Download as pdf or txt
Download as pdf or txt
You are on page 1of 102

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/319471996

CHEMICAL STRENGTHENING OF GLASS BY ION-EXCHANGE

Working Paper · October 2017

CITATION READS

1 9,478

1 author:

Guglielmo Macrelli
Isoclima Spa
44 PUBLICATIONS   173 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

HaSU - Development of Hydrophobic Windscreen Coating for Next Generation Civil Tilt Rotor View project

Diffusion studies in glasses related to residual stress build up for application in chemically strengthened glasses by ion exchange View project

All content following this page was uploaded by Guglielmo Macrelli on 18 January 2018.

The user has requested enhancement of the downloaded file.


CHEMICAL STRENGTHENING OF GLASS BY
ION-EXCHANGE
Guglielmo Macrelli
Lecture Notes

Workshop at the 2017 ICG Annual Meeting


Istanbul – Turkey October 24th, 2017

Copyright © 2017 Guglielmo Macrelli – All rights reserved

This booklet/lecture notes is not for sale and it is not for mass dissemination but just for
students and colleagues

1
GLASS STRENGTHENING BY ION EXCHANGE

Index :
Forward
1. Introduction
1.1 Glass strength: the surface flaws problem
1.2 Glass strength: fatigue as thermally activated stress assisted crack growth
1.3 Glass strengthening techniques
1.3.1 Strengthening by flaws removal/reduction
1.3.2 Strengthening by flaws healing
1.3.3 Strengthening by introduction of residual stress
1.3.3.1 Thermal strengthening
1.3.3.2 Chemical strengthening

2. Chemical strengthening by ion exchange


2.1 Ion exchange theory
2.2 Kinetics of Ion Exchange: residual concentration of invading ions
2.3 Stress build-up and relaxation
2.4 Strength issues of ion exchanged glass
2.5 LNDC anomaly and zero time surface compression
2.6 Characterization of ion exchanged glass
2.6.1 Strength measurements by flexural tests
2.6.2 Statistical methods for glass strength

3. Product and process control


3.1 Chemical depth profiling
3.2 Strength measurements
3.3 Stress depth profiling
3.4 Process design and process control
3.5 Interdiffusion coefficient definition
3.6 Reference vlues
3.7 Process control procedures

Bibliography

Appendices
A1 – Ion exchange from a deposited finite layer and post ion exchange thermal
treatment
A2 – Relationship between the exchanged molar flux and the process parameters
A3 - Terminology
A4 – Evaluation of the depth of the compression layer using the “thermal stress
analogy” equation.
A5 – Non Isothermal Ion-Exchange
A6 – Viscosity
A7 – Thermodynamics of Ion Exchange
A8 – Presentation of Glass chemical composition formulae
A9 – Molar volume determination

2
Forward
This booklet has been conceived as a lecture guide to a Workshop organized within the 2017
ICG annual meeting to be held in Istanbul (Turkey) on the subject of “Glass Chemical
Strengthening by Ion Exchange”.

The well knows brittleness of glass is in deep contrast with the structural use of this material
in a wide range of applications: architectural, transportation, pharmaceuticals and thin
strengthened covers for consumer electronics. Among the different technologies to improve
glass strength, chemical strengthening by ion exchange is increasing its range of applications
and the overall quantity of strengthened glass articles by Ion Exchange is becoming
significantly wider.

The starting point is glass strength. Without having in mind the basic concepts about glass
strength it is very unlikely that one may achieve a good understanding of glass chemical
strengthening.

One of the first evidence in testing glass strength is that it breaks suddenly without any
evidence of yield. Drawing a plot of the stress (σ) versus strain (ε), determined during a glass
strength test, we get just a straight line up to the fracture event (see figure below).

Glass strength is basically a problem related to glass surface and its interactions with the
environment. There are two basic concepts to understand:

- First lesson (Inglis and Griffith): surface flaws are responsible for brittle fracture and
for the low values of strength when compared with the theoretical values predicted by
considering the Si-O bond strength.

But this is not enough,

- Second lesson (Widerhorn, Mikalske-Freimann): surface flaws may interacts with the
environment leading to subcritical crack growth which is responsible for delayed
fracture evidences (fatigue effects in glass).

3
Strengthening methods and techniques have been developed to overcame these two aspects of
glass strength.
Technologists sometime use processes and methods without knowing the scientific
background (at least in its concepts) and this is a good way to make serious mistakes.

The purpose of the lecture to be held at the above mentioned workshop is to present in a
rational way the fundamentals of material science (in this case glass science) that generates
methods and technologies for glass strengthening by ion exchange.
The purpose is always the same: get higher strength and reliability of final products when
exposed to structural loads and service life conditions.

The booklet is organized with an initial introduction that is, basically, a glass science
refreshment and discussion about glass strength concepts covering brittle fracture and
subcritical crack growth effects leading to fatigue issues. The following main part of the
booklet is related to glass chemical strengthening by ion exchange and it is divided in five
sections:

Glass Strengthening by the introduction of residual stress

Residual concentration of invading ions in the glass during ion exchange

Stress build up and relaxation

Ion Exchange process and product control

The content of the booklet is not “easy” or “too much simplified”. Complexities have not
been hidden either from the mathematical point of view or from the physical-chemical
approach. The reader is supposed to have a background at a undergraduate / graduate level in
material science or equivalent scientific area (physics, chemistry, engineering). This is also in
the spirit of the ICG annual meeting to encourage post-docs and students at national and
international level to scientifically approach the matter and to go deep into that.

One of the important parts of this booklet is the bibliography. I have divided this into two
different parts: books and articles. The reader is strongly encouraged to dig into this huge
amount of scientific literature to deepen concepts and knowledge: good or seminal ideas are
often due to the effort to understand and eventually critically review what somebody else has
produced before.

A full presentation of about 140 slides is part of the workshop documentation. This
booklet/lecture notes has to be considered a complementary support to that.

Finally, any responsibility for typos, mistakes or any other oddities you may find here is
obviously due to myself.

Comments, due corrections, clarification requests may be addressed to my e-mail


guglielmomacrelli@hotmail.com.

4
1. Introduction
For the sake of this booklet “glass strength” will be assumed the stress value at breakage
indicated as “σb” and expressed in Pa (N/m2) units or by its natural multiplies : kPa, MPa or
GPa.
When dealing with glass strength the first striking evidence is the huge range (3-4 order of
magnitude !) of strength values. As shown in Figure I-1 this wide range of strength values is
roughly depending on a number of factors either related to the glass conditions or related to
environmental elements.

Figure I-1 – Glass strength figures

In order to get a rationale of these apparently uncorrelated evidences let us focus on the first
source of discrepancy between what can be defined “Theoretical glass strength” and
“Effective glass strength”. I am speaking of the surface flaw issue.

The main features of glass strength and glass breakage are related to:

(i) Brittle failure that accounts for failure far below the theoretical strength without
any apparent yield effect on macro scale
(ii) Time delayed failure effects at both static and dynamic loads

In paragraph 1.1 a general discussion is presented of the brittle failure problem introducing
the surface flaw model while in paragraph 1.2 it will be discussed the time delayed failure
effects using the same flaw model and introducing the flaw growth velocity as thermally
activated stress assisted crack growth effect.
Figure I-2 (inspired by a Corning® brochure of Gorilla® glass) provides a visual
representation of the glass surface flaws issue.

5
Figure I-2 – Glass suface flaws

I will present glass strength in terms of geometrical characteristics of surface flaws: an


adimensional factor Y and its critical depth ac, and material/environment characteristics
included in the so called Critical stress intensity factor KIC ( see Figure I-3). This glass
strength introduction starts with initial arguments developed in the first part of last century by
Inglis and Griffith than rationalized in linear fracture mechanics concepts (In the text and
figures of this booklet the flaw depth is normally indicated by the letter “a”. Sometime you
will find it indicated by the letter “c”. This has not been amended as I think the context is
quite obvious).

Figure I 3 Glass strength in terms of geometric characteristics of surface flaws

Today it is possible, using atomic force microscopy and nano-indentation techniques (See
Figure I-4 a), to have access to a direct characterization of surface flaws.

6
a) b)
− 11 −1
Figure I-4: a) In situ view at the tip of a propagating crack (≈ 10 m s ) in silica glass as
obtained by atomic force microscopy at relative humidity ≈ 60%, 22 ° C. (From L.
Wondraczek , A. Dittmar , C. Oelgardt , F. Célarié , M. Ciccotti C. Marlière [62])
b) Representation of an applied external stress σa and a residual stress σR on a surface flaw.

nevertheless, a complete surface scan of large area glass products, in order to detect flaws at a
micron size, is technologically nearly impossible by inspection methods or devices currently
available at production sites. Therefore, indirect control of surface flaws population is a
matter of quality control in glass selection and production procedures. Generally speaking a
high level of original glass strength is a necessary baseline for any subsequent strengthening
method but, for strengthening by ion exchange, is particularly critical because in this case the
depth of the introduced compression layer is of the same order of magnitude of surface flaws.

1.1 Glass strength: the surface flaws problem

Theoretical strength of solid materials can be expressed in terms of elastic modulus and
surface energy according to (Orowan 1934):

E ⋅γ f
σ Th =
ro (1.1)

Where:
E – Elastic (Young) modulus (MPa)
γf – Surface energy (J/m2)
ro – Equilibrium distance between the atomic component of the material (m)

Considering for soda-lime glass typical values:

7
E = 70000 MPa
γf = 3.5 (J/m2)
ro – 0.2 (nm)

it results a theoretical strength value:

σTh = 35000 MPa .

This value is of orders of magnitude far above from the experimental values for soda-lime
glass, that are around 50-100 MPa.
This great difference between theoretical and experimental values of glass strength is related
to the presence of flaws on the glass surface. Flaws act as stress amplifiers so that the applied
tensile stress is amplified at the flaw tip where it reaches the value of the theoretical strength
leading to glass breakage. A simple argument due to Inglis can explain how this is possible.
Consider an elliptical flaw of axis 2b and 2c under the action of an applied stress σa .
According to Inglis [1913] (see Figure 1) the stress at the tip is:

c
σ yy = σ a ⋅ (1 + 2 ⋅ ) (1.2)
b

The radius of curvature at the tip of the flaw is:

b2
ρ= (1.3)
c

And the stress equation results (in terms of radius of curvature):

c
σ yy = σ a ⋅ (1 + 2 ⋅ ) (1.4)
ρ

Figure 1 – Elliptical flaw in a glass plate under applied tensile stress σa

8
In the case of a very long elliptical flaw (sharp flaw) c >> ρ :

c
σ yy = 2 ⋅ σ a ⋅ (1.5)
ρ

In the assumed hypothesis of very sharp flaws we can reasonably conclude that the radius of
curvature ρ at the flaw tip is of the same order of magnitude of the Si-O bond (r0). This allows
to define so breakage condition from σyy = σTh as:

E ⋅γ f c
= 2 ⋅σ b ⋅ ;
r0 r0

1 E ⋅γ f E ⋅γ f
σb = ⋅ = (1.6)
2 c 4⋅c

Where σb is the breakage stress. Considering a flaw size of 10 µm it results:

σ b = 78MPa

A 50 µm more severe flaw will result a final strength of:

σ b = 35MPa

Both results are quite comparable with the experimental values.

Another criterion for strength can be derived using a thermodynamic argument (Griffith 1921)
considering that the flaw must reach a critical length at which it can propagate freely. This
criterion considers the total energy U of a plate with an elliptical flaw of length 2c. In the
elastic domain the elastic strain energy per unit volume (J/m3) is (see the figure in the
introduction) just the area below the stress/strain curve which is dictated by the Hooke law (ε
is the strain):

σ = E ⋅ε (1.7)

and it is 1/2 σ· ε. Considering an infinitely narrow elliptical flaw (b→0) this strain energy is
distributed over a surface around the flaw, and can be expressed as elastic strain energy per
unit length c of the flaw introducing a suitable geometric factor Y. Finally the elastic strain
energy per unit length of the crack front (J/m) (actually two crack fronts) can be defined:

σ2
U el = −σ ⋅ ε ⋅ Y 2 ⋅ c 2 = − ⋅Y 2 ⋅ c2 (1.8)
E

In case of a flaw of length 2c that grows due to this applied stress, this elastic strain energy
per unit length is reduced to create 2 new flaw surfaces that require an increasing of surface
energy per unit crack length (J/m):

9
Us = 4 ⋅γ f ⋅c (1.9)

Considering the applied stress (σ = σa), the total elastic energy per unit flaw length will result:

Y 2 ⋅ c 2 ⋅ σ a2
U =− + 4⋅c ⋅γ f (1.10)
E

This means that an applied tensile stress generates an elastic strain energy, if this energy is
enough to generate new surfaces the crack will propagate spontaneously otherwise the crack
will stay at his initial length.
The first term in equation (1.10) is the loss of strain energy used to create new surfaces
represented by a gain in surface energy (second term of (1.10). For small values of c the
second term dominates and U is positive, but as c become larger the first term dominates and
U reduces continuously as c increases. In these conditions the flaw become energetically
unstable and grows spontaneously. The critical value of the flaw length, c*, upon which it
grows spontaneously is found in correspondence at the U maximum and can be easily found
differentiating U in respect to c and setting the derivative to zero:

1 2 ⋅ E ⋅γ f
σb = ⋅ (1.11)
Y c*

With this last criterion considering (as an example) the geometric factor Y=(π)1/2 and a flaw
length of 10 µm it results for soda-lime glass:

σ b = 124MPa

A 50 µm more severe flaw will result a final strength of:

σ b = 56 MPa

This last value is again quite comparable with the values found experimentally. This analysis
indicates that glass strength is strongly related to the state of the surface by the presence of
flaws in the form of micro-fissures. These flaws extend generally from few up to tens of
micron into the glass. Few of them may be even more severe resulting in a strongly reduced
mechanical strength (see Figure 2).

10
Figure 2 – Glass Strength VS Flaw depth (From T.Rouxel, S.Yoshida [91])

Flaws may originate from a number of sources. The first source is the glass article
manufacturing process itself: contact with handling equipment and atmospheric particles and
dust may create surface flaws in glass. Subsequent processing as cutting, grinding, laminating,
packaging may induce even more severe flaws in the glass articles. Final use of glass is also a
source of flaws on glass surfaces. The result is a glass article with a wide variety (surface
population) of flaw lengths that corresponds to a wide variation of breaking strengths.
Following this argument it easily comes out that glass strength is a statistical concept strongly
related to the statistics of surface flaws.

Figure 3 – Schematic representation of a surface flaw under tensile stress

From equations (1.6) and (1.11) it can be realized that the breakage condition can be
expressed as:

σ b ⋅Y ⋅ c = 2 ⋅ E ⋅γ f (1.12)

11
Where E is a material characteristics and γf is determined by the material and its environment.
This equation can be arranged in terms of fracture mechanics concepts introducing the stress
intensity factor (SIF) KI. This factor is related to the applied stress and to the flaw
characteristics (geometrical factor Y and depth c) and it is basically the first member of
equation (1.12). Given a defined surface flaw characteristic than the SIF is proportional to the
applied stress σa according to:

KI = σ a ⋅Y ⋅ c (1.13)

This equation can be transformed in the breakage condition equation defining a critical stress
intensity factor KIC

K IC = 2 ⋅ E ⋅ γ f (1.14)

Fracture condition occurs as KI→ KIC. This condition is achieved by increasing σa or


increasing c. Units for KI are accordingly [MPa·m1/2]. From these definitions it is clear that
KIC is a material property slightly influenced by the environment because of γf. In table 1
some typical values are reported for Young modulus (E) and the critical stress intensity factor
(KIC) for glasses of different chemical compositions.

Glass E (MPa) KIC (MPa·m1/2)


Fused Silica 73000 0.74 ÷ 0.81
Soda-lime silica 70000 ÷ 74000 0.72 ÷ 0.82
Aluminosilicate 83000 ÷ 91000 0.85 ÷ 0.96
Borosilicate 64000 ÷ 89000 0.75 ÷ 0.82
Lead-silicate 58000 ÷ 65000 0.62 ÷ 0.73
Table 1 – Young modulus and stress intensity factor for different glasses

In fracture mechanics there are three critical values of the stress intensity factor depending on
the mode of crack extension (see Figure 4).
opening in tensile (KI) where failure plane and propagation is perpendicular to
principal tension,
sliding in shear failure (KII) where the applied stress acts to slide fracture planes over
each other along the direction of propagation,
tearing in shear (KIII) where direction of propagation is perpendicular to that of the
stress but failure plane is parallel to shear stress direction.
For glass strength and, in general for brittle solids, only the first condition (KI) is relevant.

a) b) c)
Figure 4 –Modes of crack extension : a) KI, b) KII and c) KIII

12
The elastic strain energy per unit crack length Uel is sometime indicated in the literature by
the letter G and it is equal to K2/E. So G is the change in the strain energy of the body for unit
change in crack length. In our case using the Griffith criterion (1.11) (Y = π):

2
K IC σ 2 ⋅ Y 2 ⋅ c*
G IC = = b (1.15)
E E

And :

1 G IC ⋅ E
σb = ⋅ (1.16)
Y c*

That results in the definition for GIC:

G IC = 2 ⋅ γ f (1.17)

Toughness of a solid is its capacity to absorb strain energy before failure. Considering a
stress/strain curve, toughness is the area under this curve. For a brittle solid this point is of
little importance in the macro scale. It will be more significant to consider the energy required
to extend a flaw from its initial length to the critical one. So GIC can be considered as a
definition of fracture toughness. Some authors call the critical stress intensity factor KIC =
(E·GIC)1/2 “fracture toughness of the material”.

Glass Environment γf (J/m2)


Soda - lime N2 (l) @ 77K 4.1
Air @ 22°C , 40% RH 3.5
Vacuum @ 0.1 torr 5.0
Fused Silica N2 (l) @ 77K 4.6
Air @ 22°C, 40% RH 3.7
Aluminosilicate N2 (l) @77 K 5.2
Air @ 22°C , 40% RH 3.7
Borosilicate N2(l) @ 77K 4.7
Air @ 22°C , 40% RH 4.0
Table 2 – Fracture surface energy for different glasses in different environments

1.2 Glass strength: fatigue as thermally activated stress assisted crack growth

The theory outlined above provides a good explanation for the huge difference between
theoretical and experimental values of glass strength. But there also some other evidences in
glass strength that need to be carefully considered:

Glass breakage may suddenly occur at constant load when leaving the load acting for
long time over the glass article. The time-to-failure is reduced as the load in increased.
In a strength test higher stress to failure (breakage stress) is achieved increasing the
stressing rate.

These two new elements introduce the “time” variable as a relevant parameter in glass
strength. Both observations indicate evidence of delayed failure as typical of glass strength.

13
These effects are normally referred as “fatigue” effects: some authors name the first “static
fatigue” and the second one “dynamic fatigue”.
Regardless of definitions, if we remain in the flaw model for the explanation of glass strength
we should conclude that, to explain the delayed failure mechanism, we should consider that in
some way flaws may grow with time as they are under applied tensile stress condition and the
flaws growth velocity has to be in some way proportional to the applied stress. This means
that at lower stress, flaws grow velocity is reduced allowing a longer time delayed failure.
In this framework the so called “static” and “dynamic” fatigue effects can be interpreted as
“sub-critical crack growth” effect which is function of the applied stress (or better of the stress
intensity factor) (Wiederhorn et al. 1967, 1970). Experiments made to evaluate the growth
velocity of cracks as function of stress intensity factor Vc(KI) put in evidence the following
features:

There is a strong dependence of Vc(KI) curve on environment conditions mainly


humidity (Water vapour concentration).
Four different regions of different subcritical crack growth behaviour were observed
after which the crack has grown to the critical size required to fulfil the Griffith energy
criterion for spontaneous failure

Region 0 – very low crack velocities where the Vc vs. KI plot become almost vertical. This
indicates the evidence of a static fatigue limit - threshold intensity factor (KIth) below which
crack do not grow, for soda lime glass KIth = 0.25 MPa·m1/2.
Region I – Crack velocity increases exponentially with applied load and with humidity values.
Region II – where crack velocity is independent from applied stress but still increasing with
humidity level
Region III – Where crack velocity does not depend on humidity level but depends once again
exponentially from the applied load.

The slope in region III is much steeper than in region I, at the end of region III KI→KIC and
spontaneous failure happens. The so called spontaneous failure is considered when velocity
exceeds 0.1 m/s. This means that velocities below 0.1 m/s are the sub-critical crack velocities
causing the crack growth up to its critical value.

