You are on page 1of 41

Chapter 8: Diffusion and electrical conduction in glasses

8.1 Introduction
8.2 Fundamentals of Diffusion
8.3 Ionic Diffusion in Glasses
8.4 Examples of Diffusion and Ionic Conduction
8.5 Impedance Spectroscopy of Glasses
8.6 Electrical Properties of Glass
8.7 Dielectric Relaxations in Glasses
Introduction

Atomic/ionic diffusion through the network controls:


• Electrical conductivity of glasses containing monovalent ions.
• Chemical dissolution.
• Ion exchange reactions.
• Dielectric and mechanical losses.
• Gas permeation.
• Diffusion-controlled reactions.
Fundamentals of Diffusion
Steady-State Diffusion

• Diffusion is time-dependant i.e. the amount of an element that is


transported within another depends on time.
• How fast diffusion occurs (rate of mass transfer) is called diffusion flux.

M J = diffusion flux (units: kg.m-2.s-1 or atoms.m-2.s-1)


J= M = mass (or number of atoms)
At
A = area across which diffusion is occurring (m2)
t = time (s)

Area A
If the diffusion flux does not
Diffusion flux J change with time, steady-
state diffusion exists
• Consider a thin metal plate, with a gas
diffusing through it.
• If PA and PB are kept constant, the plot of
C vs. x is linear and constant with time.
• C vs. x plot is the concentration profile.
• the slope is the concentration gradient.

dC
concentration gradient =
dx

For the present case:


∆C CA − CB
concentration gradient = =
∆x xA − xB

Units of concentration: kg/m3 or g/cm3

4
The mathematics of diffusion in a single (x) direction is given by Fick’s First Law

dC
J x = −D
dx

Jx = flux (flow rate) of diffusing species


in x-direction.
D = diffusion coefficient or diffusivity
(units: m2/s).
Concentration gradient is C = concentration.
the driving force which x = distance.
makes diffusion occur. dC concentration gradient of diffusing species
=
dx in the x direction

5
Non steady-State Diffusion

Diffusion path
The concentration gradient at any position on the
diffusion path changes with time.

t=0
∂C ∂  ∂C  Fick's second law
=x
D  x

∂t ∂x ∂x 

t1 > 0 Cx = concentration at position x.

If D is independent of C, then:

t2 > t1 ∂C ∂C 2

=D
x x

∂t ∂x 2

Position x
6
• To find the diffusion coefficient of alkali ions in glasses, a thin layer of a radioactive isotope of
the ion is placed on the surface of the glass and heated at a fixed temperature for a fixed time.
• The concentration profile of the radioisotope is then analyzed.
• Concentration vs. distance data are then fitted to:

c = concentration
𝑄𝑄 𝑥𝑥 2
𝑐𝑐 = 𝑒𝑒𝑒𝑒𝑒𝑒 − x = distance
𝜋𝜋𝜋𝜋𝜋𝜋 4𝐷𝐷𝐷𝐷 Q = concentration of isotope per unit area on the glass surface
D = diffusion coefficient

Q and t are constants, so plotting log c vs. x2 allows us to


calculate Q from the slope of the plot.

Mehrer et al. J Phys: Conf Ser 106 (2008) 012001


If the source of the diffusing species is a melt or a gas, the concentration profile is given by:

𝑥𝑥 Co = surface concentration (constant)


𝑐𝑐 = 𝐶𝐶𝑜𝑜 𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒
4𝐷𝐷𝐷𝐷 erfc = error function

The diffusion coefficient is obtained from a best fit to the concentration profile.

• If the diffusing species is a gas such as He, one face of a


plate of thickness L can be exposed to a known pressure of
gas while the other face can be held at zero pressure.
• In this case, steady-state flow is reached and

Δ𝑐𝑐 Δ𝑐𝑐
𝐽𝐽 = −𝐷𝐷 = slope of the linear concentration
𝐿𝐿 𝐿𝐿 gradient between the two faces

Doremus, J Phys Chem 68 (1964) 2212


If the gas concentration follows Henry’s Law: c = SP S = solubility coefficient (or solubility)
P = gas pressure
then 𝑆𝑆 𝑃𝑃2 − 𝑃𝑃1 P1 = outside face, P2 = inside face
𝐽𝐽 = −𝐷𝐷
𝐿𝐿
If a vacuum is maintained on the inside face, then P2 = 0 and P1 = applied pressure P. Hence:

𝐽𝐽𝐽𝐽
= 𝐾𝐾 = 𝐷𝐷𝐷𝐷 K = permeability coefficient (or permeability)
𝑃𝑃
The permeability allows the steady-state flow rate of gas through a sample to be calculated,
when the dimensions of the sample and the applied gas pressure are known.

