai, T., Yazawa, T., 2004. NMR study of structural changes of alkali borosilicate glasses with heat treatment. Journal of Non-Crystalline
Solids 345−346, 104–107.
Doremus, R.H., 1994. Glass Science, second ed. New York, NY: John Wiley & Sons.
Eguchi, K., Tanaka, H., Yamaguro, T., Yazawa, T., 1988. Chemically durable porous glass and process for the manufacture thereof. US Patent: 4,778,777.
Elmer, T.H., 1992. Porous and Reconstructed Glasses. Engineered Materials Handbook 4, 427–432.
Hasanuzzaman, M., Rafferty, A., Olabi, A.G., 2014. Effects of zircon on porous structure and alkali durability of borosilicate glasses. Ceramics International 40, 581–590.
Hasanuzzaman, M., Sajjia, M., Rafferty, A., Olabi, A.G., 2013. Thermal behaviour of zircon/zirconia-added chemically durable borosilicate porous glass. Thermochimica Acta
555, 81–88.
Kiefer, W., 1989. Borosilicate glass. US Patent: 4,870,034.
Kingery, W.D., Uhlmann, D.R., Bowen, H.K., 1976. Introduction to Ceramics, second ed. New York: Wiley.
Kokubu, Y., Chiba, J., Saita, K., 1987. Porous glass, process for its production and glass material used for the production. US Patent: 4,665,039.
Kord, M., Marghussian, V.K., Eftekhari-Yekta, B., Bahrami, A., 2009. Effect of ZrO2 addition on crystallization behaviour, porosity and chemical−mechanical properties of a
CaO−TiO2−P2O5 microporous glass ceramic. Materials Research Bulletin 44 (8), 1670–1675.
Krause, D., 2005. Glasses. In: Martienssen, W., Warlimont, H. (Eds.), Springer Handbook of Condensed Matter and Materials Data. Berlin Heidelberg: Springer, pp. 523–572.
Kukizaki, M., 2010. Large-scale production of alkali-resistant Shirasu porous glass (SPG) membranes: Influence of ZrO2 addition on crystallization and phase separation in
Na2O−CaO−Al2O3−B2O3−SiO2 glasses; and alkali durability and pore morphology of the membranes. Journal of Membrane Science 360 (1−2), 426–435.
Lancry, M., Régnier, E., Poumellec, B., 2012. Fictive temperature in silica-based glasses and its application to optical fiber manufacturing. Progress in Materials Science 57 (1),
63–94.
Marques, P., 2008. Borosilicate glass compositions and uses thereof. US Patent: 7,341,966.
McMillan, P.W., 1979. Glass-Ceramics. London: Academic Press.
Mojumdar, S.C., 2004. The Microstructure and optical transmittance thermal analysis of sodium borosilicate bio-glasses. Journal of Thermal Analysis and Calorimetry 78 (1),
145–152.
Newton, R.G., Davison, S., 1989. Conservation of glass, for series: Butterworths series in conservation and museology. Butterworths.
Paul, A., 1990. Chemistry of Glasses, second ed. London; New York: Chapman and Hall.
Peuchert, U., Kunert, C., Bartsch, R., 2004. Borosilicate glass with high chemical resistance and use thereof. US Patent: 6,794,323.
Pfaender, H.G., 1996. Schott Guide to Glass. London: Chapman & Hall.
Reben, M., Li, H., 2011. Thermal stability and crystallization kinetics of MgO−Al2O3−B2O3−SiO2 glasses. International Journal of Applied Glass Science 2 (2), 96–107.
Rose, P.B., Woodward, D.I., Ojovan, M.I., Hyatt, N.C., Lee, W.E., 2011. Crystallisation of a simulated borosilicate high-level waste glass produced on a full-scale vitrification
line. Journal of Non-Crystalline Solids 357 (15), 2989–3001.
Saiki, K., Sakida, S., Benino, Y., Nanba, T., 2010. Phase separation of borosilicate glass containing sulfur. Journal of the Ceramic Society of Japan 118 (7), 603–607.
Shelby, J.E., Lopes, M. (Eds.), 2005. Introduction to Glass Science and Technology. Cambridge: The Royal Society of Chemistry.
Shorrock, P., Yale, B., 1993. Borosilicate glass composition. US Patent: 5,219,801.
Uhlmann, D.R., Kreidl, N.J., 1983. Glass-forming Systems. New York: Academic Press.
Varshneya, A.K., Mauro, J.C., 2010. Comment on Misconceived ASTM Definition of "Glass" by A.C. Wright. Eur. J. Glass Sci. Technol. A 51 (1), 28–30.
Vogel, W., Kreidl, N.J., Lense, E., 1985. Chemistry of Glass. Columbus, Ohio: American Ceramic Society.
Watzke, E., Kampfer, A., Brix, P., Ott, F., 1998. Borosilicate glass of high chemical resistance and low viscosity which contains zirconium oxide and lithium oxide. US Patent:
5,736,476.
12 Properties of Glass Materials

Wheaton, B.R., Clare, A.G., 2007. Evaluation of phase separation in glasses with the use of atomic force microscopy. Journal of Non-Crystalline Solids 353 (52−54),
4767–4778.
Yazawa, T., Kuraoka, K., Du, W., 1999. Effect of Cooling Rate on Pore Distribution in Quenched Sodium Borosilicate Glasses. The Journal of Physical Chemistry B 103 (45),
9841–9845.
Yazawa, T., Tanaka, H., Eguchi, K., Yokoyama, S., 1994. Novel alkali-resistant porous glass prepared from a mother glass based on the SiO2−B2O3−RO−ZrO2 (R ¼ Mg, Ca,
Sr, Ba and Zn) system. Journal of Materials Science 29 (13), 3433–3440.
Zachariasen, W.H., 1932. The atomic arrangement in glass. Journal of the American Chemical Society 54 (10), 3841–3851.
Zarzycki, J., 1991. Glasses and the Vitreous State. Cambridge; New York: Cambridge University Press.

Further Reading

C162, C162, 1945. Compilation of ASTM Standard Definitions. The American Society for Testing Materials, Philadelphia PA.

S-ar putea să vă placă și