Crack velocity depends upon glass composition and, in region I, also on temperature and
environment conditions. This suggests that sub-critical crack growth may be a thermally
activated stress corrosion process. The behaviour in region II can be explained considering
that crack velocity is limited by the rate of water vapour transport to the crack tip.
Dependence of crack velocity to stress intensity factor is normally expressed with a power
law:

n
dc  K 
VC = ; VC = A ⋅  I  (1.18)
dt  K IC 

A and n are constants. n is named “stress-corrosion susceptibility coefficient” or “crack


growth exponent” and ranges between 12 and 35 for most glasses.

14
Figure 5 – Crack velocity Vc vs. Stress intensity factor KI

Crack extension is, in this way, connected to a stress corrosion mechanism that involves the
reaction of water molecules with silica at the flaw tip. The reaction is understood as the
transformation of a siloxane bond Si-O-Si which is the basic bond of glass structure into two
silanol groups (Si-OH) according to :

Si-O-Si + H2O → Si-OH + HO-Si

This reaction is a glass hydrolysis reaction that leads to flaw length growing; the crack tip
hydrolysis in enhanced by stress concentration.

Figure 6 – Flaw growth after glass crack tip hydrolysis enhanced by stress concentration.

By coupling equation (1.13) and equation (1.18) it is possible to calculate the time to fracture
for a crack of extension Co under a tensile applied stress of σa up to the critical value Cf as:

tf Cf n K IC
dc K IC dc
t f = ∫ dt = ∫v A ⋅
= ∫ (1.19)
0 C0 KI K In

Which results:

15
2 ⋅ K IC
n
 1 1 
tf = ⋅  n − 2 − n −2 
A ⋅ Y ⋅ σ a ⋅ (n − 2 )  K I
2 2 (1.20)
K IC 

Expression (1.20) can be further reduced as a function of crack extension c:

2 ⋅ K IC
n
2 ⋅ K IC2
tf = −
A ⋅ (Y ⋅ σ a ) ⋅ (n − 2 ) ⋅ c n / 2−1
(1.21)
n
A ⋅ Y 2 ⋅ σ a2 ⋅ (n − 2)

As an example consider a flaw of extension c=100 µm . Using equation (1.13) with a


geometric factor Y=(π)1/2 we get a fracture stress of

σ b = 42 MPa

If the same crack is exposed to a permanent tensile stress of 12 MPa, well below the breakage
stress, the time to failure will result:

t f = 82hours

This means that, even if the applied tensile stress is well below the breakage value, providing
that it is a permanent stress, the glass will break after about three days and a half . This is true
if the flaw length is such that, under the applied stress σa, the stress intensity factor (KI) is
larger than the threshold limit KIth (because this is the condition to activate the subcritical
crack growth). See below examples of threshold externally applied load leading to subcritical
crack growth ( KIth = 0.25 MPa·m1/2)

A) c=100 µm → σth = 14 MPa


B) c=50 µm → σth = 20 MPa
C) c=30 µm → σth = 25 MPa
D) c=20 µm → σth = 32 MPa
E) c=10 µm → σth = 44 MPa

The first conclusion is that good surface quality (lower crack/flaws depth) increase the limit
of applied external stress that trigger the subcritical crack growth mechanism.

This result is a good example of how the subcritical crack growth is of paramount importance
in glass strength and has to be carefully considered in structural design and applications.

1.3 Glass strengthening techniques

As above discussed, glass strength is a problem strongly related to surface conditions. Lower
values than expected are explained by the presence of surface flaws while time delayed failure
effects are explained by sub-critical crack growth produced by stress corrosion at flaws tip
enhanced by environment conditions (temperature and water vapour) and tensile stress
concentration. Anything (metals, ceramic, other glass, dust particles and even human hands)
coming in contact with glass surfaces (both in processing and end using) may cause surface
damage lowering its strength. The need to improve strength values and reliability in respect to
static and dynamic fatigue is of great practical importance. Glass breakage is in most cases in

16
correspondence of a tensile stress, breakage in compression stress is generally possible but
practically very unlikely.
Strategies commonly considered for glass strengthening under the action of a service/desgn
load are related to:

Increase of practical strength:

- Eliminate flaws or at least reduce their severity (reducing KI under a service load σa)
- Introduce residual compression stress in the near surfaces areas
- Healing flaws by suitable processing techniques.

and

Increase of service life:

- Introduction of surface coatings to prevent surface cracking (flaws formation during


service life)
- Introduction of surface protection to prevent crack stress corrosion mechanism.

In this booklet we will consider strengthening processes based on the introduction of near
surface residual compression stress by chemical ion-exchange processes. Nevertheless a short
description of other strengthening methods will be given in the following articles.

1.3.1 Strengthening by flaws removal/reduction

As discussed at point 1.1 the critical factor in flaws severity is their sharpness represented (see
equation 1.5) by the ratio c/ρ. Fire polishing during glass formation is a good technique to
achieve this because glass flowing rounds off or close the flaws decreasing c and/or
increasing the radius of curvature (ρ). In a similar way a low concentration (2-5 %) of
hydrofluoric acid (HF) can dissolve glass (etching) at a few microns per minute rate. This
action reduce flaws severity or eliminates flaws. Also glassware autoclaving under a
pressurized steam or water can reduce the flawed surface. These techniques are effective only
for strengthening glass articles with unexposed glass surfaces. Subsequent handling or
environment exposure in fact will inevitably introduce new flaws on exterior surfaces.
Very high levels of glass strength have been achieved by HF etching in the order of 1000
MPa, unfortunately if the glass surface is not immediately protected this high strength is not
retained for a long time.

1.3.2 Strengthening by flaws healing

Another method based on the concept to neutralize flaws action is not their removal but their
“filling” with a suitable material in order to reduce the stress concentration effect. This
strategy is performed by coatings that have the purpose to re-create a chemically bonded
material filling of the crack. Several recipes have been proposed using both organic, inorganic
and organic/inorganic coating. Strength values around 200-250 MPa are possible by this
technique. Compatibility and protection of surface coating are issues always to be considered
when using surface coating.

17
1.3.3 Strengthening by the introduction of residual compression stress

In this strengthening method a residual compression stress field is introduced in the glass
body by some technological process (thermal or chemical). The idea (see Figure 7) is to
introduce compression stress in the near surface areas of the glass body to pinch the surface
flaws preventing their growing under externally applied loads. For mechanical stability these
induced compression stress in the glass near surface shall be balanced by inner tensile stress
in the glass body.

Figure 7 – Residual compression stress field compared to a glass surface flaw (Flaw depth is
here indicated by “a” instead of “c”)

Let’s rewrite the expression (1.13) of the stress intensity factor under the action of an applied
service stress σa in correspondence of a flaw of length c:

KI = σ a ⋅Y ⋅ c

This equation is valid in the hypothesis that there is no residual stress. The introduction of a
residual stress field:

σ R = σ R (x) (1.22)

modifies the stress intensity equation introducing a stress intensity factor due to the residual
stress: KR. It is important to recall a mechanics convention that compression stress is negative
while tensile stress is positive. The stress intensity factor due to the introduced residual stress
can be evaluated from the residual stress field σR(x) using the Green function method:

2 ⋅Y ⋅ c
c
σ R ( x)
KR = ⋅∫ dx (1.23)
π 0 c2 − x2

This contribution can be introduced in the expression of the stress intensity factor:

KI = σ a ⋅Y ⋅ c + KR (1.24)

18
Equation (1.24) at failure reads:

K IC = σ b ⋅ Y ⋅ c + K R (1.25)
For compressive residual stress KR < 0 and:

K IC − K R K IC + K R
σb = = (1.26)
Y⋅ c Y⋅ c

This means, according to what intuitively obvious, that, for compressive residual stresses,
fracture happens under a higher applied load than for residual stress free state.

In next paragraph both thermal and chemical strengthening will be considered. In next
sections a more detailed evaluation will be carried out for chemically strengthened glass that
is the main subject of these lecture notes.

1.3.3.1 Thermal strengthening

Glasses can be strengthened significantly by a thermal process referred as “Thermal


Tempering” in which residual compressive stress is introduced in the glass surfaces. The
thermal process involves heating of the glass body to a temperature above its glass
transformation range (basically above its glass transition temperature) followed by a quick
quenching down to room temperature. The residual stress field is introduced in the quenching
phase of the process while the glass article pass through the transformation range of
viscosities.

A simplified description can be understood as follows:

The quenching step can be considered in two stages, in the first stage the surface layers
contracts more rapidly than the interior and becomes rigid while the interior remain still in its
viscous state. This could result in tensile stress at the surface that relaxes very quickly since
the interior body of the glass is viscous. In the second stage the interior glass body is cooled
down to room temperature. The glass interior will cool at a slower rate than the exterior. This
means that its final specific volume will be smaller than the exterior. (It is like that the full
contraction of the interior body is prevented by the already achieved full contraction of the
surfaces). This effect results in a tensile state of the interior balanced by a compression state
of the external surfaces.

This process introduces a parabolic shaped residual stress in the glass cross section with a
typical surface compression ranging from 80 MPa to 100 MPa and a depth of the compression
layer CD of 1/5 of the glass thickness.

19
Figure 8 – Residual stress distribution in a thermally tempered glass
The parabolic stress distribution is strictly related to the temperature distribution achieved
during quenching. The limiting factor to achieve higher stress are:

- Glass thermal expansion coefficient α. When this value is reduced (e.g. Borosilicate glass),
the capability to introduce compressive stresses by rapid quenching is reduced accordingly
- Glass thickness t. By reducing glass thickness it becomes difficult to create a significant
temperature gradient. Accordingly it results more difficult to introduce compressive stress.
Practical limit in thickness for thermal tempering in air is around 3mm
- External heat transfer coefficient – h(W/m2K) during quenching. The higher is this
coefficient the more it becomes possible to temper glasses with low thermal expansion
coefficient and thin glasses.
Special thermal tempering processes has been developed to increase the heat transfer
efficiency by means of liquid or phase transformation quenching and processes where glass is
moved over air cushions. These last improvements allows the capabilities to thermally temper
glass with low thermal expansion coefficient and also set the “temperable thickness limit” to
around 2 mm.
In mathematical terms the residual stress field σ (x) can be expressed as a function of the
thermal expansion coefficient α, the heat transfer coefficient between the impinging air jets
and the glass surface h, and the glass thickness t :

σ ( x ) = σ (α , t , h) (1.27)

The case depth of 1/5 of glass thickness creates a good condition in respect to protection to
service life cracks and to sub critical crack growth. Indicating c as typical flaw depth and CD
the case depth of thermally strengthened glass:

C D >> c ; C D = 0.2 ⋅ t ; t = glass thickness (1.28)

In this condition we can assume that stress field is almost constant over all the flaw length and
its value is just the surface compression. In other words, because the compression layer depth
is of the order of millimeters and flaws are on the order of tens of microns, it is quite
reasonable to assume that the residual compression is nearly constant over the flaw depth. In
this approximation we can write equation (1.23) as:

K R = Y ⋅ c ⋅ SC ; SC < 0 (1.29)

So that equation (1.26) becomes :

20
K IC
σb = + SC = σ 0 + SC (1.30)
Y⋅ c

Where σ0 is the original glass strength before thermal strengthening.

Therefore, in thermal tempering, glass strengthening is directly related to the amount of


surface compression since the first term in equation (1.30) is just the original glass strength
before strengthening.

As may be clear from figure 8, the amount of elastic energy stored in the tensile inner zone of
the glass is significant. This means that, upon breakage, due to the sudden release of the
stored tensile energy, the glass article will dice into a large number of small fragments. These
small fragments are dull-edged and pose limited risk of injury to people (accidental cuts). The
number of pieces and their size is determined by the degree of temper and by the thickness,
the higher is the inner tensile stress the larger the number and smaller will be the size of the
fragments (level of surface compression SC that is related, in thermal tempering, to the level
of inner tension τ(t/2) = -2·SC )
Thermal tempering can be discussed at a deeper level but this is not the purpose of this
booklet.

1.3.3.2 Chemical Strengthening

By chemical strengthening we identify those processes where the glass strength is increased
by the chemical modification of the near surface layer that results in the introduction of
residual compressive stress. This is normally achieved through ion exchange chemical
reactions, where some modifiers ions (normally alkali ions or alkaline-earth ions) in the glass
matrix are exchanged for other ions delivered to the glass surface by a suitable mean. We
identify basically two processes:

Ion exchange below the glass transition temperature


Ion exchange above the glass transition temperature

The first process that we will call shortly “Ion-Exchange” is based on the exchange of smaller
alkali ions in the glass matrix ( Na+ or Li+) for larger alkali ions delivered to the surface,
respectively, K+ or Na+. This will result in a near surface molar volume expansion that,
because of the constraint of the unexpanded inner portion of the glass body, will generate
strong compression stress on the surface balanced by a tensile stress in the inner glass body.
This process, as it is the main subject of this booklet, will be treated in greater detail in section
2 for the theoretical aspects and in section 3 for the application aspects.

The second process that we will call shortly “Glazing” is based in the development of a
surface layer whose thermal expansion coefficient is lower than the one of the glass matrix.
This layer will be chemically created at high temperatures, upon cooling the higher
contracting matrix forces the surface layer to be in a compression state while the matrix will
develop a balancing tensile stress. Typical processes that works in this way are:

- Dealkalization of glass surface by steam or SO2 that produces silica-rich surfaces


- Ion exchange with smaller ions for internal larger ions at above strain point temperature will
create a lower thermal expansion coefficient.

21
2. Chemical strengthening by ion exchange
By this definition we identify those ion exchange processes performed below the glass
transition temperature. According to its nature ion exchange is not a thermodynamic
equilibrium process, the flux of ions inward and outward glass matrices is a kinetic process
driven by gradients in the electrochemical potentials of the involved ionic species. I like a
definition of this type of strengthened glass that reads: “Ion-exchanged glass belongs to the
class of “forbidden glass” i.e., a glass with a structure that is not attainable by cooling from a
melt using any possible cooling path”( J.C. Mauro, R.J.Loucks;[67])

Let’s assume the following definition: Ion Exchange (IX) in silicate glasses is a non-
equilibrium thermodynamic process between a glass and an ion source, driven by gradients in
electrochemical potentials of relatively mobile network modifiers. This means that to activate
ion exchange we need a glass matrix with relatively mobile ions (typically monovalent alkali)
and a ion source with again available mobile ions. The contact of this glass matrix with the
ion source generates the ion exchange as a consequence of a gradient in the electrochemical
potentials of the involved ionic species. During the ion exchange process Si-O network of the
glass may be considered stable. Considering an ion A in the ion source (IS) and a ion B in the
glass matrix (GM) the IX process may be schematically written as:

AIS + BGM ⇔ BIS + AGM (2.1)

And pictorially represented in Figure 9 where “A” is Potassium K+ and “B” is Sodium Na+.

Figure 9 – Representation of IX process

Typical assumptions for IX in silicate glasses are:

- Mobilities of cations in the ion source are much higher than those in the glass matrix
- Interface reactions between ion source and glass are such fast to get an almost
instantaneous equilibrium condition at the interface.

22
From the above assumptions it is clear that the process rate determining factor is the inter-
diffusion of the involved ions in the glass matrix. It may happen that the two above conditions
are not met during real technological processes, this may occur in presence of ion source
contaminations that reduce ion mobilities or induce interface blocking mechanism.
The main physical and chemical effects of IX are, in a rational order:

Near surface molar volume change


Near surface molar concentration change
Near surface refractive index change
Residual stress induction

Both refractive index profile and residual stress profile are a consequence of the concentration
of the invading ions in the glass matrix. The refractive index profile is of relevance in optical
applications (optical waveguides) and in measurements (optical determination of residual
stress by birefringence) while the stress profile is relevant in strengthening applications.
In both cases a key role is played by the ion concentration of the invading ions as a
consequence of the IX process.

From: A.K.Vershneya, P.Kreski,G.Olson / DGG-GOMD 2014 – Aachen (Germany)

Alkali-containing silicate glass


Exchange smaller host ions for larger invading ions from a molten salt
bath, usually KNO3
Temperature managed to promote interdiffusion, and limit viscous
relaxation
Surface compression so generated increases glass strength

Figure 10 – IX Summary: pictorial and conditions

In Figure 10 (courtesy from Arun K. Varshneya) a good summary of IX process is depicted


and conditions are indicated to achieve a strengthening effect.

23
2.1 Ion Exchange theory

To better understand ion exchange theory let’s give some information on the chemical
structure of silicate glasses. For the sake of this booklet let’s refer to silicate glass as “glass”.
This material can be structurally considered a non-crystalline solid (solid without a long range
crystallographic order) made of a silicon oxide (SiO2) tetrahedral structure. This is more or
less the structure of fused silica. Such type of material has a very high temperature
transformation range of viscosity, this means that to perform technological operations like
melting, forming, bending we have to rise up temperature at very high values. To reduce
practical working temperatures alkali elements are introduced in the glass composition as
oxides (Na2O,K2O ..). The cations (positive ions) of these introduced oxides break in some
way the continuity of the Silicon/ Oxigen bonding network creating so called “NBO - non-
bridging-oxygens” that is oxygens not directly bonded to Silicon (Bridging Oxygens) but
bonded to the introduced cations. This introduction effectively reduces the working
temperatures but reduce also chemical stability of glass (Alkali leaching by water attack). To
increase stability alkaline earth elements are introduced (CaO, MgO). Finally “glass” can be
represented as in figure 11.

Figure 11 – Soda-lime glass chemical structure

From the structural point of view alkali and earth alkaline ions can be considered as relatively
mobile in the glass structure (because of their bond to the glass network through the NBO
Oxygens) and they are called glass network modifiers. They are bonded to the non-bridging
oxygen atoms of the silica network. Their mobility within the glass network is strongly
depending on temperature: as they get enough energy (e.g. by glass heating) they can move
relatively free in the glass network. In particular, if there is a chemical potential gradient of
the alkali ions, this acts as a thermodynamic force generating a flux of ions in the direction
that reduce the gradient. There is an important structural difference of the above introduced
glass structures (Soda-Lime silicates see Figure 12) with other types of glass silicates
structures, for example aluminosilicates. The introduction of alumina A2O3 in the glass
composition modifies the structure because Al is acting as glass former (see Figure 13). The
structure of Alkali-Aluminosilicate glass (up to the molar ratio limit [Al2O3]/[M2O] ≤ 1) is
such that mobility of ions is significantly increased while the network connectivity is
preserved.

24
Figure 12 – Structure of Soda-lime Silicate

Non bridging oxygen of Soda-Lime silicate glass results in a stronger bond with alkali ions
while fourfold negative [AlO4]- network forming units of aluminosilicate glass result in a
weaker bond with alkali ions maintaining the full connectivity of the glass matrix network.

Figure 13 – Structure of Sodium-Alumino Silicate

2.2 Kinetics of Ion Exchange: residual concentration of invading ions

We have already seen the relevance of ion concentration in the determination of residual
stress. In order to calculate the invading ions concentration we will enter into a discussion of
IX kinetics. A good presentation of a full mathematical model of IX including externally
applied electric field can be found in J.Albert, J.W.YLit – [45]. I will present here a shortcut
derivation and summary of main results.
Indicating with :

cA - Molar Concentration of ion A (mol/cm3)


vA - Drift velocity of ion A (cm/s)
JA - Molar Flux of ion A (mol/cm2s)
bA – Mobility (cm2/J·s)
F – thermodynamic driving force (N)

25
µA – Chemical or electro-chemical potential (J/mol)
NA – Avogadro number (1/mol)

The following flux equation holds (Molar Flux is the number of particles in moles flowing
through a cross section – cm2 – in unit time):

J A = CA ⋅ vA (2.2)

Drift velocity of the ions is expressed in terms of mobility bA and of thermodynamic driving
force F as:

v A = bA ⋅ F (2.3)

Molar flux and thermodynamic driving force are, in this way, simply related by a linear
relationship, this means that the flux is generated by the thermodynamic driving force:

J A = C A ⋅ bA ⋅ F (2.4)

The relationship between the thermodynamic driving force and the electrochemical potential
is:

1 ∂µ A
F =− ⋅ (2.5)
N A ∂x

By coupling equations (2.4) and (2.5) we can write an equation for the ion flux in terms of
chemical potential:

b A ∂µ A
J A = −C A ⋅ ⋅ (2.6)
N A ∂x

It can be introduced a self-diffusion coefficient DA:

R
D A = bA ⋅ k ⋅ T = bA ⋅ ⋅T (2.7)
NA

Where k is the Boltzmann constant and R the gas constant (R=k·NA) . With this definition we
can write equation (2.6) as:

C A ⋅ D A ∂µ A
JA = − ⋅ (2.8)
R ⋅T ∂x

The electrochemical potential, in general, may be written in terms of the chemical activity –
a(CA) - which is a function of molar concentration:

µ A (C A ) = µ0 + RT ln a(C A ) + ZFϕ (2.9)

26
Where Z is the atomic number F the Faraday constant and ϕ is the electrical potential
∂ϕ
connected to the electric field E = . Using this definition in equation (2.8) we obtain:
∂x

 ∂C ∂ ln a A Z ⋅F ⋅E 
J A = − DA ⋅  A ⋅ + CA ⋅  (2.10)
 ∂x ∂ ln C A R ⋅T 

Equations of type (2.6), (2.8) and (2.10) are called Nerst-Planck equations and are the basic
equations for the analysis of IX kinetics.