If the electrical conductivity is the result of field-induced diffusion of a single ionic species, the
electrical conductivity of the sample is related to the diffusion coefficient of that species by the
Nernst-Einstein equation:
σ = conductivity
Z = ionic charge
F = Faraday constant
𝑍𝑍 2 𝐹𝐹 2 𝐷𝐷𝐷𝐷
𝜎𝜎 = c = concentration of diffusing species in the glass
𝑓𝑓𝑓𝑓𝑓𝑓 D = diffusion coefficient
R = gas constant
T = absolute temperature
f = Haven ratio (≈ 0.2 – 1.0)
• The Haven ratio is the ratio of the tracer diffusion coefficient for the ion DT to the diffusion
coefficient calculated from the ionic conductivity, Dσ.
• Dσ is given by the Einstein equation:

𝑘𝑘𝑘𝑘𝜎𝜎
𝐷𝐷𝜎𝜎 = 2
𝑐𝑐𝑍𝑍

The temperature dependence of the diffusion coefficient is given by:


−∆𝐻𝐻𝐷𝐷 D0 = constant
𝐷𝐷 = 𝐷𝐷0 𝑒𝑒𝑒𝑒𝑒𝑒
𝑅𝑅𝑅𝑅 ∆HD = activation energy for diffusion.

Putting this equation into the Nernst-Einstein equation and combining all the constants into a single
term gives:

−∆𝐻𝐻𝐷𝐷
𝜎𝜎𝑇𝑇 = 𝜎𝜎0 𝑒𝑒𝑒𝑒𝑒𝑒
𝑅𝑅𝑅𝑅

This equation describes the temperature dependence of ionic conductivity.


This equation doesn’t always fit the experimental data, so it is often rewritten as:

−∆𝐻𝐻𝐷𝐷 Note: σ0 and ∆HD have different values from the terms in the
𝜎𝜎 = 𝜎𝜎0 𝑒𝑒𝑒𝑒𝑒𝑒
𝑅𝑅𝑅𝑅 previous equation.

Both expressions are found in the literature, so be careful when calculating σ from values of σ0
and HD.
Conductivity graphs may also be plotted as log(σ) vs. 1/T or log(σT) vs. 1/T.

Ionic Diffusion in Glasses

• Most studies have been of highly mobile ions with a radioactive isotope for tracer measurements.
• Most data is for sodium, followed by potassium, rubidium and cesium.
• Lithium does not have a radioactive isotope.
• Divalent (or higher valency) ions have very low mobilities, so are difficult to study.
• Sodium diffusion coefficients increase with
Na2O concentration in Na2O-SiO2 glasses.
• Large scatter in the data.
• Differences in thermal history of the glass e.g.
cooling rate or annealing, as well as
differences in –OH group content, can affect
sodium diffusivity.
• As Na2O concentration ↑, ∆HD ↓. D0 does not
change much.
• Potassium diffusion in K2O-SiO2 glasses has
similar behavior.
• Data for Rubidium and cesium diffusion in Rb2O-SiO2 and Cs2O-SiO2 glasses respectively is very
limited, but follows the same trends as Na+ and K+ diffusion.
• Comparing the diffusion coefficients of alkali ions in their corresponding R2O-3SiO2 glass,
diffusion coefficient decreases slightly in the order Na+ > K+ > Rb+ > Cs+.
• The difference (one order of magnitude at 400°C) is smaller than expected from the large
difference in ionic radii.
• Self-diffusion: the diffusion of an ion which is a primary component of the glass.
• Can also measure diffusion of an ion which is not a component of the glass e.g K+ diffusion in a
Na2O-SiO2 glass.
• For R2O-3SiO2 glasses (R = Na, K, Rb, Cs) Impurity ions always have smaller diffusion coefficients
than component ions.
• Diffusion coefficient of the impurity ion decreases with increasing size difference between
impurity and component ions.
• For a Na2O-3SiO2 glass, diffusion coefficient decreases in the order Na+ > K+ > Rb+ > Cs+.
• For a Cs2O-3SiO2 glass, diffusion coefficient decreases in the order Cs+ > Rb+ > Na+.
• Addition of a divalent modifier e.g. Ca2+ to a Na2O-SiO2 glass decreases the Na+ diffusion
coefficient.
• The Ca2+ ions have low mobility. They occupy interstices in the network and block the more
mobile alkali ions.
• This helps to improve the chemical durability of alkali silicate glasses.
• Na+ diffusion in Na2O-GeO2 glasses is quite different.
• Na+ diffusion coefficient decreases with increasing Na2O concentration, passing through a
minimum at 10-20 mol % Na2O.
Examples of Diffusion and Ionic Conduction

Vitreous silica and quartz

• Diffusion of 30Si is very slow.


• ∆HD ≈ 6 eV.
• Si-O bond energy ≈ 2.9 eV.
• (SiO4)4- bond energy = 2.9 eV × 4 bonds × 0.5* = 5.8 eV.
• Main barrier for Si4+ diffusion is the Si-O bond energy.
• For O2- diffusion, ∆HD should ≈ 3 eV.
• Experimental values lie between 0.85 – 3.08 eV.
• It appears that different diffusion mechanisms operate for Si
and O.
• Na+ diffusion in quartz parallel (llc) and perpendicular (⊥c) to
the c-axis shows a change in ∆HD at 575°C due to the phase
change from low to high quartz.
• Low quartz has trigonal symmetry, high quartz hexagonal.
• The higher symmetry reduces ∆HD.
• The diffusion coefficient in vitreous SiO2 lies in between the values for quartz.
• Na+ and Ca2+ high similar ionic radii but the diffusion coefficient of Ca2+ is much lower.
• Ca2+ has stronger bonding to the glass network.

*each O is shared by two (SiO4)4- tetrahedra.


Soda-lime Silicate Glass.

Composition of this glass is (mol %):


71.8 % SiO2, 14.52 % Na2O, 7.22 % CaO, 6.24 % MgO

• Viscosity data from the undercooled melt is


used to calculate the viscosity diffusion
coefficient Dη. (Ionic radius of Si = 0.042 nm)
• Dη follows Vogel-Fulcher Tammann behavior.
• Tracer diffusion coefficents for 22Na and 45Ca
are shown, along with the charge diffusion
coefficient Dσ.
• All three follow Arrhenius behavior.
• At Tg, Ca2+ diffusion is six orders of magnitude
lower than Na+ diffusion.

• The Ca2+ ions are more strongly bonded to the network than the Na+ ions and so cannot move
easily.
• The Na+ diffusion coefficient and Dσ have the same activation enthalpy and similar values.
• This shows that the electrical conductivity of soda-lime silicate glasses is due to Na+ diffusion.
• Diffusion coefficients for another glass are shown (composition in mol %: 71.37 % SiO2, 13.19 %
Na2O, 10.43 % CaO, 5.01 % MgO).
• Haven ratios for the two glasses are shown (assume only Na+ ions are mobile).
• Both Haven ratios are approximately temperature independent.
• This indicates that the Na+ diffusion mechanism does not change with temperature.
Alkali Borate glasses

• The structure of vitreous B2O3 is very different to that of vitreous


SiO2.
• In crystalline compounds, boron occurs in triangular and
tetrahedral coordination.
• In vitreous boric oxide, only the triangular coordination is formed.
• BO3/2 units are connected at all three corners by B-O-B bonds to
form a network.
• Some of the BO3/2 units join together in groups of three to form
boroxol rings.
• The B-O-B angle is variable and can twist as well as bend.
• The vitreous boron oxide network is planar.
• A 3-D structure is made by crumpling the planar network.
• The primary bonds exist only within the plane, with van der Waals
bonds in the 3rd dimension, so the structure is weak.
• Tg of vitreous boron oxide is only 260°C.
• Unlike silicates, addition of alkali oxides to boric
oxide does not create non-bridging oxygens.
• Addition of alkali oxide causes some trigonal
BO3/2 units to change to tetragonal BO− 4⁄2 units.
• Formation of two BO− 4⁄2 tetrahedra uses the
oxygen provided by one R2O molecule.
• Each R+ cation balances the excess negative
charge of one BO− 4⁄2 unit.
• As a result, network connectivity increases with
R2O addition.
• Thermal expansion coefficient decreases and Tg increases with R2O addition.
• Above 25 mol % R2O addition, non-bridging oxygens begin to form.