If a glass with network modifiers ions B is put in contact at relatively high temperature with a
means (for example a molten salt of specie A) that can provide other ions A with the same
oxidation state (valence) two electrochemical potential gradients are generated and a chemical
ion-exchange reaction (subscripts G and S means Glass Matrix and Salt Bath respectively)
holds (see Figure 14)

AS + BG ⇔ BS + AG (2.11)

Figure 14
Glass in contact with a Potassium salt batch Glass in contact with a Potassium salt batch
Before Ion Exchange After Ion Exchange

In this case species are B→Na+ and A→K+. The ion exchange process can be represented
according to figure 15.

27
Figure 15 – Ion Exchange in terms of gradients of electro-chemical potential and IX effects

From Figure 15 the relevant effects of IX are evident: a residual concentration of invading A
ions and a residual stress in the glass matrix.
The kinetics of the ion exchange process that involve two monovalent (Z=1) ions A and B
according to (2.11) can be described by the following kinetic equations:

 ∂C ∂ ln a A F ⋅E
J i = − Di ⋅  i ⋅ + Ci ⋅  ; i = A,B
 ∂x ∂ ln Ci R ⋅ T 
(2.12)
∑J
i
i =0

Where the second condition simply states the equimolar flux condition: JA+JB = 0 for. This
last condition is no more true in case of an externally applied electric field E. An internally
build up electric field result from differences in ion mobility of the involved species in IX.
This means that the electric field in the flux equation (2.12) can be either an externally applied
electric field or an “internal” electric field build up because of different mobilities of the
involved ions. This last condition is evaluated applying the equimolar flux condition. With the
equimolar flux condition and by the continuity equation it can be derived the diffusion
equation for IX without application of external electric field.
A good way to go ahead in this task is to introduce the molar fraction concentrations :

28
CA CB
ca = cb = c a + cb = 1
C A + CB ; CA + CB ; (2.13)

∂ ln a A
Approximating the activity term ∂ ln C ≅ 1 the problem (2.12) can be transformed:
A

 ∂c F⋅E  ∂c F ⋅E
− J a = Da ⋅  a + c a ⋅  ; − J b = Db ⋅  b + cb ⋅ 
 ∂x R ⋅T   ∂x R ⋅T 
(2.14)
Ja + Jb = 0

This set of equations, duly algebraically developed, allows to express the flux of the ions “A”
entering in the glass as:

∂c a
− J a = D AB ⋅
∂x
(2.15)
D A ⋅ DB DA D
D AB = ; D AB = ; α = 1− A .
D A ⋅ c a + D B ⋅ cb 1 + α ⋅ ca DB

Where DAB is defined: interdiffusion coefficient. Through the continuity equation:

∂c a ∂
= − (J a ) (2.16)
∂t ∂x

We can finally establish a non-linear diffusion equation for the concentration:

∂c a ∂  DA ∂c 
=  ⋅ a  (2.17)
∂t ∂x  1 − α ⋅ c a ∂x 

This equation can be further manipulated considering DA and DB (and so α) constants leading
to:

∂c a DA  ∂ 2 ca α  ∂c a  
2

=  2 + ⋅   (2.18)a
∂t 1 − α ⋅ c a  ∂x 1 − α ⋅ c a  ∂x  

In case of application of an external field “E” the equation will be modified as:

∂ca DA  ∂ 2ca α  ∂ca  e ⋅ E ∂ca 


2

=  2 + ⋅ − ⋅  (2.18)b
∂t 1 − α ⋅ ca  ∂x 1 − α ⋅ ca  ∂x  k ⋅ T ∂x 

These last non-linear diffusion equations (2.18)a and (2.18)b can be further approximated in
linear, more familiar equations, considering an “almost constant interdiffusion” coefficient.
Equation (2.18)a reduces to:

29
∂c a ∂ 2c
= D AB ⋅ 2a (2.19)a
∂t ∂x

Similar linearization arguments may be used with the equation for externally applied electric
field (2.18)b leading to the following linearized diffusion equation with externally applied
electric field:

∂ca ∂ 2c ∂c e ⋅ DAB
= DAB ⋅ 2a − µe ⋅ E ⋅ a ; µe = (2.19)b
∂t ∂x ∂x k ⋅T

Where µe is the electrical mobilities of incoming ions. Linear approximation of (2.18)a and
(2.18)b holds when the self-diffusion coefficients of both ions are close to each other (that
means α≈0) and almost independent of concentrations.

Even though, under some conditions, inter-diffusion coefficients can be reasonably


considered “independent” from concentration, its dependence from temperature cannot be
neglected. An acceptable approximation is to consider that it follows an Arrhenius-type
dependence on the temperature

 ∆E 
− 
DAB (T ) = Do ⋅ e  k ⋅T 
(2.19)c

where ∆E is the energy barrier for a jump between two adjacent sites and Do is the pre-
exponential factor. The first attempt to derive an expression for the energy barrier is due to
Anderson and Stuart, which assumed that it was constituted by two terms, related to the
electrostatic interaction with the neighbors and to the elastic strain of the network, treated as a
continuum.

Equation (2.19)a is well known in the scientific literature as “The diffusion equation”. The
solution to equation (2.19)a provides the concentration of the incoming ions “A” at any time
and at any point of the glass matrix. The solution exists and is unique only if an initial
condition and boundary conditions are defined:

c a ( x,0) = c t =0 ( x ) - Initial condition


c a (0, t ) = c s (t ) - Boundary condition of the first kind (Surface concentration)

[c a ( x, t )]x =0 = α ⋅ c s (t ) - Boundary condition of second kind
∂x

Let’s write down some important solution to the diffusion equation (2.19)a [B13] .

(i) Semi-infinite medium ; constant surface concentration with zero initial


concentration:

ca(0,t) = cs
(2.20)a
 
( )
z
c a ( x, t ) = c s ⋅ erfc
x  ; erf ( z ) = 2 ⋅ exp − β 2 dβ
 2⋅ D ⋅t 
 AB  π ∫0

30
Where the erf function is called “error function” erfc(z) = 1-erf(z) is the complementary error
function (see the erfc plot in figure 16)
A solution to equation (2.19)b under boundary condition (i) that is constant source, can be
found as:

cs   x − µe ⋅ E ⋅ t  µ ⋅E⋅x x + µe ⋅ E ⋅ t 
ca ( x, t ) = erfc   + exp( e ) ⋅ erfc( ) ⋅ (2.20)b
2  2⋅ D ⋅t  DAB 2 ⋅ D ⋅ t 
  AB  AB

For large electric fields it can be shown that the second term in eq. (2.20)b can be neglected
and diffusion does not play a significant role giving a step-like profile

1,0

0,9

0,8

0,7

0,6
erfc (z)

0,5

0,4

0,3

0,2

0,1

0,0
0,0 0,5 1,0 1,5 2,0 2,5

Figure 16 –Plot of the complementary error function – y = erfc (z)

(ii) Semi-infinite medium ; constant surface concentration with co initial


concentration:

ca(0,t) = cs; ca(x,0) = co


(2.21)
ca ( x, t ) − cs  x 
= erf  ;
co − cs  2⋅ D ⋅t 
 AB 

As co=0, (ii) reduces to (i).

(iii) Semi-infinite medium ; diffusion from an infinitesimally thin layer with zero
initial concentration:

x2
Q − Q
ca ( x, t ) = ⋅ e 4⋅D⋅t ; ca (0, t ) = (2.22)
π ⋅ D ⋅t π ⋅ D ⋅t

Where Q is the amount of ions per unit area in the layer.

31
(iv) Semi-infinite medium – constant surface flux (boundary condition of second
kind) where co is the concentration remote from the surface.

∂c a α
− D AB ⋅ = α ⋅ (c o − c s ) ; h =
∂x D AB
(2.23)
    
c a ( x, t ) = co ⋅ erfc
x
 2⋅ D ⋅t
( )
 − exp h ⋅ x + h 2 ⋅ D AB ⋅ t ⋅ erfc

x
 2⋅ D ⋅t
+ h ⋅ D AB ⋅ t 

  AB   AB 

(v) Slab –l ≤ x ≤ l – constant surface concentration - cs, initial concentration


c1(x,0)=0

ca(0,t) = cs

 4 ∞ (− 1)n  (2 ⋅ n + 1)2 ⋅ π 2 ⋅ D AB ⋅ t    2 ⋅ n + 1  
c a ( x, t ) = c s ⋅ 1 − ⋅ ∑ ⋅ exp −   cos  ⋅ π ⋅ x   (2.24)
 π n=0 2 ⋅ n + 1  4⋅l2   2 ⋅ l  

with initial concentration ca(x,0) = co

4 ∞ ( −1)  ( 2 ⋅ n + 1)2 ⋅ π 2 ⋅ DAB ⋅ t    2 ⋅ n + 1


n
ca ( x, t ) − co 
= 1− ⋅ ∑ ⋅ exp  −   cos  ⋅ π ⋅ x  (2.25)
cs − co π n=0 2 ⋅ n + 1  4⋅l 2
   2 ⋅ l 

In Appendix A1 it will be presented a solution for an IX with a time variable boundary


condition where the Ion Source is constant up to a certain time (as in case (i) supplying ions
and maintaining a constant surface equilibrium concentration) and, when all available ions
 ∂c 
have been introduced in the glass it becomes constant flux zero  a  = 0 . This boundary
 ∂x  x =0
condition may be assimilated to a IX from a thick deposited layer with a limited quantity of
available ions or to an IX with a subsequent thermal treatment outside the continuous ion
source.
The typical and most implemented condition used in glass chemical strengthening by ion
exchange is by the immersion of the glass in a well stirred molten salt ion source bath. In this
case the process can be depicted as in Figure 17.

32
Figure 17 – Ion Exchange Process

Surface equilibrium condition is achieved in a far shorter time τ << t in respect to the total
immersion time so that surface concentration assumes in a very short time and maintain for all
process duration an equilibrium constant values. This matches condition (i) and (ii). In
general, if:

co is the initial concentration of invading ions in the glass matrix,


CsK is the equilibrium surface concentration of invading ions during the IX process,
CsNa is the surface concentration of un-exchanged ions

the concentration of invading ions is:

 x 
cK ( x, t ) = csK + (co − csK ) ⋅ erf   (2.26)
 2⋅ D ⋅t 
 AB 
That, as co=0, reduces to:

 x 
cK ( x, t ) = csK ⋅ erfc   (2.27)
 2⋅ D ⋅t 
 AB 

The concept of exchange ratio “r” is important to define the condition of maximum of surface
concentration of invading ions. In this condition CsNa=0 and r=1.

csK − co
r= (2.28)
csK + csNa − co

By this definition we can define the percentages of exchanged ions on the glass surface: if r=1
we have 100% exchange, if r=0.8 we have a 80% of IX if r=0 we have 0% of IX.
One of the effects of ion exchange is the weight increase of the glass articles. This effect is of
some importance for the estimation of the diffusion coefficient and for process control. Let’s
introduce some definitions related to weight gain during IX.

33
QAB – exchanged molar flux (mol/cm2)
S – Glass article surface (cm2)
∆MK – Molecular weight difference between ion species =MA-MB (g/mol)

The weight increase during IX may be expressed as:

∆W = Q AB ⋅ S ⋅ ∆M K (2.29)

QAB can be expressed in a very general way (see Appendix A2)as:

 ∂c 
Q AB = −2 ⋅ D AB ⋅ t ⋅   = k ⋅ t (2.30)
 ∂x  x=0

In this case k is a constant. The flux ΦW (g/cm2) is:

∆W
ΦW = = k ⋅ ∆M K ⋅ t = χ ⋅ t (2.31)
S

χ = k ⋅ ∆M K (2.32)

In a Potassium K+/Sodium Na+ ion exchange, setting the following definitions:

Φ K = k ⋅ M K ⋅ t - Flux (g/cm2) of Potassium ions entering in the glass

Φ Na = k ⋅ M Na ⋅ t - Flux (g/cm2) of Sodium ions entering in the salt bath

it results:

Φ W = k ⋅ (M K − M Na ) ⋅ t = k ⋅ M K ⋅ t − k ⋅ M Na ⋅ t = Φ K − Φ Na (2.33)

The ion-exchange kinetics is, in general, a non-linear diffusion problem. The solution of the
ion-exchange kinetic equations has the purpose to evaluate the concentration of the incoming
ions in the glass matrix. Analytical solution to the non-linear equations are difficult to find.
Linearization (basically considering a constant interdiffusion coefficient) is not such a bad
approximation considering that the ion exchange process is normally performed in isothermal
conditions. This last approach leads to classical diffusion equations that can be easily
solved. Numerical approaches are also possible. Non-linear diffusion equation (2.17) can also
be written as:

∂c a ∂  ∂c 
=  D AB (c a ) ⋅ a  (2.34)
∂t ∂x  ∂x 

Any approach either non-linear either linearized needs the previous evaluation of the
interdiffusion coefficient. A suitable technique is known in literature as Boltzmann-Matano
technique. According to this method the concentration depth profile has to be known :

34
c a = c a ( x, t ) (2.35)

Than the interdiffusion coefficient at concentration ca is:

c
1  dx  a
2 ⋅ t  dt  ∫0
D AB (c a ) = − ⋅   ⋅ x ⋅ dc (2.36)

Details of this method may be found in [B 16] pages 230-232.


A simple method to evaluate the interdiffusion coefficient is based on the linearized approach
of equation (2.19) and using the weight gain results . According to equation (2.19) taking
solution (2.20), the molar flux QAB can be written for an Ion-Exchange process:

DAB ⋅ t
QAB = 2 ⋅ CsA ⋅ (2.37)
π

Where Cs,A is the surface concentration of ion A. If the molar flux QAB is known from weight
gain (2.29) than a “Mass average diffusion coefficient DM” can be calculated:

QAB2
⋅π
DM = (2.38)
4 ⋅ Cs , A ⋅ t
2

DM is a quick acceptable estimation of the “interdiffusion coefficient”.With this parameter it


is also possible, under the same approximation assumptions, to estimate the depth of the
chemically modified (exchanged layer as)

Cδ ≈ DM ⋅ t (2.39)

2.3 Stress build-up and relaxation

The introduction in the glass structure of larger ions in the place of smaller ones at
temperatures below the transformation range generates a relative volume increase ∆V/V. So
the strain produced by this “stuffing” effect, in absence of viscous relaxation, in a not
constrained glass volume simply is:

1  ∆V 
ε = ⋅  (2.40)
3  V 

Stress is generated as a result of constraints introduction to the initial volume V that, in some
way, prevent its free expansion as invading ions are coming in.
Development of stresses during IX requires consideration of two competitive processes:

Stress build up due to molecular stuffing


Stress relaxation due to glass viscosity and other mechanism

The stress build up by molecular stuffing can be calculated by analogy to thermal stress
[14],[39]. In order to set up this calculation let’s consider an infinite slab of thickness 2L

35
exposed from two sides in the x axis direction to ion exchange (see Figure 18). Ion stuffing
occurs only in the +/- x direction.

Figure 18 – Stress build-up

The unrestrained free strain generated in the y and z directions by ion stuffing is (Figure 18
(B)) directly and linearly related to the ion concentration through a Linear Network Dilatation
coefficient (LNDC) – B :

ey ( x, t ) Free = ez ( x, t ) Free = B ⋅ c ( x, t ) ; (2.41)

In this case we have free strain that means free expansion of the glass element. Clearly no
stresses are generated in a totally free expansion. Dilatation (expansion) occurs because the
incoming, invading ions have a larger volume in respect to out-going ions. If the process is
performed below the glass transition temperature than the glass body is not free to expand and
the uneven dilatations in the y and z directions (see Figure 19) are suppressed because of
compatibility criteria. So the body swells freely in the x direction but can expand only
uniformly in the y and z directions. The net expansions along y and z directions are given by
the average expansions of the x elements over the 2L thickness (Figure 18 (C)).

L
1
2 ⋅ L −∫L
[ey ( x, t )]Net = [ez ( x, t )]Net = B ⋅ c( x, t ) ⋅ dx (2.42)

The “true strain” of an x element is the difference between the free body expansion (2.41) and
the effective net expansion (2.42), resulting

L
1
[ε y ( x, t )]x = [ε z ( x, t )]x = B ⋅ c( x, t ) −
2 ⋅ L −∫L
B ⋅ c( x, t ) ⋅ dx (2.43)

Residual stress can be calculated from (2.43) as follows:

36
B ⋅ E ⋅ c ( x, t )
L
E
[σ y ( x, t )]x = [σ z ( x, t )]x = −
2 ⋅ (1 −ν ) ⋅ L −∫L
+ B ⋅ c( x, t ) ⋅ dx (2.44)
1 −ν

Where E is the elastic (Young) modulus and ν is the Poisson ratio of the glass. Figure 19
represents stress generated due to ion stuffing through x direction and enforcement of
compatibility criteria in y and z directions.

Figure 19 – Geometric representation of stress build-up after IX with diffusion from x and –x
directions with enforcement of compatibility criteria along y and z directions.

This is true in case c(x,t) is an even function over x. In case c(x,t) is an odd function a bending
stress appears. The balancing condition of zero moments over the cross section introduce an
additional term in the (2.44) and finally it results:

B ⋅ E ⋅ c ( x, t ) 3⋅ x ⋅ E
L L
E
σ ( x, t ) = − ∫ 2 ⋅ L3 ⋅ (1 −ν ) −∫L
+ B ⋅ c( x, t ) ⋅ dx + ⋅ x ⋅ B ⋅ c( x, t ) ⋅ dx (2.45)
1 −ν 2 ⋅ (1 −ν ) ⋅ L − L

For the sake of this booklet we will always consider even concentration distribution and
constant B values. Under these hypothesis the main stress equation 2.45 reduces to:

B⋅E
σ ( x, t ) = − ⋅ c ( x, t ) − c (t )  (2.46)
1 −ν 

having recognized that the average concentration over the glass thickness is just:

L
1
2 L −∫L
c (t ) = c ( x, t ) ⋅ dx (2.47)

37
The first term in equation (2.46) represent compression stress while the second represents the
inner tensile stress balancing near surface compression as will be formally demonstrated later
on. From equation (2.46) we can define three important characteristics: (the glass slab
thickness d=2L is ranging from –L to L).

Surface Compression – SC
B⋅E B⋅E
S C = σ ( ± L, t ) = − ⋅ [ c ( ± L, t ) ] = − ⋅ co (2.48)
1 −ν 1 −ν

Compression layer depth (CaseDepth) - CD (defined as distance from glass surface to zero of
compression).

B⋅E
σ (cD , t ) = − ⋅ c(cD , t ) − c (t )  = 0 ; → c(cD , t ) = c(t ) (2.49)
1 −ν 

Central tension

B⋅E
ST = σ (0, t ) ≈ − ⋅ c (t ) (2.50)
1 −ν

The above introduced parameters are described in Figure 20.

Figure 20 – Concentration profile and residual stress in a glass article – Parameters


identification

As ion-exchange is performed at a relatively high temperature for a rather long process time
stress relaxation cannot be ignored or neglected. This effect can be included introducing a
normalized relaxation function of time: R(t) such that: R(t)=1 if no viscous relaxation is
involved, R(t)=0 for totally relaxed stress.