• Other structural units also exist:


Tetraborate – a boroxol ring in which one
triangle has been converted into a tetrahedron.
Diborate - a boroxol ring in which two triangles
have been converted into tetrahedrons.
• The number of these units varies with R2O
concentration.
• Variation of physical properties with
composition is complex.
• dc conductivity shows Arrhenius behavior and
increases with increasing alkali content.
• Conductivity decreases in the order Li > Na > K >
Rb.
• Conductivity is determined by the number of
mobile ions and their mobility.
• Smaller alkali ions are more mobile.
Impedance Spectroscopy of Glasses

• Ion-conducting materials: almost no electronic conduction, conductivity results from the


hopping motion of ions.
• Examples: ionic crystals, ion-conducting glasses and ion-conducting oxides.
• For ion-conducting materials, electrical conductivity measurements also give information about
diffusion.
• dc conductivity is related to the charge diffusivity by:

Dσ = charge diffusivity.
σdc = dc conductivity
𝜎𝜎𝑑𝑑𝑑𝑑 𝑘𝑘𝐵𝐵 𝑇𝑇
𝐷𝐷𝜎𝜎 = Nion = number density of mobile ions
𝑁𝑁𝑖𝑖𝑖𝑖𝑖𝑖 𝑞𝑞 2 q = electrical charge per ion
kB = Boltzmann coefficient

• Measurements of ionic conductivity are made using an ac bridge.


• This avoids polarization effects at the electrodes.
• Two identical electrodes are applied to the faces of a circular cylindrical or rectangular
parallelepiped sample.
• An electrical ac stimulus (a known voltage or
current) of frequency ν is applied to the sample.
• The response (current or voltage) of the sample is
then observed.
• A frequency range from 10-3 Hz to several MHz is
usually covered.
• A variable frequency generator, a vector ammeter
and vector voltmeter are required.
• Current, voltage and phase shift must be
measured.


𝑉𝑉(𝜈𝜈) 𝑍𝑍̂ 𝜈𝜈 = impedance
Complex impedance is defined as: ̂
𝑍𝑍(𝜈𝜈) =
̂
𝐼𝐼(𝜈𝜈) 𝑉𝑉� 𝜈𝜈 = voltage
𝐼𝐼̂ 𝜈𝜈 = current

It is composed as: 𝑍𝑍̂ 𝜈𝜈 = 𝑍𝑍 ′ − 𝑖𝑖𝑖𝑖𝑖 Z’ = real part of impedance


Z” = imaginary part of impedance
i = the imaginary unit

Complex conductivity 𝜎𝜎� 𝜈𝜈 = 𝜎𝜎 ′ 𝜈𝜈 + 𝑖𝑖𝑖𝑖𝑖(𝜈𝜈) σ’ = real part of conductivity


is composed as: σ” = imaginary part of conductivity
1 𝑑𝑑0
𝜎𝜎� and 𝑍𝑍̂ are related via: 𝜎𝜎� 𝜈𝜈 = D0 = sample thickness
𝑍𝑍� 𝐴𝐴
A = electrode area

For a circuit containing an ohmic resistance R and a


capacitance C in parallel:
𝑅𝑅
𝑍𝑍̂ =
1 + 𝑖𝑖𝜔𝜔𝜔𝜔𝜔𝜔

ω = 2πν is the angular frequency. The real and imaginary parts of the impedance are written as:

𝑅𝑅 𝜔𝜔𝜔𝜔𝑅𝑅2
𝑍𝑍 ′ = and 𝑍𝑍𝑍 =
1+𝜔𝜔2 𝑅𝑅2 𝐶𝐶 2 1+𝜔𝜔𝑅𝑅2 𝐶𝐶 2

• A plot of impedance in the complex Z’ vs. – Z” plane ω = (RC)-1


is called a Cole diagram.
̂
• For an RC circuit, the values of 𝑍𝑍(𝜔𝜔) fall on a
semicircle of diameter R. ω=0
• The semicircle should pass through the origin for ω→∞
• ω → ∞,through (R,0) for ω = 0 and through (R/2, R/2)
for ω = (RC)-1.
• For several RC circuits connected in series , the Cole
diagram is a series of semicircles.
• Three RC circuits in series can represent a polycrystalline
sample plus electrodes.
• Each RC circuit represents a conductivity process.
• RV and CV are volume conduction ω = (RbCb)-1

-
Rb and Cb are grain boundary conduction ω = (RVCV)-1
Re and Ce are electrode conduction ω = (ReCe)-1
• RC circuits are often oversimplified.
• Other equivalent circuits may better represent the
process.
• A Cole diagram for an ion-conducting alkali-borate glass is
shown.
• The semicircles represent volume conduction at different
temperatures.
• Grain boundaries are absent and the electrode conduction
is located at lower frequencies not displayed in the figure.
• dc resistance is given by the intercept of the arcs with the
Z’ axis. Resistance decreases with increasing temperature.
• dc conductivity can be deduced from the ohmic resistance
observed at the intersection of the semicircle with the Z’
axis.
• The real part of the conductivity times
temperature (σ’ × T) is plotted as a function of ν
at different temperatures.
• The low frequency plateau corresponds to dc
conductivity.
• This plateau reflects the long-range movement
of mobile cations.
• dc conductivity increases with temperature
following the Arrhenius law.
• Using the Nernst-Einstein equation, charge
diffusivity Dσ can be calculated.
• Dσ is shown with tracer diffusivity of 22Na.
• The Haven ratio HR = D*/Dσ can give information
about the atomic mechanism of diffusion and
the correlation effects involved.
• In this case, HR is temperature-independent.
• This indicates that the mechanism of ion
movement does not change with temperature.
• The increase of conductivity at higher
frequencies is called dispersion.
• This dispersion is because ionic jumps are
correlated.
• An onset frequency of dispersion, νonset, is
defined by σ’(ν) = 2σdc.
• The onset frequencies lie on a straight line with
slope = 1.
• This means that σdc × T and νonset are thermally
activated with the same activation enthalpy.
• This implies that the same jump processes take
place at different temperatures.
Electrical properties of glass

Basic theory of d.c. conductivity in solid glass

• In most oxide glasses, mobile cations such as Na+ are the


main charge carriers.
• Consider a Na+ ion in a potential well i.e. in a gap in the
continuous random network.
• To move to another gap in the network, the Na+ ion has to
squeeze past Si4+ and O2- ions i.e. overcome a potential
energy barrier.
• At T > 0K, the Na+ ion has thermal vibrational energy and has a finite and
equal probability of jumping into any of the adjacent potential wells.
• Assume that the Na+ ion moves in one dimension parallel to the x-axis.
• It jumps over a potential barrier of height ∆H/N (N = Avogadro’s
number).
• The probability of the ion moving either to the right or to the left is:

p = b exp (-∆H/NkT)

• b = vibrational frequency of the ion in its well. A Boltzmann distribution


of ion energies is assumed.
• If an electric field E is applied, the potential barrier on one side if the well is lowered and is
raised on the other side.
• This favours the motion of ions from left to right in this case.
• λ = distance between two neighbouring wells.
• The potential barrier separating each well to the right will be
reduced by 0.5 Fλ. F = eE, the force on the ion.
• The potential barrier separating each well to the left will be E
increased by 0.5 Fλ.
• The probability of motion to the right is now
∆𝐻𝐻 1
𝑃𝑃+ = 12𝑏𝑏 𝑒𝑒𝑒𝑒𝑒𝑒− − 2𝐹𝐹𝜆𝜆 /𝑘𝑘𝑘𝑘
𝑁𝑁

• The probability of motion to the left is

1 ∆𝐻𝐻 1
𝑃𝑃− = 2𝑏𝑏 𝑒𝑒𝑒𝑒𝑒𝑒−
+ 2𝐹𝐹𝜆𝜆 /𝑘𝑘𝑘𝑘
𝑁𝑁

• The Na+ ion will drift towards the right. The mean velocity of the drift motion is

𝐹𝐹𝜆𝜆
𝜐𝜐̅ = 𝑃𝑃+ − 𝑃𝑃− = 12𝑝𝑝 𝑒𝑒𝑒𝑒𝑒𝑒 𝐹𝐹𝜆𝜆
2𝑘𝑘𝑘𝑘
−𝑒𝑒𝑒𝑒𝑒𝑒
−𝐹𝐹𝜆𝜆
2𝑘𝑘𝑘𝑘 or 𝜐𝜐̅ = 𝑝𝑝 sinh
2𝑘𝑘𝑘𝑘
𝜆𝜆2 𝑝𝑝 𝐹𝐹𝐹𝐹
• For low field strength ( ½Fλ « kT), 𝜐𝜐̅ ≈ 2𝑘𝑘𝑘𝑘 𝐹𝐹. For very large field strength, 𝜐𝜐̅ ≈ 𝐶𝐶 exp 2𝑘𝑘𝑘𝑘 ,
where C is a constant.
• At room temp, Fλ is small compared to kT for field strengths of up to 10 V/cm.
• For weak electric fields, the current is proportional to field strength in accordance with
Ohm’s law.
• Current density i due to the drift of the ions of charge e in the electric field is given by
𝑖𝑖 = 𝑛𝑛𝑛𝑛𝜐𝜐̅ N = number of ions / cm3
• For moderate fields 𝑒𝑒𝑒𝑒𝜆𝜆2 𝑝𝑝𝑝𝑝 𝑒𝑒 2 𝑛𝑛𝜆𝜆2 𝑝𝑝𝑝𝑝
𝑖𝑖 = =
2𝑘𝑘𝑘𝑘 2𝑘𝑘𝑘𝑘