0 < R (t ) < 1 ; R (0) = 1 ; R (∞ ) = 0 (2.51)

Using this definition the resulting stress, including viscous relaxation effects, is:

B⋅E ∂
t
σ ( x, t ) = − ⋅ ∫ R(t − t ') ⋅ c( x, t ') − c(t )  ⋅ dt ' (2.52)
(1 −ν ) 0 ∂t

38
When relaxation effect are negligeable R(t)=1 for all t and equation (2.52) reduces to (2.46) as
c(0,t)=0 and, obviously c(t ) = 0 . Another form of this stress equation can be obtained by
integration by parts [87],[90]:

B⋅E  ∂R(t − t ') 


( )
t
σ ( x, t ) = − ⋅  c( x, t ) − c(t ) − ∫ ⋅  c( x, t ') − c(t ')  ⋅ dt ' (2.53)
(1 −ν )  0
∂t ' 

The first term of equation (2.53) is the unrelaxed part of the stress while the second term is the
relaxation contribution.

A well known and popular relaxation function is the Maxwell visco-elastic normalized
stretched relaxation function written as:

  t γ 
R(t ) = exp  −    (2.54)
  τ  

Where τ is the relaxation time and γ accounts for the non-Maxwellian behaviour (γ ≈ 3/7
value usually considered for structural relaxation). If we use the (2.20) expression for
concentration :

 x 
c( x, t ) = C 0 ⋅ erfc  (2.55)
 2⋅ D⋅t 

It results:

∂ C0 ⋅ x  x2 
c ( x, t ) = ⋅ exp −  (2.56)
∂t t ⋅ 4 ⋅π ⋅ D ⋅t  4 ⋅ D ⋅ t 

The average concentration value is:

4 D ⋅t d c(t ) 2 D
c(t ) = ⋅ ; = ⋅ (2.57)
d π dt d π ⋅t

And equation (2.52) results:

B ⋅ E ⋅ co 
t
x  − x2  2 D    t − t ' γ 
σ ( x, t ) = −
(1 −ν ) ∫0  t '⋅ 4 ⋅ D ⋅ t '
⋅  ⋅ exp  − ⋅  ⋅ exp  −    ⋅ dt ' (2.58)
 4⋅ D ⋅t '  d π ⋅t '    τ  

Using the alternative stress equation (2.53) and taking the derivative of the relaxation function
(2.54) we obtain an equivalent expression of residual stress that may have some advantages in
computational algorithms.

39
B ⋅ E ⋅ co  x 4 D ⋅t 
σ ( x, t ) = − ⋅  erfc( )− ⋅   −
(1 −ν )  2⋅ D ⋅t d π  
γ (2.59)
γ
t
 t −t'
γ −1  t −t '    x 4 D ⋅t ' 
−  τ 
−∫  ⋅  ⋅e  ⋅  erfc( )− ⋅  ⋅ dt '
τ  τ    2⋅ D ⋅t ' d π 
0
 

Relaxation effects in chemically strengthened glass are evident from the shift of the residual
stress maximum towards the inner section of the glass and from the reduction of the residual
stress values (see figure 21).

Figure 21 – Stress profile - Blue line relaxation effect not significant – Red line significant
relaxation effect.

The important point to have in mind is that the relaxation process has a relevant importance
when the difference between process temperature and strain point temperature is less than
about 100°C. For soda-lime glass the strain temperature is around 514°C, this means that at a
process temperature of about 400°C we can expect stress relaxation effects when the process
is held for a significant time (more than 8 hours).

The front term of equations (2.58) and(2.59) is just the surface compression at zero time:

B ⋅ E ⋅ co
Sc (0) = − (2.60)
1 −ν

It is an experimental evidence that surface compression relaxes and the functional dependence
of this relaxation can be found considering equation (2.52).

With the values of surface compression σS obtained in this way and the data about CD the
strength increase of the glass ∆σR can be evaluated knowing the original glass strength (flaw
length) and using the equations that will be given in next chapter.

Equation (2.52) can be written separating the two terms under derivative:

B⋅E 1 ∂
t

σ ( x, t ) = − ⋅ ∫ R (t − t ') ⋅  c( x, t ') − c(t ')  ⋅ dt ' =


(1 −ν ) 0 ∂t '  
(2.61)
B⋅E ∂ B⋅E ∂
t1 t1

= ⋅ ∫ R (t − t ') ⋅ c(t ')  ⋅ dt ' −


  ⋅ ∫ R (t − t ') ⋅ [ c( x, t ') ] ⋅ dt '
(1 −ν ) 0 ∂t ' (1 −ν ) 0 ∂t '

40

applying condition (in this case “d” is glass plate thickness) [c( x = d / 2, t )] = 0 ; (even
∂t
condition in the mid section of the glass plate) we can identify the central tension ST term:

B⋅E ∂
t
ST (t ) = ⋅ ∫ R (t − t ') ⋅ c(t ')  ⋅ dt ' (2.62)
(1 −ν ) 0 ∂t '  

That, in case of no relaxation (R(t) = 1 for any t>0), results:

B⋅E
ST (t ) = ⋅ c (t ) (2.63)
(1 −ν )

From this argument surface compression SC(t) can be easily obtained considering that we
have initially (t=0) a sudden variation of surface concentration that we can represent with the
following boundary condition at glass surface (using the Heaviside θ function):

c(0, t ) = co ⋅ ϑ (t )
∂ϑ (t )
t ≤ 0............ϑ (t ) = 0 ; = δ (t ) (2.64)
∂t
t > 0............ϑ (t ) = 1

(where δ(t) is the Dirac delta function) surface compression results:

B ⋅ E ⋅ co B ⋅ E ⋅ co
t
SC (t ) = ST (t ) − ⋅ ∫ R(t − t ') ⋅ δ (t ') ⋅ dt ' = ST (t ) − ⋅ R(t ) (2.65)
(1 −ν ) 0 (1 −ν )

Again, when relaxation is not relevant, R(t)=1 and surface compression results:

B ⋅ E ⋅ co B⋅E
SC (t ) = ST (t ) − =− ⋅ (co − c(t )) (2.66)
(1 −ν ) (1 −ν )

Temperature (°C) Zero time Surface Relaxation time τ (s)


Compression SC(0) (MPa)
425 790 1.94·106
450 744 6.30·105
475 723 1.37·105
Table 3 Zero time surface compression and relaxation time for Soda lime silicate glass (From
A.K.Varshneya et al JNCS – 2015)

Recently an analytical solution (closed form solution) has been proposed [95] for the
calculation of the residual stress in ion exchanged silicate glass for both constant and not
constant relaxation coefficients. The non-constant coefficients either in relaxation parameters
[95] or in the LNDC [58] seems to be the way to justify the residual stress shift of the
maximum (see Figure 21) that is usually found for longer immersion times or higher
temperature IX processes. In [95] a Prony series (generalized Maxwell model) approach has
been used for the evaluation of the relaxation function.

41
In table 3 some values are reported for zero time surface compression. The time evolution
with viscous relaxation of surface compression can be calculated using the values of table 3 in
equation (2.65) and considering a relaxation function (2.54) with γ=3/7.

2.4 Strength issues of ion-exchanged glass

As discussed in the first part of this booklet glass strength is depending from surface flaws
statistics and finally from the interaction of residual stress with the flaws geometry and the
environment. Let’s recall the main points already introduced at point 1.3.3.
The introduction of a residual stress field:

σ R = σ R (x) (2.67)

Modifies the stress intensity equation introducing a stress intensity factor due to the residual
stress as KR: that can be calculated using the Green function as:

2 ⋅Y ⋅ c
c
σ R ( x)
KR = ⋅∫ dx (2.68)
π 0 c2 − x2

This contribution is introduced in the expression of the stress intensity factor as:

KI = σ a ⋅Y ⋅ c + KR (2.69)

Equation (2.96) at failure reads:

K IC = σ b ⋅ Y ⋅ c + K R (2.70)

For compressive residual stress KR < 0 and:

K IC − K R K IC + K R
σb = = (2.71)
Y⋅ c Y⋅ c

This means that for compressive residual stresses fracture happens under a higher applied load
than for residual stress free state.

In chemical strengthening we have a shallow depth of compression layer with a pretty high
steep of stress curve. The stress gradient is of the order of hundreds of MPa over tens of
micron. A possible approximation of the residual stress field can be proposed by the
following first order linear expression:
 x 
σ R ( x) = − S C ⋅ 1 −  ; x ≤ CD
 C D 

(2.72)
σ R ( x) = ST ; x > CD

Where Sc is the surface compression and ST the central tensile stress and SC >> ST

42
In Figure 22 the representation of this approximated residual stress profile (2.72) is
represented as a linear curve.

Figure 22 – Stress profile in chemically strengthened glass – Linear approximation

Following this approach and using the stress profile represented in equation (2.72) the integral
of equation (2.68) can be easily evaluated analitically so that KR can be calculated. Two cases
can be distinguished:

(A) – CD > c (compression layer depth - Case Depth – larger than flaw depth
– c) see figure 23.

Figure 23 – Flaw depth smaller than case depth

− 2 ⋅Y ⋅ c π c
KR = ⋅ SC ⋅ ( − ) (2.73)
π 2 CD

2 ⋅Y ⋅ c π c
K IC = σ b ⋅ Y ⋅ c − ⋅ SC ⋅ ( − ) (2.74)
π 2 CD

K IC 2 π c   2 c 
σb = + SC ⋅ ⋅  −  = σ 0 + SC ⋅ 1 − ⋅ 
Y⋅ c π  2 C D   π CD 
(2.75)

43
Where σ0 is the glass strength before strengthening. Equation (2.77) express the strength
increase after chemical strengthening:

 2 c 
∆σ = σ b − σ 0 = S C ⋅ 1 − ⋅  (2.76)
 π CD 

In the approximation CD >> c (valid for example in thermal tempering) it results:

σ b = σ 0 + SC (2.78)

Which is exactly equation (1.30) derived in another way just for thermal tempering.

B) CD < c (compression layer depth - Case Depth – smaller than flaw depth – c) see
Figure 24

Figure 24 – Flaw depth larger than case depth

−2 ⋅ Y ⋅ c  C  c c2 
KR = ⋅ SC ⋅  arcsin  D  − + −1  +
π   c  CD
2
CD 
  (2.79)
2 ⋅Y ⋅ c π C 
+ ⋅ ST ⋅  − arcsin  D 
π 2  c 

2 CD 2 c 2 c2
∆σ = ⋅ ( SC + ST ) ⋅ arcsin( ) − ⋅ SC ⋅ + ⋅ SC ⋅ − 1 − ST (2.80)
π c π CD π CD2

An example of the above results: let’s consider two glass types with same chemical
composition but different surface flaws depth characteristics. The glasses have been submitted
to the same chemical strengthening process that results in a surface compression of 450 MPa
and a case depth of 30µm summarizing:

A) c = 15 µm ; Chemical strengthening – SC = 450 MPa ; CD = 30 µm


B) c = 50 µm ; Chemical strengthening – SC = 450 MPa ; CD = 30 µm

Strength is calculated following the above analysis. Results are reported in table 4.

44
Glass Flaw Original Surface Case Strength Final
Type Depth Strength Compression Depth increase Strength
c(µm) σ0 (MPa) SC (MPa) CD (µm) ∆σ (MPa) σb(MPa)
Glass A 15 98 450 30 306 404
Glass B 50 53 450 30 90 143
Table 4 – Effect of chemical strengthening on two types of glass (same chemical composition
different surface defects) evaluated with the linear stress approximation.

Results of Table 4 are not very far from a realistic situation. Nevertheless the rude assumption
of linear behaviour of residual stress is very unlikely. In Figure 25 a comparison is reported of
residual stress calculated (red line) according to equation (2.59) for a IX process of 24 hours
on a soda lime silicate glass thickness 10mm with a diffusion coefficient of 8.25·10-12 cm2/s
(8.25·10-4 µm2/s)resulting in a measured surface compression of 450 MPa and a measured
case depth of 35 µm. Surface compression has been calculated (red line) according to Table 3
data, resulting in 488 MPa.

Figure 25 – Comparison of Linear vs Effective residual stress profile

The strength prediction using the effective residual stress profile requires a consideration of
flaw geometric characteristics conservation after IX. In order to compare with experimental
strength results it is necessary to consider that during ion exchange, as stress builds up and
relaxes, a mechanism is introduced that modifies the geometrical characteristics of pre-
existing surface flaws. This has been modelled introducing a flaw severity reduction factor, in
our example this has been evaluated fcR=0.69.
In the following figures strength versus original flaw depth is reported for both the first order
linear model and for the effective stress model. The two models provide results in good
agreement. The overestimation of the linear model close to the flaw tip is compensated by the

45
introduction of the flaw geometry non conservation assumption. In conclusion, the linear
model provides acceptable results only fortuitously, it can be considered only for quick
estimation of final strength after IX.

Fig.26 – Original strength (blue dashed curve) and strength after IX (24h @ 450°C) calculated
according to equation

Fig.27 – Comparison of strength values vs original flaw depths (range 10µm to 35µm). Initial
strength before IX (magenta dashed curve) Linear first order model (blue dashed curve)
effective stress (red curve)

46
Fig.28 – Comparison of strength values vs original flaw depths (range 35µm to 100µm).
Initial strength before IX (magenta dashed curve) Linear first order model (blue dashed
curve) effective stress (red curve)

This discussion demonstrates that, when in presence of severe flaws, even very high surface
compression may result in a not significant breakage strength after chemical strengthening.
The other important point to be considered about strength issues is related to sub-critical crack
growth leading to static or dynamic fatigue effects. Strengthening glass by introduction of
surface compressive residual stress is considered a way to prevent sub-critical crack growth.
The growth of surface flaws up to their critical length needs tensile stress and water vapour
environment to be activated. By strengthening, with introduction of residual compressive
surface stress, cracks are submitted to compressive stress and no sub-critical crack growth
may be activated providing that service tensile stress do not exceed surface compression. This
argument is fully true for thermally strengthened glass where we have CD >> c. For
chemically strengthened glass there are two cases under which this argument is no more true
and sub-critical crack growth may be activated and becomes an issue in the following two
cases:

(i) when c>CD and the flaw is accessible by water vapour. In this case the tip of the
flaw is in the tensile zone of the glass cross section and this means that it is
permanently under tensile stress. Because of water vapour accessibility sub-
critical crack growth may be activated leading to delayed “spontaneous”
fracture.
(ii) when a crack c>CD is generated during service life (after chemical
strengthening). In this case, locally, the effect of chemical strengthening is lost
and the glass behaves (limited to the local surface where the crack has been
generated) as a not strengthened glass. In this case the glass has lost local
strength and local protection from sub-critical crack growth mechanism.

47
2.5 LDNC Anomaly and zero time surface compression

The general build-up and relaxation model can be summarized by the following equations:

B⋅E  ∂R(t − t ') 


( )
t
σ R ( x, t ) = − ⋅  c( x, t ) − c(t ) − ∫ ⋅ c( x, t ') − c(t ')  ⋅ dt ' (2.81)
(1 −ν )  0
∂t ' 

 x    t γ 
c( x, t ) = c0 ⋅ erfc  ; R(t ) = exp  −   
 2⋅ D⋅t    τ  

It has been already pointed out that the front term:

B ⋅ E ⋅ co
Sc (0) = − (2.82)
1 −ν

Is just the zero time surface compression. In the previous discussion we have derived zero
time surface compression from surface compression values measured at different process time
through equation

B ⋅ E ⋅ co
SC (t ) = ST (t ) − ⋅ R (t ) (2.83)
(1 −ν )

We need to establish a calculation method to evaluate zero time surface compression. The key
point is the Linear Network Dilatation Coefficient (LNDC also named Cooper coefficient in
honour of professor A.R.Cooper who introduced this idea in the discussion of stress build up
in glass strengthening). By definition:

1 ∂ ln Vm
B= ⋅ (2.84)
3 ∂C

Where Vm is the molar volume of the glass. This can be calculated as follows:

1 [V − VmNa ]
B = ⋅ mK (2.85)
3 VmNa ⋅ co

Where VmNa is the molar volume of the glass before IX, VmK is the molar volume of the
glass after IX and co is the mol% of alkali in glass that exchanges. The quantity Bco is just the
linear expansion of glass network which is constrained by the un-exchanged inner layers
generating zero time surface compression:

B ⋅ E ⋅ co  E  VmK − VmNa 
SC (0) = − = ⋅  (2.86)
(1 −ν )  1 −ν   3 ⋅VNa 

Considering a Soda lime glass of composition 15Na2O·10CaO·75SiO2 the molar volume can
be calculated using the method outlined in A.K.Varshneya Fundamentals of Silicate Glasses
Ch 7 using the coefficients of Huggins and Sun (see also Appendix A9). The molar volume

48
VmK after IX calculated according to this method is the so called: ”compositionally equivalent
as-melted – CEAM. As already pointed out, ion exchanged glasses have non equilibrium
structures that does not match a CEAM one (forbidden glass [67]). This means that it will be
very unlikely that the molar volume after IX calculated in this way will represent the “after
IX” condition. Another possibility to evaluate molar volume after IX is by using a Molecular
Dynamic simulation program ( Varshneya,A.K.,Olson,G.A,Kreski,P.K.Gupta,P.K. [87] ).
Following the CEAM approach a surface compression value around 2700 MPa (2.7 GPa) is
found in large excess (factor 2.5-3.5) from typical values that can be experimentally measured
(750 MPa). This discrepancy is known in the literature as LNDC anomaly and has been
recently extensively discussed in the literature ([71],[72],[73],[74],[75],[83],[88],[89]). MD
simulation allows a large portion recovery of the discrepancy between measured and
predicted values (see table 5) nevertheless some disparity remains between the 1.2 GPa
ending stress value (predicted by the MD approach) and the observed 0.5-0.75 GPa values.

Molar Volume and Sc(0) Literature calculated MD simulation (at 450°C)


3
VmNa cm /mol 23.95 23.92
3
VmK cm /mol 26.12 25.25
Sc(0) – (MPa) 2713 1600
Sc(0) – (MPa) – cooled to 2713 1200
room temperature (20-25%
reduction)
Table 5 – Zero time Surface compression calculated from literature data for a compositionally
equivalent as melted (CEAM) and from MD Ion stuffed simulation.

For the purpose of a mathematical model, starting from a predicted value of Sc(0) I have
proposed to introduce an adimensional factor to take into account this anomaly. The idea is to
calculate molar volume in CEAM conditions and calculate Sc(0) as follows:

B ⋅ V ⋅ E ⋅ co
SC (0) = − 0 <V <1 (2.87)
(1 −ν )

Factor “V” can be determined either experimentally and by MD calculations. Typically


V≈0.3-0.4. My proposal [87],[90] is to name this factor “Varshneya factor” in honour of
A.K.Varshneya who widely contributed to the discussion and solution of the LNDC (Cooper
coefficient) anomaly.

2.6 Characterization of ion exchanged glass

The effect of chemical strengthening by ion-exchange can be summarized as reported in


Figure 29 with two types of curves in the glass cross section:

49
Figure 29 – Stress and concentration profiles

- A concentration profile curve (green in Figure 29)


- A residual stress profile curve (red in Figure 29)

The main characteristics of the two curves to be considered for the characterization are
reported in table 6

Characteristics Description
Surface compression – SC (MPa) Compression stress value at the glass surface
Inner mid tension –ST (MPa) Tensile stress at the glass cross section mid
point
Case Depth – CD (µm) Depth of the compression layer (DOL) –
Distance between glass surface and internal
cross section point xCD at which compression
stress is 0 - σ(xCD) =0
Chemically modified layer depth -Cδ(µm) Depth of the chemically modified layer –
Distance between glass surface and internal
cross section point xCδ at which incoming ion
concentration is equal to value before in-
exchange.
Table 6 –Main characteristics of chemically strengthened glass by Ion-Exchange

When discussing characterization of chemically strengthened glass there are some preliminary
information that has always be provided:

Glass Chemical composition


Type of ion exchange process that has been performed

Glass chemical composition can be expressed both in weight percentages of glass constituents
(wt%) or in molar percentages (mol%) or molar fraction (see Appendix A8). The important
point is that the units of composition shall be clearly indicated. For ion exchange the molar
expressions are preferable over weight percentages.