• For high fields 𝑒𝑒 2 𝜆𝜆2 𝑛𝑛𝑛𝑛𝑛𝑛 −Δ𝐻𝐻


𝑖𝑖 = exp
2𝑘𝑘𝑘𝑘 𝑅𝑅𝑅𝑅

2𝑘𝑘𝑘𝑘 Δ𝐻𝐻
• Resistivity ρ = E/i. ∴ 𝜌𝜌 = exp (low electric field)
𝑒𝑒 2 𝜆𝜆2 𝑛𝑛𝑛𝑛 𝑅𝑅𝑅𝑅

2𝑘𝑘𝑘𝑘 Δ𝐻𝐻
or log 𝜌𝜌 = log + (high electric field)
𝑒𝑒 2 𝜆𝜆2 𝑛𝑛𝑛𝑛 𝑅𝑅𝑅𝑅

• ∆H is usually called the activation energy for dc conductivity.


Effect of temperature on the d.c. conductivity of glass

• For measurements made near or in the molten state:

• log ρ = P + QT + RT2 + … (P, Q, R are constants).

𝐵𝐵
• At other temperatures: log 𝜌𝜌 = 𝐴𝐴 + (A, B are constants).
𝑇𝑇

Δ𝐻𝐻 2𝑘𝑘𝑘𝑘
• A and B may be interpreted as 𝐵𝐵 = − and 𝐴𝐴 = log
𝑅𝑅 𝑒𝑒 2 𝜆𝜆2 𝑛𝑛𝑛𝑛

• For alkali-silicate or alkali-lime-silicate glasses, Li-containing glasses have lower resistivity than Na-
containing glasses at equivalent concentration.
• Both have lower resistivity than K-containing glasses.
• Resistivity can vary widely with composition.

• For systems with only one alkali oxide,


increasing the alkali content tends to increase
conductivity and lowers activation energy
(provided ≥ 10 mol % alkali oxide is present).
Binary alkali silicate glasses

• Increasing Na2O content and / or temperature reduces


resistivity.
• A change in slope appears at ∼33 mol % Na2O.
• Plot of activation energy vs. Stevel’s parameter Y shows a
discontinuity at Y = 3. Y is the average number of
bridging oxygens linked to each Si. Y = 4 for SiO2 and 3 for
Na2O.2SiO2.
• Compositions with Y > 3 would have some Si atoms with
four bridging oxygens, making the network very rigid.
• However plots of log ρ vs. number of Na+ ions / cm3 do
not show a change in slope.
• The change in slope is due to experimental scatter in
values of ρ.
Mixed alkali effect

• In a binary alkali-silicate or alkali-borate glass, substituting


one alkali oxide for another e.g. Li2O for Na2O produces a
curious effect.
• Resistivity does not change linearly but goes through a large
maximum when the two alkali ions are present in approx.
equimolar amounts.
• This is called the mixed alkali effect.
• The reason for this effect is not known.
• A similar effect also happens with glass viscosity (glass
viscosity goes through a minimum point at a fixed
temperature).
• Isokom temperature (temperatures where viscosity is
constant) goes through a minimum.
• This is more pronounced at lower temperatures than in
the glass melting range.
• Addition of two alkali oxides extends the working range
of the glass.
Effect of binary or ternary additions on resistivity
• Replacement of SiO2 with an alkali oxide causes a rapid drop
in ρ.
• Replacing SiO2 with other oxides tends to increase ρ.
• This is very marked for CaO.
• BaO and PbO substitution cause ρ to increase then
decrease.
• Glass ρ is generally proportional to its viscosity.
• However, B2O3 and Al2O3 addition show the
opposite behavior.
• B2O3 addition reduces viscosity but increases ρ, whereas
Al2O3 addition increases viscosity but reduces ρ.

Binary alkali borate and germanate glasses

• For low alkali content, increasing the amount of alkali oxide


may cause a decrease in conductivity and an increase in
activation energy.
• For high alkali content the opposite behavior occurs.
• Resistivity shows a weak maximum or plateau at low alkali
concentration, then decreases steeply.
• Sodium borate and germinate glasses are more resistant than
silicate glasses of the same soda concentration.
Dielectric relaxations in glass

Static dielectric constant εs is defined as follows:

A capacitor of two parallel plates placed a small distance apart


in a vacuum with +σ and -σ charges on the plates produces an
electric field D given by:
D = 4πσ

The space between the plates is now filled with a dielectric


material. The electric field decreases to a smaller value E

4𝜎𝜎𝜎𝜎 𝐷𝐷
𝐸𝐸 = Thus = 𝜀𝜀𝑠𝑠 or 𝐷𝐷 = 𝐸𝐸𝜀𝜀𝑠𝑠
𝜀𝜀𝑠𝑠 𝐸𝐸

The drop in field strength can also be achieved by reducing surface density of charge σ by the amount:

𝜎𝜎 𝜀𝜀𝑠𝑠 − 1
𝑃𝑃 = 𝜎𝜎 − =
𝜀𝜀𝑠𝑠 𝜀𝜀𝑠𝑠

P is called the polarization of the dielectric material. D = E + 4πP


If we place the two capacitors in an alternating field, there will be a phase difference δ between D
and E (and hence between P and E). Using complex notation:

𝐸𝐸 ∗ = 𝐸𝐸0 exp 𝑖𝑖𝜔𝜔𝜔𝜔


ω = angular frequency of the applied field.

𝐷𝐷 = 𝐷𝐷0 exp 𝑖𝑖 𝜔𝜔𝜔𝜔 − 𝛿𝛿

∴ D* = ε*E* where ε* is the complex dielectric constant.

ε* = ε’ - iε”
𝐷𝐷0
Hence 𝐷𝐷∗ = exp −𝑖𝑖𝛿𝛿 = 𝜀𝜀𝑠𝑠 exp −𝑖𝑖𝛿𝛿 = 𝜀𝜀𝑠𝑠 cos 𝛿𝛿 − 𝑖𝑖 sin 𝛿𝛿
𝐸𝐸0

The real and imaginary parts of the complex dielectric constant are ε’ = εs cos δ and ε” = εs sin δ
𝜀𝜀′′
Therefore tan 𝛿𝛿 = 𝜀𝜀′

ε’ is sometimes called the dielectric constant, ε” is called the dielectric loss factor.
tan δ is called the loss angle or dissipation factor. sin δ is the power factor.
• d.c. conductivity in glasses requires the movement of
charge over long distances.
• But in an alternating field, charged particles of dipoles
which can only make limited movements can also make a
contribution to the dielectric properties.
• Charged particles oscillating between two adjacent potential minima (arrows b) cause a dielectric
loss ε” with a maximum value at a frequency ωmax = 1/τ. τ = relaxation time.
• τ is related to the height of the potential barrier separating the two minima.
• This dielectric loss acts as an ac conductivity σac = ωε”
• This ac conductivity is added to any dc contribution to the conductivity.
• Any dipoles (which can only change their orientation in response to the changing electric field)
will also make a contribution to the ac conductivity.
• The electric properties of a glass in an alternating field depend not only on mobile ions e.g. Na+
but also on relatively immobile ions e.g. Ca2+ and on dipoles which may form part of the glass
network.
• The general behavior of a glass in an alternating electric field
can be represented by a spectrum of loss (tan δ) which
consists of four mechanisms: conduction losses, ionic
relaxation losses, vibration losses and deformation losses.
• Each mechanism has its own characteristic behavior.

Conduction losses (curve 1)

• These are caused by a continuous chain of ionic jumps leading to charge transfer in the
direction of the field.
• These losses decrease rapidly with increasing frequency of the applied field and become
negligible above 50 Hz.