For the process types we can identify some of the most significant:

- Single ion exchange – Glass single ion exchanged for single incoming ion. Example is
Sodium cation in glass exchanged for Potassium in molten salt bath :

50
[Na+]G ↔[K+]S

- Multiple Glass ion exchange – Multiple Glass ion exchanged for single incoming ion.
Example is Lithium and Sodium cations in glass exchanged for Potassium in molten
salt bath :
[Na+]G ↔[K+]S
[Li+]G ↔ [K+]S

- Multiple Bath ion exchange – Glass single ion exchanged for multiple incoming ion.
Example is Lithium cations in glass exchanged for Sodium and Potassium in molten
salt bath :

[Li+]G ↔[Na+]S
[Li+]G ↔ [K+]S

- Multiple Glass and Bath ion exchange – Multiple Glass Ion exchanged for multiple
incoming ion. Example is Lithium and sodium cations in glass exchanged for Sodium
and Potassium in molten salt bath :

[Li+]G ↔[Na+]S
[Nai+]G ↔ [K+]S

Obviously the kinetics of multiple ions ion exchange processes is far more complex than
single ion exchange processes. This is because there could be coupling mechanism and effects
both in the glass matrix and in the salt bath than can limit or speed up the process.
Nevertheless there could be even significant advantages in performing processes as the
“Multiple Glass and Bath ion exchange” to achieve high compression depth with the first
exchange and high residual stress level (Surface compression) with the second parallel
exchange process.

Together with these processes there can be other processes where two or more steps are
performed sequentially:

- Multiple sequential ion exchange processes. Where two or more of the above ion-
exchange processes are performed sequentially.
- Multiple hybrid or complex strengthening processes. Where an ion exchange process
is coupled to other strengthening processes. Examples Thermal strengthening
followed by Ion exchange (the reverse makes no sense because of the complete ion
exchange stress relaxation during the heating phase of thermal tempering. Ion-
exchange followed by crack healing/ bridging coating processes.

In most of these processes the important parameters to be evaluated are related to the residual
stress profile and they are: Surface compression – SC and depth of the compression layer or
case depth - CD. In some applications it is also important the compression value at a certain
depth from glass surface. The critical and finally important value is the strength of the glass
after the strengthening process (σb) compared to original glass strength (σ0) measured before
strengthening process.

51
2.6.1 Strength measurement by flexural tests

Even if methods have been discussed to predict glass strength the only way to asses strength
values for structural applications is by mechanical testing. There are basically two methods
normally used for glass, and in general for brittle materials.
The first is a flexural bending test where a bar of the material of thickness t is submitted to a
geometrically well defined bending test. In this test the edge of the specimen is under test, so
what is tested is both glass surface and edge finishing (grinding and polishing) level.
Normally failures starts from edge. The preferred test method is the so called four points
bending test where we have an area of constant bending moment between the two loading
rolls, in a three points bending test the maximum bending moment is in correspondence of the
single central loading roll.

Figure 30 - Four Points Bending Test

Figure 30 - Testing method: Four points bending test (ASTM C 158 on small specimens –
EN 1288-3:2000 on large specimens 1100mm x 360mm; L = 1000mm; L-2a= 200mm )

Definitions (see figure 30):

a - Moment arm (mm)


b - Bar width (mm)
t - Bar thickness (mm)
F - Force at breakage (N)
σb - Stress at breakage (Final stress) (N/mm2 = MPa)

3⋅ F ⋅ a
σb = (2.88)
b⋅t2

The method can also be applied to specimens with a circular cross section of radium r. In this
case we have:

52
3⋅ F ⋅ a
σb = (2.89)
π ⋅ r3
The stress at breakage is called by ASTM C 158 as M.O.R (Modulus of Rupture).

If (like in ASTM C 158):

a = 50 mm
b = 38 mm

we have:

3 ⋅ (50 )F F
σb = = 3,9474 ⋅ 2 (2.90)
38 ⋅ t 2
t

If (like in EN 1288-3)

a = 400 mm
b = 360 mm

3 ⋅ (400 )F F
σb = = 3,3333 ⋅ 2 (2.91)
360 ⋅ t 2
t

Another bending test method is based on biaxial flexure (ring-on-ring test). In this case
circular or squared samples supported by a circular rounded ring of radius r2 are submitted to
a flexural load by a loading concentric ring of radius r1. In this method the edge of the glass is
outside the stressed area and the obtained result is not influenced by edge finishing and is
representative of glass surface.
Testing method: Coaxial double ring test ⇒ Large surfaces EN 1288-3:2000
Small surfaces EN 1288-5:2000

Definitions:

r1 - Radius of the loading ring (mm)


r2 - Radius of the supporting ring (mm)
r3 - radius or characteristic dimension of the specimen
ν - Poisson ratio (0.22)
t - Specimen thickness (mm)
F - Force at breakage (N)
σb - Stress at breakage (Final stress) (N/mm2 = MPa)

3 ⋅ (1 + ν )   r2  1 − ν r2 2 − r1 2  F
σb = ⋅ ln  + ⋅ 2 
⋅ 2 (2.92)
2 ⋅π   r1  1 + ν 2 ⋅ r3  t

If we will have:

r1 = 6 mm
r2 = 30 mm
r3 = 33 mm ( for squared specimen this is half of the side length)

53
it results:

F
σ R = 1,09 ⋅ (2.93)
t2

for circular specimen and:

F
σ R = 1,04 ⋅ (2.94)
t2

for squared specimen (side L = 66 mm) using a value of r3 = 0.6 L.

In case deflection is significantly higher than half of sample thickness geometrical non linear
effects may be such relevant that the above formulas (2.92), (2.93) and (2.94) are no more
valid. In this last case (which is indeed pretty common with chemically strengthened glass)
the strength shall be evaluated from the force at breakage by using a non-linear Finite Element
model of the testing configuration. Neglecting this last point may results in significant
mistakes in strength determination.

Recalling discussions of the introduction, testing glass strength at room temperature in normal
laboratory environments (RH range between 25% to 80%) is like making a dynamic fatigue
test where surface flaws are stressed allowing their subcritical cracks to growth up to the
spontaneous propagation. For this reason results are very depending from the rate at which
load is applied (stress rate). Glass strength tests shall always have indications about test
environment conditions (temperature and humidity) and stress rate (MPa/s) applied. In glass
with a residual stress profile, load is applied up to the point where it exceed the residual
surface compression value and the mechanism of crack growth is activated.

It is clear that testing a number N of specimens in same testing conditions will result a
statistical distribution of values:

σ b (i ) ; i=1..N

The statistical distribution is not significantly due to small changes in testing conditions but is
mostly due to the surface flaws statistics. Herewith, we will introduce two methods for
statistical description of glass strength.

2.6.2 Statistical methods for glass strength

The characterization of glass strength is strongly influenced by the surface flaws statistics.
Another important factor influencing glass strength is edge finishing. This last factor is
relevant when edges are submitted to structural stress coming from load actions or thermal
actions.
To introduce statistical description of glass strength a short outline of statistics is introduced.

Let us consider a random variable x (related to breakage strength). If we introduce a


frequency distribution F(x) we can calculate the fraction of samples that would survive the
stress level x as (Survival probability):

54

S = ∫ F ( x) ⋅ dx (2.95)
x

If we introduce the Weibull distribution:

F ( x) = m( x) m−1 ⋅ exp − x m ( ) (2.96)

The survival S(σ) probability at stress level σ (x =σ/σ0 )will be:

 σ 
m

S = exp−    (2.97)
  σ 0  

This equation may also be written as:

1 σ
− ln(ln( )) = m ⋅ ln( ) = m ln(σ ) − m ⋅ ln(σ 0 ) (2.98)
S σ0

This is the equation of a straight line in logarithmic scale where m (called Weibull modulus)
is an indicator of statistical dispersion and σ0 is the stress level at which the survival
probability is equal to 1/e (0.37). Once m and σ0 are determined from a set of experimental
strength results the survival probability can be easily calculated at any stress level from
equation (2.97).
A method to calculate m and σ0 from a series of strength data can be set up as follows:

- Rank the N specimens in order of increasing strength (1,2,3..j..N); σ1,σ2,..σj..σN


- Determine the survival probability Sj of the jth specimen as:
j − 0,3
S j = 1− (2.99)
N + 0,4
- Plot the variable [ -ln(ln(1/Sj)] versus ln(σj)
- The least squares fit to the resulting line gives the Weibull modulus and the strength
value σ0.

In table 7 an example is reported of a typical distribution of strength data. Here data are
analyzed according to the above presented method. The Weibull plot is reported in figure 22.
The parameters of Weibull distribution obtained from data reduced in table 7 are.

m = 15,9
σ0 = 291 MPa

From these parameters, using equation (2.120), it is possible to calculate the survival
probability at any stress level . As an example consider a stress level of 200 MPa than the
survival probability results:

  σ m    200 15.9 
S = exp−    = exp−    = 0,9974 99,74 (%) (2.100)
  σ 0     291  

55
Rank-j Break Strength- Surv. Prob.-Sj ln(sj) ln(ln(Sj)
sj (MPa)
1 250 0,933 5,521461 2,663843
2 258 0,837 5,55296 1,723263
3 270 0,740 5,598422 1,202023
4 276 0,644 5,620401 0,821667
5 282 0,548 5,641907 0,508595
6 285 0,452 5,652489 0,230365
7 289 0,356 5,666427 -0,03292
8 298 0,260 5,697093 -0,29903
9 305 0,163 5,720312 -0,59398
10 310 0,067 5,736572 -0,99269
Table 7 – Examples of glass strength data reduction

From the statistical distribution also average and variance can be calculated:
 1
< S >= σ o ⋅ Γ 1 +  (2.101)
 m

  2  1 
Variance = σ o 2 ⋅  Γ 1 +  − Γ 2 1 +   (2.102)
  m  m 

In our case it results: <S>=282 MPa, Variance = 20.3 MPa

Weibull Plot y = -15,901x


2
+ 90,217
R = 0,9858
3

2,5

1,5
-lnln1/Sj

0,5

0
5,5 5,55 5,6 5,65 5,7 5,75
-0,5

-1

-1,5

ln sj

Figure 31 – Weibull plot of Table 7 data

The way to use this statistical approach is simple and straightforward. An acceptable survival
probability limit is fixed let us say for example S=99.999 % this means an acceptable
breakage level of (1-S) that is 1 over 100000 so that from equation (2.97)
  σ 15,9 
0.99999 = exp −    (2.103)
  291  

56
Solving for σ we get the allowable stress level as:

σ = 141MPa

Another alternative method sometime used for small samples series (below 20 samples) for
the statistical evaluation of glass strength is based on the concept of taking into account the
dispersion of data by a suitable factor that depends from the size of sampling.

N - Number of test specimens (at least 10)


σi - Breaking stress (MPa) for each test sample when tested according to EN 1288-3.
σav - Average value
1 N
σ av = ⋅ ∑ σ i (2.104)
N i =1
Sx - Standard deviation
N

∑ (σ − σ av )
2
i
(2.105)
SX = i =1

N −1

Vx - Coefficient of variation
S
VX = X (2.106)
σ av
Kn - Statistic coefficient corresponding to 95% confidence limit (This value depends from the
number of test specimens – N) according to the t-Student statistical distribution.
With the above definitions σC - Characteristic breaking strength is:

σ C = σ av ⋅ (1 − Kn ⋅ Vx ) (2.107)

Number of tested specimens Kn


10 2.26
11 2.23
12 2.20
13 2.18
14 2.16
15 2.14
20 2.09
25 2.06
30 2.05
40 2.02
60 2.00
120 1.98
∞ 1.96
Table 8 – Values of the statistical coefficient Kn vs. number of tested specimens

Taking the data of table 7 it results:

57
σ C = σ av ⋅ (1 − Kn ⋅ Vx ) = 282,3 ⋅ (1 − 2,26 ⋅ 0,068912) = 238,3MPa (2.108)

To have an acceptable value of design strength this characteristic breaking strength is divided
by a design factor. Very conservative approaches consider a factor equal to 4 so the
acceptable strength results:

σC 283,3
σ Design = = = 59,6MPa (2.109)
DF 4

That corresponds to a survival probability of 0,99999999999! This last approach is indeed


pretty conservative.
The choice of the design factor is in some way questionable. Less conservative approaches
use factors DF=2,5 in this last case :

σC 283,3
σ Design = = = 113MPa (2.110)
DF 2,5

That corresponds to a survival probability of 0,9999997 quite comparable with the result
achieved in (2.103). Providing that the samples tested are representative of the real conditions
of the loaded articles this last approach is probably more realistic than the one based on a
conservative design factor 4.

In our statistical description we have not mentioned the specimen size. Larger specimens will
have a higher probability of containing a deeper flaw which will cause lower strength. This is
an important point to be evaluated because data obtained from testing are coming from test
specimens usually smaller than final glass articles.
Let us follow this argument:

The survival probability of a sample of volume Vo to survive stress σ is:

  σ m 
S (VO ) = exp−    (2.111)
  σ 0  

The probability that a batch of n samples of the same volume Vo all survive stress σ is lower :

S (n ⋅ VO ) = [S (VO )]
n
(2.112)

Now, defining a larger volume V=n·Vo according to equation (2.111) it will be:

 V  σ 
m

S (V ) = [S (VO )] = [S (VO )]
V
= exp−   ⋅   
n
Vo
(2.113)
  VO  σ0  

There are some arguments to be discussed:

(i) In deriving equation (2.111) we have assumed that only one flaw population is
controlling strength. We may have different flaw populations coming from

58
different sources as processing (raw glass fabrication), machining (cutting,
grinding and polishing), further handling and so on. Different flaw populations
may have different strength distributions.
(ii) In deriving equation (2.113) we have assumed that volume flaw are responsible for
failure. For glass it is probably more realistic to consider that surface flaws cause
failure. In this case we can get the same equation using same arguments just
changing surface A in place of volume V:

  A  σ 
m

S ( A) = exp−   ⋅    (2.114)
  AO  σ 0  

Let us consider test specimens with a surface Atest and corresponding components (to be
designed with data taken from strength test on test specimens) with surface Acomp. Equating
survival probabilities of the two types of samples it results:

1
σ comp  Atest  m
=  (2.115)
σ test  Acomp 

Finally another argument to be considered is the correspondence between stress distribution in


a component and in a sample submitted to a three points or four points bending strength. In a
component fully submitted to tensile stress any critical flaw anywhere located in the sample
will propagate with equal probability. In the tested sample only a part of the sample is
submitted to the tensile stress. As a result the effective surface (volume) tested is reduced. A
good relationship between the strength obtained from test and the one effective on the
component is:

σ bendingtest
= [2 ⋅ (m + 1) 2 ]
1 m

σ component (2.116)

In other words the samples subjected to flexural bending test will appear to be stronger by a
factor depending from m. For m=5 the ratio is about 2, for m=20 the ratio reduces to 1.4.

When dealing with chemical strengthening an issue is the residual strength of the glass when
this is submitted to abrasion actions during service life. Because of the shallow depth of the
compression layer subsequent abrasion actions during service life may introduce deeper flaws
on the glass surface reducing its strength. This effect can be evaluated simulating the effect of
abrasion by external agents like sand, abrasive particles (silicon carbide) and than testing the
glass articles with the abraded surface in the tensile side of the test. Same mechanical test
arrangements can be used as described in 2.4.1.

59
3 – Product and process control

There can be considered several different methods to be used for product and process control
for chemically strengthened glass. The important point to have in mind is that a method
should provide reliable numbers. This because the results of the tests will be used to take
decisions about:

Products acceptance or rejection


Ordinary or extraordinary maintenance actions
Exceptional or emergency actions on the plant

In industrial production environments all these decisions may be very serious and also very
expensive, so it is of great importance that they are taken with reliable numbers on the table.

The fundamental technological parameters characterising a chemically strengthened glass are:

The final strength (Modulus Of Rupture MOR) σb (MPa). This parameter can be a
glass characteristic (without the edge) or a glass + edge characteristic.
Surface compression SC (MPa)
Depth of the compression layer CD (µm)

other related parameters are:

Central tensile stress introduced with the strengthening process - ST (MPa)


Depth of the ion exchanged layer Cδ (µm)
Strength increase - ∆σ = σb - σo (MPa) where σo is the original strength of the glass.

It is very important to recognise which parameter is measured, this will prevent misleading
conclusions and will help in the discussion of the results.

3.1 Chemical depth profiling

Chemical depth profiling methods measures directly either the introduced concentration
profile or the depth of the ion-exchanged (Chemically modified) layer. We will now introduce
the most popular methods:

Weight Gain

By recording the weight gain of a glass sample before and after ion exchange (∆W), it is
possible to calculate the total amount of moles (Q) of the introduced ions per unit surface area
of glass exposed to the ion exchange process (SIX). Knowing the molecular weight of the
incoming ions (MA) and of the outgoing ions (MB):

∆W
Q= (3.1)
S IX ⋅ ( M A − M B )

this relationship will allow the determination of the Mass Average Diffusion Coefficient
DM(cm2/s) integrating equation (2.20)a and assuming constant mean interdiffusion
coefficient:

60
DM ⋅ t
Q = 2 ⋅ cs ⋅ (3.2)
π

where t is the process time. A similar relationships (Qa= f[(DMot)1/2] holds:

Q = f ( DM 0 ⋅ t ) 
1/2
(3.3)
 

even if the diffusion coefficient is a function of concentration, in that case DMo is the
diffusion coefficient at the glass surface. So by using the value of the mass average diffusion
coefficient it is possible to reconstruct the concentration depth profile and evaluate the depth
of the exchanged layer.

Sectioning and chemical analysis

The concentration depth profile can be measured sectioning the glass samples submitted to
ion exchange. Thin successive layers of glass can be removed by 2-5% solutions of
hydrofluoric acid (HF). The depth of the layer removed can be controlled easily with the
accuracy of 1 µm by weigh control. Analysis of the etchates by, AAS or ICP spectrometry
provides the chemical composition of the removed layer. This is the basis of a method
recently proposed by GLAFO-RISE (SAC) [65],[80]. If the bath is tagged with radiotracer of
the exchanging specie than counting the etchates of remaining sample will give information
about the concentration profile.

Electron microprobe

In this method the concentration is measured using an electron microscope equipped with an
X-rays microprobe. Positioning the beam in a defined point of the glass sample section and
focusing this point with an electron beam generates x-rays emission that can be detected by a
dispersive microprobe giving qualitative and quantitative concentration information of the
exchanging and exchanged ions.

In this method care has to be taken to minimise alkali migration due to electron beam local
heating and charging, neglecting this point may cause serious errors. Also inaccuracy in
establishing a sharp boundary at the glass surface with fixing and mounting materials can
cause other accuracy problems. A good advice is to scan the beam in the diffusing direction to
minimize artifacts. Sometime this method can lead to controversial results when used in
inappropriate conditions.

3.2 Strength measurements

MOR strength measurements can be performed mechanically using dynamometers equipped


with load cells. Methods have been developed both for bar tests in which glass and edge are
part of the loaded section and double coaxial bending test in which the edge effect is excluded
as discussed in point 2.6.1.

61
The strength increase evaluation from initial strength and final strength should be made using
the same method (bar test or double coaxial bending test) on original glass samples and
strengthened samples.

The bar test should be considered when the edge has some structural importance in the glass
article application.

It should be observed that the mechanical MOR tests are the only methods to assess the
technological quality of the strengthened glass. In coaxial double ring test a special attention
shall be paid to the limit of linearity conditions. When deflection is of the order of magnitude
of sample thickness the test conditions are outside of the linearity limits and strength cannot
be evaluated using the “linear” formulas, in this case a FEA model of the test is strongly
recommended for data reduction. Abrasion procedures before testing (ASTM C 158 or others)
can be developed to evaluate the effect of surface abrasion or other tribology effects
(scratching) on the residual strength.

3.3 Stress depth profiling

These methods have the purpose to measure the surface residual compression SC or/and the
depth of the compression layer (case depth) CD. We can define optical methods based on
photoelasticity and other mechanical methods based on microindentation techniques.

Optical methods

Different instruments have been developed to measure the surface compression resulting from
chemical strengthening. The point is that stressed glass became birefringent and under certain
illumination conditions with polarised light, interference fringes result related to the level and
distribution of residual stresses.

The analysis of the fringes gives information on the amount of the residual compression stress
on the glass surface and/or the refractive index and as a consequence stress profile.

Direct methods to measure stress profile and namely surface compression and case depth is by
using polarising microscopes with compensators (Berek or others). A detailed procedure is
established in a recent ASTM standard –ASTM 1422.

Other popular instruments are:

DSR - Differential surface refractometer. A version based on a paper of Kishi [31] is actually
produced by Luceo (Japan) with the brand name FSM . This can be used both for surface
compression and stress profile measurement.

Stratorefractometer - Developed by St.Gobain Research is basically the same as the Luceo


DSR.

Babinet compensator. While DSR and Stratorefractometer can be used on samples and glass
articles, Babinet compensator requires a special produced sample very well polished. It works
looking perpendicular to the ion exchanged section and illuminating from the other side with
polarised light.

62
Another class of methods are based on the use of scattered light (SCALP 5 by GlassStress in
Estonia).

Mechanical methods

If the glass surface is submitted to a Vickers Pyramidal microindentation test with a constant
load the length of the radial cracks can be related to the residual surface compression. If the
test is performed etching off subsequent layers a µ-Hardness profile can be generated.

3.3 Process Design and Process Control

From what said above it is pretty clear that a careful control of process parameters is of
paramount importance to design process condition according to the product specification.
This discussion identify the product specifications as driving parameters in process design.
The product specifications should define the following technological parameters:

Strength or Modulus of Rupture (MOR) - σb (MPa). This parameter can be measured


according to a mechanical test. It should be specified if (bar test) the edge is a part of
the tested cross section (4 points flexural bar test) or if the edge is outside the loaded
section(coaxial ring-on-ring test).
Abraded MOR - With this test the samples are intentionally and artificially abraded to
simulate the effect of the service conditions on the strength MOR. Surface abrasion or
scratching conditions shall be expressed in a detailed way.
Surface compression. This parameter is of paramount importance for both guarantee
final strength and to assess production conditions.
Depth of the compression layer or case depth CD. This parameter is of great
importance for the service life of the strengthened articles.

Other parameters (see figure 23) can be defined to characterize products :

Depth of the ion exchanged (Chemically modified) layer Cδ and the concentration
depth profile of the exchanged ions C = C(x).
Central tension ST value
The µ-Hardness value (or its profile after subsequent etching)

For glass product service life interesting results can be deduced comparing the strength
increase as a function of depth of compression layer in case of abraded and not abraded
glasses. In Figure 32 relevant results are reported for soda-lime glasses at different process
temperatures.

The most evident result is that, to retain the original strength of the chemically strengthened
glass, it is needed a depth of compression layer of at least 35 µm .This result is consistent with
the measurements of “ordinary scratches” induced by service life on glass articles. This means
that strengthening processes of few microns of depth, on weak (poor original strength) glass is
practically not useful for most of the applications.

63
Strength increase ∆σ (MPa) as a function of Case depth CD(µm)
Process temperatures : 475°C - Blue line / 450°C - Red line / 400°C Black line
Dotted lines - Without abrasion / Solid line - After abrasion
400

350
Strength increase - ∆σ(MPa)
300

400°C
250
450°C

200 475°C

150

100

50

0
0 10 20 30 40 50 60

Compression Layer Depth - CD(µm)

Figure 32 – Strength increase ∆σ as a function of Compression layer depth CD for different


process temperatures and with abrasion effects.

The identification of the product specifications defines a suitable process that is defines main
process parameters:

Ion source characteristics


Temperature - TP
Process time - τ.

This is the process design in which process parameters have to be selected to get products
with the specified characteristics.

The variables to be considered in process design are:

Glass Chemical composition, and glass original strength σo


Temperature dependence of the interdiffusion coefficient DAB(T) and Temperature and
process time dependence of the stress relaxation σS(T,τ).

Salt quality (basically their contamination) is something related primarily to surface


compression and (when it become serious) on case depth.
Process control methods should be economical, simple and straightforward enough to allow
quick evaluation. A possible way to establish a simple process control can be based on the
measurement of at least two variables:

A parameter related to ion-exchange efficiency that should be weight gain or case


depth CD.
A parameter related to the strengthening effect that is: surface compression SC or
central tension ST (I personally prefer Surface compression).

64
For strength or residual strength after abrasion only direct methods based on mechanical
tests can be used.

3.4 Interdiffusion coefficient definition

If we know or we establish an original glass strength we can also define a characteristic flaw
depth c (µm) this means that we can calculate the strength increase according to the methods
outlined in section 2. The initial step of all calculation methods is the residual concentration
determination. This can be performed in a pure theoretical way by knowing the interdiffusion
coefficient.

Circles - Soda lime glass type E


Squares - Soda lime glass type F
1E-10
Mass Average Diffusion Coefficient - DM (cm /s)
2

1E-11

1E-12
400 425 450 475 500
Process Temperature - T (°C)
IONICS LAB.

Figure 33 – Mass Average Diffusion coefficient DM (cm2/s) versus process temperature for
two different types of soda-lime glasses.

As already pointed out the dependence of the diffusion coefficient versus process temperature
is of “Arrhenius type” typical of thermally activated processes with an activation energy H0
for the diffusion process (see Figure 33):

 H0 
D = D0 ⋅ exp −  (3.14)
 R ⋅ TP 

Activation energy for Soda-Lime glass is typically ranging between 120 to 170 kJ/mol
depending on the glass chemical composition and types of ions to be exchanged.
Another way (to be preferred in scientific reporting) to plot D vs TP can be in logarithmic
scale: Log [DM] but versus 1000*(1/TP) (see figure 34).

65
Arrhenius Plot of The Mass Average Diffusion Coeff.DM
for two types of SL glasses

-9,5
1,32 1,34 1,36 1,38 1,4 1,42 1,44
-10
Log[DM(T)]

-10,5

-11

-11,5

-12
[1000*1/T(°K)]

Figure 34 Arrhenius plot of D vs TP for two SLS glasses

3.5 Reference Values

Design procedures require reference values, these values are normally referred to glass
chemical composition and to the salt bath conditions. This is the reason why a careful process
control has to be established to be able to compare continuously running production to
reference values. In the next section we will discuss the process control issue.

Reference values should be given both for ion-exchange kinetics parameters:

Mass average diffusion coefficient – DM = D (TP) or depth of ion exchanged layer Cδ


or weight gain

and strength parameters:

Surface compression SC or Residual stress profile - σ = σ (x)

Also the original glass strength should be defined because this parameter characterise the
typical flaw length and this is of critical importance in evaluating the strength increase after
the strengthening process. Original glass strength may be also related to surface quality and
edge finishing, as a general statement no surface visible scratches are allowed .
The two soda-lime glass types mentioned in figure 33 and a Lithium Alumino Silicate LAS
glass mentioned have the following chemical composition :

66
Glass Chemical SL-F SL-E LAS
composition by Weight (%) Weight (%) Weight(%)
Fluorescence -
X
SiO2 73,0 71,5 67,2
Al2O3 0,47 1,20 20,1
Na2O 13,53 14,20 0,35
K2O 0,14 0,70 0,23
CaO 8,60 7,40 0,05
MgO 3,95 4,50 1,06
SO3 0,25 0,22 /
As2O3 / 0,15 0,87
Fe2O3 0,08 0,02 0,03
TiO2 0,048 0,02 2,67
ZrO2 / / 1,73
ZnO / / 1,65
Li2O / / 3,15
Table 9 - Chemical Composition (by % weight) of the two types of glass

and Data reported in Figures 32 and 33 are coming from measurements reported in table 10.

Mass Average Diffusion


Coefficient DM (cm2/s)

Process Glass type Glass type


Temperature SL-F SL-E
(°C)
425 4,0 * 10-12 9,7 * 10-12
450 9,4 * 10-12 2,2 * 10-11
-11
475 2,3 * 10 5,1 * 10-11
Table 10 -Mass Average diffusion Coefficient for two types of soda-lime glasses

Other available data for Soda-Lime silicate (SLS) float glass and for Lithium
AlluminoSilicate (LAS) glasses are reported in Table 11 and in Table 12 respectively.

Process Temperature Mass Average Diffusion Coefficient


T(°C) DM (cm2/s)
364 5.2 *10-13
395 1.0 *10-12
425 5.0 *10-12
455 8.8 *10-12
476 2.2 *10-11
504 6.3 *10-11
Table 11 - Mass average Diffusion Coefficient at different process temperature:
Glass: Soda-Lime Salt Bath: KNO3 Pure, Analytical grade

67
Process Temperature Mass Average Diffusion Coefficient
T(°C) DM (cm2/s)
380 3.4 *10-9
400 5.7 *10-9
430 9.8 *10-9
Table 12 - Mass average Diffusion Coefficient at different process temperature:
Glass: Lithium Allumino Silicate (LAS) Salt Bath: NaNO3 Pure, Analytical grade

Mass Average Diffusion Coefficient

-10
1,2 1,3 1,4 1,5 1,6
-10,5
Log (DM(cm/s2)

-11

-11,5

-12

-12,5

-13
[1/T(K)]*1000

Figure 35 - Mass average Diffusion Coefficient at different process temperature (Arrhenius


plot): Glass: Soda-Lime - Salt Bath: KNO3 Pure, Analytical grade

68
Arrhenius Plot of The Mass Average Diffusion Coeff.DM
of LAS glass

-7,6 1,4 1,42 1,44 1,46 1,48 1,5 1,52 1,54

-7,8

-8
Log[DM(T)]

-8,2

-8,4

-8,6

-8,8

-9
[1000*1/T(°K)]

Figure 36 - Mass average Diffusion Coefficient at different process temperature (Arrhenius


plot): Glass: LAS – Lithium Allumino Silicate - Salt Bath: NaNO3 Pure, Analytical grade

69
Arrhenius Plot of The Mass Average Diffusion Coeff.DM
for SL glass (Blue line - Ion Exchanged with KNO3) and
LAS glass (Red line -Ion Exchanged with NaNO3)

-7
1,3 1,35 1,4 1,45 1,5 1,55 1,6
-8

-9
Log[DM(T)]

-10

-11

-12

-13

[1000*1/T(°K)]

Figure 37 – Comparison of Mass average Diffusion Coefficient at different process


temperature (Arrhenius plot)for SLS and LAS glasses

3.4.1 Process Control Procedures

It is important to stress, once again, the critical role of process control. Design data have a
meaning only if they can be related with a certain grade of confidence to process condition.
The problems that can reduce the effectiveness of chemical strengthening may be addressed
to:

Poor quality (grinding/polishing defects or scratches) of edge finishing of the original


glass
Contamination of the salt bath
Poor control in Process parameters (bad temperature control and uniformity)

Process control system can be of two types:

Test on samples introduced at each dip (or at several dips)


On line test on the salt batch

test on samples can be:

70
Mechanical MOR tests (bar tests ASTM C 158 or/and ring tests EN 1288-5)
Depth of the ion-exchanged layer (compression layer) by weigh gain or by
Electron Micro-Probe Analysis or by optical methods (DSR-
Stratorefractometer)
Surface residual stress or stress profile by optical methods (DSR-
Stratorefractometer)

Tests on the batch are:

Chemical analysis of the salt batch , pH, chemical composition and mainly:
Na+ molar fraction ([Na+]/([K+]+[Na+]) - that should always stay below 1 %
(better if 0.5 %); concentration of divalent ions: Ca++ , Mg++, Pb++…., that
should stay below 10-20 ppm.
Electrochemical potentiometric analysis of the salt batch.

Test on samples should be performed at each process dip/run and carefully recorded a data
base collecting the plant history is strongly recommended, basic information to be collected
are:

Process identification data (Glass type, glass articles, process time, process
temperatures)
Test results on samples and, if existing, on the salt bath.

71
Bibliography
Books
[B1] Varshneya, A.K.: Fundamentals of Inorganic Glasses, 2nd Ed. Sheffield UK: Society of
Glass Technologies; 2006.
[B2]. Le Bourhis, E.: Glass: Mechanics and Technology, 2nd Ed. Weinheim Germany: Wiley
VCH Verlag, 2014.
[B3] Green, D.J.: An introduction to the mechanical properties of ceramic , Cambridge UK,
Cambridge University Press, 1998
[B4]. Aben, H., Guillemet, C.: Photoelasticity of Glass, Berlin Heidelberg: Springler-Verlag,
1993.
[B5] Barton, J., Guillemet,C.: Le verre, science et technologie, Les Ulis France, EDP
Sciences, 2005
[B6] Doremus, R.H.: Glass Science, 2nd Ed. New York – USA, John Wiley & Sons, Inc, 1994
[B7] Lawn, B.: Fracture of brittle solids, 2nd Ed. Cambridge UK, Cambridge University Press,
1993
[B8] Barsoum, M.W. : Fundamentals of ceramics, 2nd Ed. Bristol UK, IOP Publishing Ltd,
2003
[B9] Elasticity and strength in Glasses in: Glass: Science and Technology Edited by
Uhlmann, D.R. and Kreidl, N.J., New York USA, Academic Press Inc, 1980
[B10] Strength of inorganic glass, Edited by Kurkjian C.R., New York USA, Plenum Press
1985
[B11] Frischat, G.H. Ionic diffusion in Oxide Glasses, Aedermannsdorf Switzerland,
TransTech Publications S.A. 1975
[B12] Wachtman, J.B.,Cannon, W.R.,Mattewson, M.J: Mechanical properties of Ceramics,
2nd Ed,New York USA, John Wiley & Sons, 2009.
[B13] Munz,D., Fett,T.; Ceramics: Mechanical properties, Failure Behaviour, Materials
Selection, Berlin Heidelberg DE, Springer-Verlag, 1999
[B14] Scholze, H.: Glass – Nature, Structure and Properties,1990 Berlin-Heidelberg DE,
Springer-Verlag, 1990 (English translation)
[B15] Vogel,W : Chemistry of glass, Westerville OH, The American Ceramic Society, 1985
(English translation)
[B16] Crank,J.: The Mathematics of Diffusion, 2nd Edition, Oxford UK, Oxford University
Press 1975] .
[B17] Carslaw,H.S.,Jaeger,J.C.: Conduction of heat in solids, 2nd Edition, Oxford UK, Oxford
University Press 1959
[B18] Jost,W: Diffusion in solids, liquids, gases, 3rdPrinting, New York USA, Academic
Press Inc. 1960
[B19] Shelby, J.E. : Introduction to glass science and technology, 2nd Edition, Cambridge UK,
The Royal Society of Chemistri (RSC) , 2005

72
Articles

[1] Doremus, R.H. : Exchange and diffusion of ions in glass, The Journal of Physical
Chemistry, 1962 Vol 8, No 08 pp 2212-2218
[2] S. S. Kistler, "Stresses in Glass Produced by Nonuniform Exchange of Monovalent Ions",
J. Am. Ceram. Soc. 45(2), 59-68 (1962).
[3] P. Acloque and J. Tochon, "Measurement of Mechanical Resistance of Glass After
Reinforcement", in Union Scientifique Continentale du Verre, p 1044, Charlroi, Belgium,
1962.
[4] H. M. Garfinkel, D. L. Rothermel, and S. D. Stookey, “Strengthening by Ion Exchange”,
in Adv. in Glass Tech., Proc. VI Intl Cong on Glass, Plenum Press 1962.
[5] M. E. Nordberg, E. L. Mochel, H. M. Garfinkel, and J. S. Olcott, "Strengthening by Ion
Exchange", J. Am. Ceram. Soc., 47(6) 215-219 (1964).
[6] A. J. Burggraaf and J. Cornelissen, “The Strengthening of Glass by Ion exchange: Part1”,
Phys. Chem. Glasses, 5(5) 123-129 (1964).
[7] R. W. Ansvin, " The Nondestructive Measurement of Surface Stresses in Glass", ISA
Trans., 4, 339 -43 (1965).
[8] S. Bateson, J. Hunt, D. A. Dalby and N. K. Sinha, "Stress Measurements in Tempered
Glass by Scattered Light Method with a Laser Source", Bull. Am. Ceram. Soc. 45(2), 193-98
(1966).
[9] A. J. Burggraaf and J. Cornelissen, “The Strengthening of Glass by Ion exchange: Part2”,
Phys. Chem. Glasses, 7(5) 169-172 (1966).
[10] W. J. Spoor and A. J. Burggraaf, “The Strengthening of Glass by Ion exchange: Part3”,
Phys. Chem. Glasses, 7(5) 173-177 (1966).
[11] M. J. Kerper and T. G. Scuderi, “Mechanical properties of Chemically Strengthened
Glasses at High Temperatures”, J. Am. Ceram. Soc., 49(11), 613-618 (1966).
[12] H. M. Garfinkel, “The Thermal Fatigue of Glasses and Glass-ceramics Strengthened by
Ion Exchange”, In Proceedings of Symposium sur la Surface du Verre et ses Traitements
Modernes, Luxembourg, pp 57-71, June 1967,
[13] Hale, D.K. Strengthening of silicate glasses by ion exchange. Nature.1968;217:1115-
1118.
[14] Cooper, A.R., Krohn, D.A. Ion Exchange strengthening of glass fibers II.J Am Ceram
Soc.1969;52:665-669.
[15] H. M. Garfinkel and C. B. King, “Ion Concentration and Stress in Chemically Tempered
Glass”, J. Am. Ceram. Soc. 53(12) 686-691 (1970).
[16] T. R. Kozlowski and G. A. Chase, “Parameters of Chemical Strengthening and Impact
Performance of Corning Code 8361 (White Crown) and Corning Code 8097 (Photogray)
Lenses”, Am. J. Optometry, 50(4) 79-88 (1973).
[17] A. K. Varshneya and M. E. Milberg, "Ion Exchange in Sodium Borosilicate Glass", J.
Amer. Ceram. Soc., 57(4), 165-69 (1974).
[18] Schaeffer, H.A., Heinze, R. Stress build up by ion exchange in glass. Glastech
Berichte.1974;47:199-207.
[19] H. Ohta, "The Strengthening of Mixed Alkali Glass by Ion Exchange", Glass Technol.,
16(1) 25- 29 (1975).
[20] A. K. Varshneya, “Kinetics of Ion Exchange in Glasses”, J. Non-Cryst. Sol. 19, 355-365
(1975).
[21] A. K. Varshneya, "Influence of Strain Energy on Kinetics of Ion Exchange in Glass", J.
Am. Ceram. Soc., 58(3-4), 106-109 (1975).
[22] A. K. Varshneya and R. J. Petti, "Finite Element Analysis of Stresses in Ion-Exchanged
Glass", J. Am. Ceram. Soc., 59(1-2), 42-46 (1976).