Ionic relaxation losses (curve 2)


• Because of the absence of long range order, the network
modifying cations are placed in different environments with
varying potential barrier heights.
• Limited cation movements cause dielectric polarization (b).
• Movements over the highest barriers (a) give continuous
conduction.
• When a steady electric field is applied, the mobile cations will try to follow it.
• This disturbs the equilibrium distribution of the cations, leading to polarization.
• This polarization is not reached instantaneously as the cations need time to move.
• For the simplest model, the heights of the b-type potential barriers are all equal.
• The frequency dependence of the complex dielectric constant in an alternating field is given by
the Debye equation:
ε = complex dielectric constant
𝜀𝜀𝑠𝑠 − 𝜀𝜀∞ εs = static dielectric constant
𝜀𝜀 = 𝜀𝜀∞ + ε∞ = dielectric constant at very high frequency
1 − 𝑖𝑖𝑖𝑖𝑖𝑖
τ = relaxation time
• Writing ε = ε’ - iε” and separating the real and imaginary parts gives:

𝜀𝜀𝑠𝑠 − 𝜀𝜀∞ 𝜀𝜀𝑠𝑠 − 𝜀𝜀∞ 𝜔𝜔𝜔𝜔


𝜀𝜀 ′ = 𝜀𝜀∞ + and 𝜀𝜀 " =
1 + 𝜔𝜔 2 𝜏𝜏 2 1 + 𝜔𝜔 2 𝜏𝜏 2

• ε” gives the power loss in the glass. It has a peak at ω = 1/τ.


• The potential barriers are partly electrical (repulsion of ions) and partly geometrical in
nature.
• When an electric field is applied, the mobile cations try to move in the same direction.
• The electric field alone is not enough to carry the ions over the potential barriers.
• The ions have to gather extra energy from collisions with other ions due to thermal
motion.
• The time needed for the ion to gather enough energy to overcome the potential barrier is
called the relaxation time τ.
• If the electric field is alternating with angular frequency ω, and ω is not large compared to
1/τ, then during each half-cycle the mobile cations have enough time to jump over the
potential barriers.
• They absorb energy from the electric field, which is transmitted to the network as heat.
• At ω « 1/τ, the losses are small. The ions jump over the barrier when the value of the
electric field is close to zero. The ions absorb little energy from the field.
• When ω ≈ 1/τ, then the ions cannot jump over the barrier until the field has reached its
maximum value. The ions absorb much energy and the losses are large.
• When ω » 1/τ, the ions jump independently of the phase of the field and the losses are
small.

𝜔𝜔𝜔𝜔
tan 𝛿𝛿 = Losses occur only at ω ≈ 1/τ
1 + 𝜔𝜔 2 𝜏𝜏 2
• In a real glass, the local structure around different cations will also be different, leading to a
range of potential barrier heights.
• Hence there will be several relaxation times and several relaxation loss peaks at different
frequencies.
• A distribution function Y(τ)dτ in introduced.
• This gives the contribution to εs from ions having relaxation times lying between τ and (τ + dτ).
• The Debye equation is changed to

𝑌𝑌 𝜏𝜏 𝑑𝑑𝑑𝑑
𝜀𝜀 = 𝜀𝜀∞ + �
1 − 𝑖𝑖𝑖𝑖𝑖𝑖
0

• To analyze the loses, the knowledge of the function Y(τ) is needed.


• The losses from ionic relaxation are small at ω « 1/τm and at ω » 1/τm.
• Losses have a maximum near ω = 1/τm.
• τm is the most probable value of the relaxation time.
Vibration losses (curve 4)

• All ions in the glass vibrate around their equilibrium positions.


• Vibration frequency is determined by their mass and the restoring force in the potential well.
• If the applied electric field has a frequency close to that of one of the ions, it will vibrate at
high amplitude.
• The vibrating ions exchange energy with their surrounding ions, causing damped motion.
• Energy proportional to the square of the amplitude is absorbed from the applied electric
field.
• This causes high dielectric losses (resonance absorption)
• There are different types of ion in the glass, and their surroundings vary.
• The resonance absorption curve is not sharp but is spread about the most probable
frequency.
• For most glasses, this frequency is in the infra-red region.
• The wide spread of the loss curve causes losses in the microwave region.

𝛼𝛼 ωres = resonant frequency


𝜔𝜔𝑟𝑟𝑟𝑟𝑟𝑟 = α = proportionality factor depending on the restoring force
𝑀𝑀
M = mass of the ion
Deformation losses (curve 3)

• At low temperatures (50 K), an extra loss peak can appear at ∼1 MHz
• This is caused by tiny displacements of the ions.
• This could be a small bond rotation or change in bond angle.
• Whole sections of the network move simultaneously.
• Activation energies and relaxation times are small.
• Loss maxima are in the region of 1013-1014 Hz at room temp.
• These losses are hidden by the vibration losses at room temperature.

You might also like