73
[23] J. R. Varner and R. Lang-Egelkraut, in Non-Crystalline Solids, pp 465-470, Trans Tech
Publications, Aedermannsdorf, Switzerland, 1977. (Proc. Fourth Intl. Conf. on the Physics of
Non-Crystalline Solids).
[24] A. Y. Sane and A. R. Cooper, "Anomalous Stress Profiles in Ion-Exchanged Glass", J.
Am. Ceram. Soc., 61(7), 359- 362 (1978).
[25] M. Abou-el-Leil and A. R. Cooper, “Analysis of Field-Assisted Binary Ion Exchange”, J.
Am. Ceram. Soc., 62(7-8) 390-395 (1979).
[26] W. Bradshaw, “Stress Profile Determination in Chemically Strengthened Glass Using
Scattered Light”, J. Mater. Sci., 14, 2981-2988 (1979).
[27] R. F. Bartholomew and H. M. Garfinkel, "Chemical Strengthening of Glass", in Glass
Science and Technology, v.5 (Elasticity and Strength in Glasses), pp.217-270, Academic
Press, New York, 1980.
[28] M. Abou El-Leil, “Abrasion-Resistant High Strength (1000 MPa) Glass, Glass Technol.,
21(4) 206-208 (1980)
[29] C. Windisch and W. M. Risen, Jr., “Comparative Raman Study of Mixed Alkali and
Silimar Ion-Exchanged Glasses”, J. Non-Cryst. Sol. 44, 345-364 (1981).
[30] H. Ono, “The Production of Chemically-Strengthened Glass Containers”, Glass 60(10)
380-383 (1983).
[31] T.Kishi, “ Surface stress meters utilizing the optical waveguide effect of chemically
tempered glasses”, Optics and Lasers in Engineering, 4 (1983) 25-38
[32] A. K. Varshneya and G. Dumais, "Influence of Externally Applied Stress on Kinetics of
Ion Exchange in Glass", J. Am. Ceram. Soc., 68(7),C165- 66 (1985).
[33] Y. Jiang and L. Jiang, “Effect of Additives in the Salt Bath on Glass Strengthening”, J.
Non-Cryst. Sol. 80, 300-306 (1986).
[34] X. Zhang, O. He and Y. Zheng, “The Effect of Impurity Ions in Molten KNO3 salt on
Ion Exchange and Strengthening of Glass, “J. Non-Cryst. Sol., 80, 313-318 (1986).
[35] L. B. Glebov, V. G. Dokuchaev, N. B. Nikonorov, and G. T. Petrovskii, “Change in the
Volume of Glass After Low-Temperature Ion Echange”, Fiz. i Khim Stekla, Translation, 14(2)
pp 232-239, March-April 1988. Original article: December 1986.
[36] S. Carson and W. C. LaCourse, “Effect of Surface Flaws on Ion Exchange
Strengthening”, Proc. XIVth Intl. Cong. On Glass, New Delhi (India), 2, 71-77 (1986).
[37] M. Abou-El-Leil and A. R. Cooper, “Impact Fracture of Chemically Tempered Glass
Helicopter Windshields”, J. Am. Ceram. Soc., 69(9) 713-716 (1986).
[38] V. Jain and A. K. Varshneya, “Finite Element Analysis of Network Dilatation in Ion-
Exchanged Glass After Slicing”, J. Am. Ceram. Soc., 70(8) 585-598 (1987).
[39] A. Y. Sane and A. R. Cooper, "Stress Buildup and Relaxation During Ion Exchange
Strengthening of Glass", J. Am. Ceram. Soc., 70(2), 86-89 (1987).
[40] W. C. LaCourse, "How Surface Flaws Affect Glass Strength", Glass Ind.,14- 23, June
1987.
[41] I. W. Donald and M. J. C. Hill, “Preparation and Mechanical Behavior of Some
Chemically Strengthened Lithium Magnesium Alumino-Silicate Glasses”, J. Mater. Sci., 23,
2797-2809 (1988).
[42] D. Connoly, “Fracture Analysis of Chemically Strengthened Glass Disks”, J. Am. Ceram.
Soc. 72(7), 1162-1166 (1989).
[43] Albert, J., Lit, J.W.Y.: Full modeling of field assisted ion exchange for graded index
buried channel optical wavwguides. Appl. Optics 1990 vol.29 No. 18 pp 2798-2804
[44] H. Wang, W. C. LaCourse and P. F. Johnson, “Sonic Strengthening Effect on Soda-
Lime-Silicate Glass”, Proc. 1991 Dalian Conf. On Glass, p.226, Beijing (China).
[45] A. K. Varshneya and W. C. LaCourse, "Technology of Ion Exchange Strengthening of
Glass: A Review”, Ceram. Trans. 29, pp365-376 (1993).

74
[46] I. W. Donald, M. J. C. Hill, B. L. Metcalfe, D. J. Bradley and A. D. Bye, “The
Mechanical Properties and Fracture behavior of Some Chemically Strengthened Silicate and
Borate Glasses”, Glass Technol. 34(3) 114-119(1993).
[47] S. T. Gulati, “Frangibility of the Tempered Soda-Lime Glass Sheet”, Proc. of Glass
Processing Days, 13-15 Sept 1997, Tampere (Finland).
[48] Y-K Lee, Y. L. Peng, and M. Tomozawa, “IR Reflection Spectroscopy of a Soda-Lime
Glass Surface During Ion Exchange”, J. Non-Cryst. Sol., 222, 125-130 (1997).
[49] V. Tyagi and A. K. Varshneya,”Measurement of Progressive Stress Buildup During Ion
Exchange in Alkali Aluminosilicate Glass” J. Non-Cryst. Sol. 238, 186-192 (1998).
[50] D. J. Green, R. Tandon and V. M. Sglavo, “Crack Arrest and Multiple Cracking Through
the Use of Designated Residual Stress Profiles”, Science, 283(5406) 1295-1297 (1999).
[51] ASTM C-1422—99, “Standard Specification for Chemically Strengthened Flat Glass”,
Amer. Soc for Testing Materials, W. Conshohocken PA.
[52] S. J. Glass, M. Abrams, and R. V. Matalucci, ”New Glass Technologies for Enhanced
Architectural Surety: Engineered Stress Profile (ESP) for Soda-Lime-Silica Glass”, Sandia
Research Laboratories, Report SAND2000-3001.
[53] A. K. Varshneya, “Ion exchange: Physical Properties of Ion-Exchanged and Melt-
Processed Glasses Differ”, GlassResearcher, 10(20/11(1), 21-26, 51 (2001).
[54] G. Macrelli, “Strength Issues in Chemically Strengthened Glass”, Riv. della Staz. Sper.
del Vetro no. 4, 69-76 (2001).
[55] C. W. Sinton and W. C. LaCourse, “Experimental Survey of the Chemical Durability of
Commercial Soda-Lime-Silicate Glasses”, Mat. Res. Bull., 36, 2471-2479 (2001).
[56] J. Shen and D. J. Green, “Variable Temperature Ion-Exchanged Engineered Stress
Profile (ESP) Glas”s, J. Am. Ceram. Soc., 86(11) 1979-1981(2003).
[57] J. Shen, D. J. Green, R. E. Tressler, and D. L. Shellman, “Stress Relaxation of a Soda
Lime Silicate Glass Below the Glass Transition Temperature”, J. Non-Cryst. Sol., 324, 277-
288 (2003).
[58] J. Shen and D. J. Green, “Prediction of Stress Profiles in Ion Exchanged Glasses”, J.
Non-Cryst. Sol., 344, 79-87 (2004).
[59] Jill Glass, Raj Tandon, Arun Varshneya and Ian Spinelli, “Processing and Properties of
Ion Exchanged Glases”, PowerPoint presentation, Glass & Optical Materials Division
Meeting, Amer. Ceram. Soc. , Cape Canaveral FL, November 6-12, 2004.
[60] J. Shen and D. J. Green, “Effect of the K/Na ratio in Mixed-Alkali Lime Silicate Glasses
on the Rheological and Physical Properties, “J. Non-Cryst. Sol., 344, 66-72 (2004).
[61] R. Tandon and S. J. Glass, “Controlling the Fragmentation Behavior of Stressed Glass”,
in Fract. Mechan. Ceram. R. C. Bradt, D. Munz, M. Sakai, and K. W. White editors, v. 14 p.
77-92 (2005).
[62] Wondraczek, L., Dittmar, A. , Oelgardt,C. , Célarié,F. , Ciccotti,M. Marlière,C. , J. Am.
Ceram. Soc. 2006 , 89 , 746 .
[63] V. M. Sglavo, A. Prezzi and D. A. Green, “In Situ Observation of Crack Propagation in
ESP (Engineered) Stress Profile) Glass”, Eng. Fract. Mechan. 74, 1383-1398 (2007).
[64] J. W. Smail, D. J. Green, and M. B. Abrams, “Onset of Cone Cracking in Ion-Exchanged
Glass”, J. Am. Ceram. Soc., 90(1) 333-335 (2007).
[65] A. Quaranta, E. Cattaruzza, F. Gonella, “Modelling the Ion exchange Process in Glass:
Phenomenological Approaches and Perspectives”, Mater. Sci. & Eng. B, 149, 133-139 (2008).
[66]. Gy, R. Ion exchange for glass strengthening. Mater. Sci. Eng. B, 2008;149: 159-165,.
[67] Mauro,J.C., Loucks,R.J.; J.Non-Cryst.Solids, 355 (2009) 676
[68]. Karlsson, S., Jonson, B. The technology of chemical glass strengthening – a review.
Glass Technol. Eur. J. Glass Sci. and Technol.2010;51:41-54.

75
[69]. Varshneya, A.K. Chemical strengthening of glass: lessons learned and yet to be learned.
Int. J. Appl. Glass Sci.2010;1:131-142.
[70]. Varshneya, A.K. The physics of chemical strengthening of glass: room for a new view.
J. Non-Cryst. Solids.2010;356:2289-2294.
[71] Kreski,P.K.,Varshneya,A.K.,Cormack,A.N. Investigation of ion exchange “stuffed” glass
structure by molecular dynamicssimulation, J. Non-Cryst.Solids., 358(2012)3539-3545
[72] Tandia, A.,Vargheese,K.D.,Mauro J.C.,Varshneya,A.K.: Atomistic understanding of the
network dilation anomaly in ion exchanged glass, J. Non-Cryst.Solids., 358(2012)316-320
[73] Tandia,A.,Vargheese,K.D.,Mauro,J.C,: Elasticity of ion stuffing in chemically
strengthened glass, J. Non-Cryst.Solids., 358(2012)1569-1574
[74] Fu,A.I.,Mauro,J.C. Mutual diffusivity, network dilation and salt bath poisoning effects in
ion exchanged glass, J. Non-Cryst.Solids., 363(2013)199-204
[75] Vargheese, K.D.,Tandia,A.,Mauro,J.C. Molecular dynamic simulations of ion exchanged
glass, J. Non-Cryst.Solids., 403(2014) 107-112
[76] Mazzoldi, P. Carturan, S., Quaranta, A.,Sada,C.,Sglavo, V.M.: Ion Exchange Process:
History, evolution and applications, Riv. Nuovo Cimento, 2013 Vol 36, No9 pp 397-460
[77]. Macrelli, G:,Poli, E. Ion concentration and stress profile modifications of ion exchanged
glass after thermal treatment. In: Materials Science and Technology (MS&T) October 27-31,
2013 Montreal, Quebec, Canada 2013:2434-2442.
[78] Sglavo,V.M., Quaranta, A.,Allodi,V.,Mariotto, G. Analysis of the surface structure of
soda lime silicate glass after chemical strengthening in different KNO3 salt baths, J. Non-
Cryst.Solids., 401(2014)105-109
[79]. Macrelli, G.,Poli, E. Chemically strengthened glass by ion exchange: residual stress
profile and strength evaluation. In: Engineered Transparency. International conference at
glasstec, Dusseldorf Germany 21-22 October 2014:231-240.
[80] Macrelli, G.:Glass chemical strengthening by ion exchange: from ion exchange kinetics
to strength determination. In Proceedings of Glass Performance Days 2015 406-410 –
Tampere June 2015
[81]. Morozumi, H., Nakano, H, Yoshida, S, Matsuoka, J. Crack initiation Tendency of
Chemically Strengthened Glasses.Intl. J. Appl. Glass Sci.2015;6:64-71
[82] Rogozinski, R Ion exchange in glass – the changes of glass refraction, INTECH – doi.
10.5772/51427
[83] Wang,M.,Bauchy,M.: Ion exchange strengthening of glass: atomic topology matters,
arXiv:1505.08880v1[cond-mat.mtrl-sci] 28 May 2015
[84] Karlsson,S.,Ali, S.,Limbach,R.,Strand,M. Wondraczeck, L. : Alkali salt vapour
deposition and in-line ion exchange on flat glass surfaces, Glass Technol:Eur.J.Glass
Sci.Technol.A, 2015, 56(6) 203-213
[85] Macrelli,G : Critical evaluation of chemically strengthened glass in structural glazing
applications. In Engineered Transparency 2016. Glass in Architecture and Structural
Engineering. First Edition Edited by Jens Schneider, Bernhard Weller, Ernst & Sohn Gmbh &
Co. KG. 2016
[86] Sglavo, V.M.;Effects of salt impurities on chemical strengthening of float glass by ion
exchange. Presented at the S Society of Glass Technology Centenary conference (SGT100) &
13th Symposium of the European Glass Society of Glass Science and Technology ,4-8
September, 2016 Sheffield UK
[87]. Macrelli, G. Some unconventional analytical solutions to diffusion equations in thermal
ion exchange in silicate glasses. Presented at the Society of Glass Technology Centenary
conference (SGT100) & 13th Symposium of the European Glass Society of Glass Science and
Technology (ESG2016); 4-8 September, 2016: Sheffield UK

76
[88]. Varshneya, A.K., Olson, G.A., Kreski, P.K., Gupta, P.K. Buildup and relaxation of
stress in chemically strengthened glass. J. Non-Cryst. Solids.2015;427:91-97.
[89]. Varshneya,A.K. Mechanical model to simulate buildup and relaxation of stress during
glass chemical strengthening, J. Non-Cryst. Solids.433 (2016) 28-30
[90] Macrelli, G.: Chemically strengthened glass by ion exchange: strength determination. Int
J Appl Glass Sci. 2017; 00:1-11 doi./10.1111/ijag.12291
[91] Rouxel, T, Yoshida,S.: The Fracture Toughness of Inorganic Glasses,
J.Amer.Ceram.Soc. 2017, doi 10.1111/jace.15108
[92] Kennedy, D.P., Murley, P.C.: Impurity atom distribution from a two step diffusion
process, Proceedings of the IEEE, 1964 pp 623-624
[93] Malkovich,R.Sh.: Impurity diffusion from a deposited layer, Fiz.metal.metalloved., 1963,
15,No 6,pp. 880-884
[94] Huggins,M.L., Sun,K.H. J.Am.Ceram. Soc. 26,4 1943
[95] Dugnani,R.;Residual stress in ion exchanged silicate glass: An analytical solution, J.Non-
Cryst.Solids. 471 (2017) 368-378

77
APPENDICES
A1 Ion exchange from a deposited finite layer and post ion exchange heat
treatment
When the ion source that supplies cations to glass is limited because is made of a deposited
layer on the glass surface or because the glass has been extracted from the bath and kept at
high temperatures the conventional constant source boundary condition is no more valid. A
model to handle such type of IX processes can be proposed with a variable boundary
condition

First step from time t=0 to time t=τ

Diffusion equation:

∂ca ∂ 2c
= D1 ⋅ 2a
∂t ∂x

Boundary condition: ca (o, t ) = cs 0 ≤ t ≤ τ ; Initial condition: ca ( x, 0) = 0 x>0

Solution is:
 x 
(i) ca ( x, t ) = cs ⋅ erfc  ;
 2⋅ D ⋅t 
 1 
Second step from time t=τ to t=∞

Diffusion equation

∂ca ∂ 2c
= D2 ⋅ 2a
∂t ∂x

∂ca x
Boundary condition: =0 τ ≤ t ≤ ∞; Initial condition: ca ( x,τ ) = cs ⋅ erfc( )
∂x x =0 2 ⋅ D1 ⋅τ

It can be noted that the diffusion coefficient at step 1 (D1) is, in general, different from the
diffusion coefficient at step 2 (D2). The reason for this is to keep the solution as general as
possible so that step 2 can be considered, in full generality, at a different process temperature
of step 1. Solution to this problem can be found in the literature [15], [92], [93] nevertheless
reference [15] identifies the application to post IX thermal treatments and provides in the
appendix a not so practical solution to the general case, reference [92] reports a typo mistake
in the final expression and reference [93] treats the case only for D1=D2. A corrected solution
has been presented by me in [77] and [87]:


2 ⋅ cs
ca ( x, t ) = ⋅ ∫ e − y ⋅erf ( k ⋅ y ) ⋅ dy
2
(ii)
π x
γ

78
D1 ⋅τ
γ = 2 ⋅ D1 ⋅τ + D2 ⋅ t k=
D2 ⋅ t

The overall (Step1+Step2) process time is t’=τ+t. It can be demonstrated that, in the limiting
x
case t’→τ (that is t=0) , as expected, the solution (ii) → ca ( x,τ ) = cs ⋅ erfc( ) that is to
2 ⋅ D1 ⋅τ
the boundary condition. More interesting is the limiting case for large times t→∞. It can be
demonstrated that, in this last limiting case, the solution is:

2 ⋅ cs D1 ⋅τ − x
2
4⋅D2 ⋅t '
ca ( x, t ) = ⋅ ⋅e
π D2 ⋅ t '

The total quantity of ions Q entered into the glass during step 1 can be easily evaluated
(equation (2.37)):

D1 ⋅τ
QAB = 2 ⋅ cs ⋅
π

This allows the expression of D1τ in terms of QAB and the limiting solution results:

2 ⋅ cs π QAB
2 2
1 −x QAB −x
4⋅ D2 ⋅t ' 4⋅ D2 ⋅t '
(iii) ca ( x, t ) = ⋅ ⋅ ⋅ ⋅e = ⋅e
π 2 cs D2 ⋅ t ' π ⋅ D2 ⋅ t '

Nice to realize that is exactly the solution for a semi-infinite medium with diffusion from an
infinitesimally thin layer with zero initial concentration ( equation (2.22)

This means that we can summarize the following conditions:

If t’≈τ we can use the constant source solution (i):

 x 
ca ( x, t ) = cs ⋅ erfc  ;
 2⋅ D ⋅t 
 1 

If t > 4·τ we can use the infinitesimally thin layer solution (iii)

2
QAB −x
4⋅ D2 ⋅t '
ca ( x, t ) = ⋅e
π ⋅ D2 ⋅ t '

In between these times τ < t’ < 4·τ we shall use the exact solution (ii)


2 ⋅ cs
ca ( x, t ) = ⋅ ∫e
− y2
⋅erf ( k ⋅ y ) ⋅ dy
π x
γ

79
D1 ⋅τ
γ = 2 ⋅ D1 ⋅τ + D2 ⋅ t k=
D2 ⋅ t

A1 – Figure 1 Boundary Conditions: A) Continuous Source from a virtually infinite ion


reservoir and B) Finite source of ions from a deposited layer

In Figure 1 it is expressed the boundary condition for a continuous source with ions supplied
by a virtually infinite reservoir (A) and the boundary condition (B) for a finite source of ions
from a deposited thick layer. In both cases it is assumed that the interface Ion Source/ Glass
reaches an almost instantaneous equilibrium condition.

80
Calculation examples:

A - Typical for post heating treatments after IX

Initial bath immersion time = τ


Post Heat treatment = t1 and t2

D1=8.5·10-12cm2/s D2=2.0·10-12cm2/s - τ=8h t1=0.5h t2=32h

Cht1 - Exact solution at time t’=τ+t1


Cht2 - Exact solution at time t’=τ+t2
Cerfc – Limit Solution for continuous source at time τ
Cthin – Limit Thin film solution at time t’=τ+t2

81
B - Typical for post heating treatments after IX

Initial bath immersion time = τ


Post Heat treatment = t1 and t2

D1=8.5·10-12cm2/s D2=2.0·10-12cm2/s - τ=8h t1=0.5h t2=64h

Cht1 - Exact solution at time t’=τ+t1


Cht2 - Exact solution at time t’=τ+t2
Cerfc – Limit Solution for continuous source at time τ
Cthin – Limit Thin film solution at time t’=τ+t2

82
C - Typical for IX from deposited layers

Initial constant source time = τ


Drive in time = t1 and t2

D1=5·10-10cm2/s D2=5·10-10cm2/s - τ=360s (6 minutes) t1= 60s t2=1440s (24 minutes)

Cht1 - Exact solution at time t’=τ+t1


Cht2 - Exact solution at time t’=τ+t2
Cerfc – Solution for continuous source at time τ
Cthin – Thin film solution at time t’=τ+t2

83
A2 - Relationship between the exchanged molar flux and the process
parameters
Here I will give a general relationship between the total exchanged molar flux and the process
parameters without making any assumption or approximation for concentration dependence of
the interdiffusion coefficient or of the concentration profile.

We start from the flux equation:

 ∂c 
J = − D AB ⋅  
 ∂x 
making the variable transformation (Boltzman transformation)

x
z=
D ⋅τ

The derivative transformation is

∂ 1 ∂
= ⋅
∂x D ⋅ t ∂x

and the flux equation is :

D ∂c
J =− ⋅
τ ∂z

We can calculate the total exchanged molar flux integrating the flux at the glass surface
boundary, J0 = J(z=0)

 ∂c  Do  ∂c 
J o = − Do ⋅   = − ⋅
 ∂x  x =0 τ  ∂z  z =0

(Do is the interdiffusion coefficient at the glass surface) obtaining:

τ τ
 ∂c  1  ∂c 
Q AB = ∫ J o ⋅ dτ = − Do ⋅   ⋅ ∫ dτ = −2 ⋅ Do ⋅ τ ⋅ 
0  ∂z  z =0 0 τ  ∂z  z =0

that is related to the measured weight gain (∆W) by:

∆W
Q AB =
S ⋅ (M A − M B )

if the ion concentration profile is:


 x   z
C ( x , t ) = Co ⋅ erfc  = Co ⋅ erfc 
 2⋅ D⋅ t   2

84
it results:

 ∂C   Co  z 2  Co
 ∂z  =  − ⋅ exp  −  = −
  z =0  π  2   z =0 π

and the total exchanged molar flux:

D ⋅τ
Q AB = 2 ⋅ Co ⋅
π

85
A3 – Terminology
Mechanical Characteristics

Residual stress Profile σ = σ (x)

Surface Compression SC = σ (0)

Case Depth CD - Depth of compression layer σ (xo) =0 ⇒ CD = x o

Internal Tension - t = glass sheet thickness - τ = σ (t/2)

Strength - MOR

Testing methods:

Bar test ⇒ ASTM C 158 – EN 1288-3


Ring on ring test ⇒ EN 1288-5

Original glass strength - σ0

Final glass strength - σb

Strength increase - ∆σ = σb - σ0

Abraded Strength - σ*

Kinetic characteristics

Ion Concentration profile - c = c(x;t)


Units: mol/cm3; %weigth/cm3;
rel. Conc - c(x;t)/c(0;t)

Depth of ion-exchanged layer Cδ:

c(Cδ) = 0

Ion concentration profile is the difference between the ion concentration and the initial bulk
ion concentration

c(x;t) = CK+(x;t) - Ck+

Chemical Interdiffusion Coefficient - DAB (cm2/s)

∂c
J = − D AB ⋅
∂x

86
∂c ∂  ∂c 
=  D AB (c ) ⋅ 
∂t ∂x  ∂x 

Mass Average Diffusion Coefficient - DM (cm2/s)

∂c ∂2
= D M ⋅ 2 c ( x, t )
∂t ∂x

87
A4 Evaluation of the depth of the compression layer using the “thermal
stress analogy” equation.
Let us recall the equation we have introduced in (2.62) for the evaluation of the stress
generated in a plate (-L,L) by a not uniform concentration distribution c(x,t)

B⋅E   B⋅E  1 L 
L
1
σ ( x, t ) = − ⋅ c ( x , t ) − ⋅ ∫ c( x, t ) ⋅ dx  = ⋅ ⋅ ∫ c( x, t ) ⋅ dx − c( x, t )
(1 − ν )  2 ⋅ L −L  (1 − ν )  2 ⋅ L − L 

Indicating with:

L
M t = ∫ c( x, t ) ⋅ dx
−L

The amount of ions entered in the slab at time t; and with:

L
M ∞ = ∫ c s ⋅ dx = 2 ⋅ L ⋅ c s
−L

The amounts of ions entered at infinite time we can define the ratio:

L
Mt 1
M ∞ 2 ⋅ L ⋅ c s −∫L
= ⋅ c( x, t ) ⋅ dx

With this definition the stress equation results:

B⋅E  M 
σ ( x, t ) = ⋅ c s ⋅ t − c ( x , t ) 
(1 − ν )  M∞ 

Specifically surface stress (Surface compression - SC) results:

B ⋅ E ⋅ cs M 
σ ( L , t ) = SC = ⋅  t − 1
(1 − ν ) M∞ 

Introducing the normalized stress function Fs [ Ref. H.M.Garfinkel, C.B.King “Ion


concentration and Stress in a Chemically Tempered Glass, J.Amer.Ceram.Soc. Vol 53, No12
– December 1970, pp.686-691]

Mt M t c ( x, t ) c ( x , t ) M t
cs ⋅ − c ( x, t ) − −
σ ( x, t ) M∞ M∞ cs cs M∞
Fσ = = = =
σ ( L, t ) M Mt M
cs ⋅ t − cs −1 1− t
M∞ M∞ M∞

Let us now recall the concentration solution introduced in (2.36):

88
 4 ∞ (− 1)n  (2 ⋅ n + 1)2 ⋅ π 2 ⋅ D AB ⋅ t    2 ⋅ n + 1  
c a ( x, t ) = c s ⋅ 1 − ⋅ ∑ ⋅ exp −   cos  ⋅ π ⋅ x  
 π n=0 2 ⋅ n + 1  4 ⋅ L2   2 ⋅ L  

And let us calculate :

_
Mt 1
L
c 8 ∞ 1  (2 ⋅ n + 1) 2 ⋅ π 2 
= ⋅ ∫ c( x, t ) ⋅ dx = = 1− 2 ⋅ ∑ ⋅ exp − ⋅ D AB ⋅ t 
M ∞ 2 ⋅ L ⋅ cs −L cs π n = 0 (2 ⋅ n + 1) 2
 4⋅ L 2

In fact:

_ L L
1 1
c= ⋅ ∫ c( x, t ) ⋅ dx = ⋅ ∫ c( x, t ) ⋅ dx =
2 ⋅ L −L L 0

 2 ∞
(− 1)
n
 (2 ⋅ n + 1)2 ⋅ π 2 ⋅ D AB ⋅ t  L   2 ⋅ n + 1  
= c s ⋅ 1 − ⋅∑ ⋅ exp − ∫ cos  ⋅ π ⋅ x   ⋅ dx =
 π ⋅ L n =0 2 ⋅ n + 1  4 ⋅ L2  − L  2 ⋅ L  

 2 ⋅ 4 ⋅ L ∞ (− 1)n (−1) n  (2 ⋅ n + 1)2 ⋅ π 2 ⋅ D AB ⋅ t  


= c s ⋅ 1 − ⋅∑ ⋅ ⋅ exp −  =
 π ⋅ L ⋅ π n=0 2 ⋅ n + 1 (2 ⋅ n + 1)  4 ⋅ L2  
 8 ∞ 1  (2 ⋅ n + 1) 2 ⋅ π 2 
= c s ⋅ 1 − 2 ⋅ ∑ ⋅ exp − ⋅ D AB ⋅ t  
 π n =0 (2 ⋅ n + 1) 4⋅ L
2 2
 

the above result is because:

  2⋅ n +1  4⋅L
L

∫ cos ⋅ π ⋅ x  ⋅ dx = ⋅ (−1) n
−L
2⋅l  (2 ⋅ n + 1) ⋅ π

With these positions the depth of the compression layer – CD is the zero of the normalized
stress function Fs, that is CD is the x value for which:

Fσ ( x = C D , t ; L) = 0

89
A5 – Non Isothermal Ion-Exchange
We are now dealing with Ion Exchange processes where the salt bath temperature is not
constant but it is a function of time:

T = T(t)

Still considering equation:

∂c( x, t ) ∂ 2 c ( x, t )
= D⋅
∂t ∂x 2

As temperature is changing, diffusion coefficient will be:

D = D (T )

according to:

 H 
D (T ) = DO ⋅ exp − 
 R ⋅T 

Where H is activation energy of the diffusion process (typically ranging from 120 kJ/mole to
170 kJ/mole), T is the temperature in Kelvin, R is the universal gas constant :

R = 8,314 J/(mol.K).

This means that D will be a time function through T=T(t) and diffusion equation will result:

∂c ( x, t ) ∂ 2 c ( x, t )
= D (t ) ⋅
∂t ∂x 2

In order to solve this equation under the boundary conditions:

c( x ≥ 0; t = 0) = 0
c( x = 0; t ≥ 0) = C o

We introduce the following definition:

t
τ (t ) = ∫ D(η )dη
0

So it results:

∂τ ∂
t
= ∫ D (η )dη = D(t )
∂t ∂t 0

90
∂c ∂c ∂τ ∂c
= ⋅ = D (t ) ⋅
∂t ∂τ ∂t ∂t

And diffusion problem will be:

∂c( x,τ ) ∂ 2 c( x,τ )


=
∂τ ∂x 2

c ( x ≥ 0;τ = 0) = 0
c ( x = 0;τ ≥ 0) = C o

With the solution:

  
c( x,τ ) = C o ⋅ erfc
x


 2 ⋅ (τ )
1/ 2
 

That is:

 x t −1 / 2
 −1 / 2

c( x, t ) = C o ⋅ erfc ⋅  ∫ D (η )dη 
  = C ⋅ erfc x ⋅ ( D * ⋅ t  
o  
 2   2  
   0    

Where:

∫ D(η )dη
D =
* 0

The above definition of the diffusion coefficient for a Non isothermal process allows a
straightforward interpretation.

Considering an overall process time tF where we had partial times tj with different process
temperatures Tj we can define a kind of “averaged” diffusion coefficient as:

tF N
D * ⋅ t F = ∫ D(η )dη = D1 (T1 ) ⋅ t1 + D2 (T2 ) ⋅ t 2 + .... + D N (TN ) ⋅ t N = ∑ D j (T j ) ⋅ t j
0 j =1

That allows the treatment of Non-Isothermal process in a similar way of Isothermal processes.

91
A6 - Viscosity –
Viscosity is a measure of the resistance of a medium to shear deformation with time. If σxy is
the shear stress and exy the shear strain (deformation) viscosity η is defined as:

σ xy = η ⋅ eɺ xy (1)

Where:

d
eɺ xy = (e xy ) (2)
dt

Viscosity units are Pa·s, sometime another unit is used as poise as:

1 Pa·s = 10 poise (3)

In pure elastic bodies (where viscosity may be considered infinite) a similar relationship
between shear stress and strain hold :

σ xy = G ⋅ e xy (4)

Where G is the elastic shear modulus in units (MPa). This means that when a shear stress acts
on an pure elastic body the body undergoes instantaneous equilibrium shape change, on the
other side if the shear stress acts on a viscous body the body starts flowing or deforming with
time.

Viscosity reference points – One of the most important glass properties is the variation of its
viscosity with temperature. This so important for manufacturing processes that five reference
points have been defined as:

Strain point η = 1013,5 Pa·s


Annealing point η = 1012 Pa·s
Softening point η = 106,65 Pa·s
Working point η = 103 Pa·s
Melting point η = 10 Pa·s

Viscoelasticity and viscoelastic solid models - In general terms, apart when viscosity is
really very high (room temperature), the response of a glass body to an applied mechanical
stress is a time-dependent structural change combining a solid like and liquid like behaviour.
We define this behaviour as “viscoelastic”. One important parameter in the viscoelastic
behaviour is the relaxation time τ, defined as:

η
τr = (5)
G

In terms of relaxation times the first two viscosity reference points may be defined as:

Strain point → Stress relaxation is a matter of hours


Annealing point → Stress relaxation is a matter of minutes

92
The viscoelastic response of a solid is normally represented by the association of an elastic
(spring like) element and a viscous (dashpot like) element. The simplest ways to do that is
coupling them in series (Kelvin model) or in parallel (Maxwell model) according to the
following models:

Kelvin solid:

σ = σ1 + σ 2 ; σ 1 = G ⋅ e ; σ 2 = η ⋅ eɺ (6)

So that:

σ = G ⋅ e + η ⋅ eɺ (7)

Where it has been eliminated the (xy) shear pedices. The Kelvin solid represent the series of a
spring and dashpot .

Maxwell solid :

e = e1 + e2 ; σ = G ⋅ e1 ; σ = η ⋅ eɺ2 (8)

So that: σɺ = G ⋅ eɺ1 ; eɺ = eɺ1 + eɺ2 we have:

σɺ σ
+ = eɺ (9)
G η

The Maxwell model represent a parallel between a spring and a dashpot. To better understand
what this means, let’s make the following considerations.

Consider the Kelvin solid under constant applied stress σ0:

σ 0 = G ⋅ e + η ⋅ eɺ (10)

The solution for e(t) is:

σ0
e(t ) =
  t  (11)
G ⋅ 1 − exp − 
  τr 

This means that the solid deforms or flows under constant stress once time is of the order of τr
(see figure7). This property is called “creep”.

93
Figure 1 – Kelvin solid – Creep

Consider now The Maxwell solid under constant deformation e0 so that:

σɺ σ
eɺ0 = + =0 (12)
G η

The solution for σ(t) is (σ0 = σ(t=0)):

t
σ (t ) = σ 0 ⋅ exp(− ) (13)
τr
This means that to apply a constant deformation a vanishing stress has to be applied (see
figure 2) and stress vanishes once time is of the order of τr . This property is called
“retardation”.

Figure 2 – Maxwell solid – Stress relaxation

Because retardation (creep) and stress relaxation are both present in real glass behaviour, a
more realistic solid model can be considered coupling in series a Kelvin model and a Maxwell
model. This is called Burger solid model. As told above relaxation time τr is a good parameter
to understand at which time scale the viscoelastic solid will behave elastically or relax. In the
glass transition viscosity range (1012 ÷ 10 12,5 Pa·s ) glass shear modulus is in between 1010

94
÷1011 Pa so that relaxation time is in the order of 100 s. This means that at glass transition
temperature stress relaxation occurs in about a couple of minutes.

The Burger solid model is an association of a Maxwell (G1, η1) element and a Kelvin element
(G2, η2) in series. Under applied constant stress, strain is resolved as:

1 1  t  t 
e(t ) = σ 0 ⋅  + ⋅ 1 − exp(−  +  (14)
 G1 G 2  τ 2r  τ 1r 

Where we have defined the two relaxation times:

η1
τ 1r = ; (15)
G1

η2
τ 2r = (16)
G2

Solution (14) is plotted in figure 9

Figure 3 – Plot of the Burger model

σ0
Initial elastic strain - ei = ;
G1
σ0   t 
Delayed elastic strain - ed = ⋅ 1 − exp −  ;
G2   τ 2r 
σ0 ⋅t
Viscous strain - ev = τ 1
r

The first term of equation (14) produces an instantaneous elastic strain, the second term is a
delayed elastic strain and the third term is a viscous strain. As stress is removed the first two
terms recover while viscous terms remains (irreversible term).

95
A7 Thermodynamics of Ion Exchange
When a glass matrix containing alkali ions “A” is put in contact with a reservoir containing
alkali ions “B” than an ion-exchange reaction process can be activated according to the
following scheme

A + B ⇔ A+ B (1)

Where A and B are the ions in the glass matrix and A,B the ions in the reservoir. Let’s
evaluate the thermodynamics of ion-exchange. The process is driven by the electrochemical
potentials of the ions in the glass and reservoir matrices,
In the glass matrix the electrochemical potential can be written as:

η i = µ i + zi ⋅ F ⋅ ϕ
i=A,B (2)
µ i = µ i' + µ i''

Where µ i is the composition dependant chemical potential while µ i is the contribution due to
' ''

change in the mechanical strain energy as the incoming ion is taking the place of the
outcoming ion.

In the reservoir (typically molten salt medium) the electrochemical potential will be:

ηi = µi + zi ⋅ F ⋅ ϕ i=A,B (3)

At equilibrium and considering ion exchange for only monovalent ions ( zi =1) we will have
for ions A and B:

Ion A:

µ A' + µ A'' + F ⋅ ϕ = µ A + F ⋅ ϕ
(4)
F ⋅ (ϕ − ϕ ) = µ A − µ A' − µ A''

Ion B

µ B' + µ B'' + F ⋅ ϕ = µ B + F ⋅ ϕ
(5)
F ⋅ (ϕ − ϕ ) = µ B − µ B' − µ B''

It will results:

µ B − µ B' − µ B'' = µ A − µ A' − µ A'' (6)

Considering the relationship between chemical potential and activity :

96
µ A = R ⋅ T ⋅ ln(a A )
µ A' = R ⋅ T ⋅ ln(a A )
µ B = R ⋅ T ⋅ ln(a B )
µ B' = R ⋅ T ⋅ ln(a B )

Equation (2.22) will result:

a A ⋅ aB
µ B'' − µ A" = R ⋅ T ⋅ ln (7)
aB ⋅ a A

Activity of an ion specie “A” is related to concentration of ion specie “A” through the activity
coefficient γA:

aA = γ A ⋅ cA (8)

A
So (introducing a selectivity coefficient K B ) :

a A ⋅ aB γ γ c ⋅ c  γ γ
ln = ln  A ⋅ B ⋅ A B  = ln A + ln B + ln K BA
aB ⋅ a A γ B γ A cB ⋅ c A  γB γA
(9)
c ⋅c
K = A B
A

cB ⋅ c A
B

Indicating with wM the mechanical work involved in the replacement of 1 mole of ion A by 1
mole of ion B (it includes any work that may be done against external forces)

wM = µ B" − µ A" (10)

It finally results:

γB γ w
ln K BA = ln + ln A + M (11)
γA γ B R ⋅T

Assuming that the ratio of activity coefficients is the same in glass matrix and salt melt:

γA γA
= (12)
γB γB

Equation (11) results:

wM
ln K BA ≅ (13)
R ⋅T

97
A
Evaluating K B in some way (as an example through equation (13)) than it is possible to
evaluate the fraction of exchanged ions at glass surface of the glass matrix versus the mole
fraction of “A” ions in molten salts.

A
From the definition of K B :

cA c
= K BA ⋅ A (14)
cB cB

The ratio ca/cb is the mole fraction of ion “A” in molten salt, the fraction of exchanged ions at
surface is:

cA
χ A⇔ B = 1 − (15)
cB

And through equation (14) it results:

cA
χ A⇔ B = 1 − K BA ⋅ (16)
cB

A
The selectivity coefficient K B , can be evaluated through equation (13) by the evaluation of
wM. This can be done, at a first approximation, using the “Misfitting Spheres Theory” as:

E (∆Vo ) 22

wM = ⋅ + ⋅ S c ⋅ ∆V o
3 ⋅ (1 − ν ) 3 ⋅ V B 3

E – Glass Matrix Young Modulus


ν – Glass Matrix Poisson ratio
∆Vo – Increase of volume in glass matrix due to exchange of ion A with ion B
VB – Volume of 1 mole of ion B (incoming ion)
Sc – Surface compression

In this way the selectivity coefficient can be calculated as:

w 
K BA = exp  M  (17)
R ⋅T 

98
A8 – Presentation of Glass chemical composition formulae[B1]
Reviewing scientific and technical literature it can be found a number of different ways in
which glass chemical composition formulae are presented: mol%, wt% mole formula, mole
fraction and so on. I think it can be useful to establish general conventional criteria to be
considered when presenting glass chemical composition.
The general first accepted convention in a glass formula is:

List glass network modifiers in increasing order of valency finishing with network formers

If we have a glass made of 1 mole of Sodium Oxide, 1 mole of Calcium Oxide and 5 moles of
Silicon Dioxide the mole formula will be:

1 Na2O 1CaO 5 SiO2 Mole Formula

Conversion of this into mole fraction is straightforward, in total there are 7 moles so the molar
fraction for each oxide will be the ratio between the singular mole number with the total
moles number:

0.143 Na2O 0,143 CaO 0,715 SiO2 Mole Fraction Formula

Conversion to mol% is obtained just multiplying the mole fraction by 100:

14.3 Na2O 14.3 CaO 71.5 SiO2 mol% Formula

To convert this into wt% molecular weights shall be considered:

MW(Na2O) = 62 g /mol
MW(CaO) = 56 g/mol
MW(SiO2) = 60 g/mol

Based on this a glass represented by the above first mole formula contains 62 g of Na2O, 56 g
of CaO and 300 g of SiO2 over a total weight of glass - total = 62+56+300 = 418g. Now it is
obvious the conversion:

14.8 Na2O 13.4 CaO 71.8 SiO2…………wt % Formula

The back conversion from wt% to mol% is, on the above indicated steps quite obvious. One
nice thing to notice is that, for sodalime silicate glasses, the chemical composition formula
from mol% to wt% does not change too much. Other minor constituent in glass chemical
composition (colorants, glass fining composites, other oxides coming from raw chemicals) are
normally listed at the end of glass formula.

In general formulation based on wt% are used in glass melting literature while formulation
based on mol% are used in glass studies of structure and physical and chemical properties.

99
A9 – Molar Volume determination[B1]

From the chemical point of view silicate glasses may be viewed as solution of inorganic
oxides. Molar volume defined as the volume of one gram of glass can be considered as an
extensive property and may be expressed in terms of partial molar volume of individual
species or structural groups. A good computational method for glass molar volume (which is
of relevance in calculating zero time surface compression)is presented in [B1] based on the
concept of additivity of component oxides. The method works as follows:

fM is the weight fraction of each component oxide MnOm in the glass.


sM is the ratio of the number of Oxygen atom in the molar formula to the formula weight and
it can be taken from Table A1 column 2.(Again from [B1] taken from Huggins and Sun [94]).

Calculate :

∑ M
sM ⋅ fM

f Si
N Si =
60.06 ⋅ ∑ M s M ⋅ f M

This last number NSi is the number of gram atoms of Si per gram atom of Oxygen in the glass.
According to the value of NSi the value vM can be selected for any component oxide in table
A1 from the corresponding NSi column. This, finally, allows the calculation of the specific
volume of the glass:

v = ∑ M vM ⋅ fM

The last column in Table I1 is for “not well annealed” glasses. Incidentally the glass density
is:

1
ρ=
v

100
Table – A-1 From [B1] and [94]

101

View publication stats

You might also like