You are on page 1of 18

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

net/publication/333757607

Enhancing mechanical endurance of chemical-tempered thin soda-lime


silicate float glass by ion exchange

Article · June 2019


DOI: 10.1007/s41779-019-00375-x

CITATIONS READS

4 418

6 authors, including:

Aydın Güzel Meryem Sarıgüzel


Gebze Technical University TUBITAK Marmara Research Center
2 PUBLICATIONS   4 CITATIONS    10 PUBLICATIONS   38 CITATIONS   

SEE PROFILE SEE PROFILE

Melis Özdemir
TUBITAK Marmara Research Center
5 PUBLICATIONS   19 CITATIONS   

SEE PROFILE

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

Development of Nanometal Glass Hybrid Nanocomposites View project

All content following this page was uploaded by Melis Özdemir on 16 June 2019.

The user has requested enhancement of the downloaded file.


Journal of the Australian Ceramic Society
https://doi.org/10.1007/s41779-019-00375-x

RESEARCH

Enhancing mechanical endurance of chemical-tempered thin


soda-lime silicate float glass by ion exchange
Aydın Süleyman Güzel 1,2 & Meryem Sarıgüzel 1 & Melis Can Özdemir Yanık 1 & Esin Günay 1 & Metin Usta 1,2 &
Yusuf Öztürk 1

Received: 27 September 2018 / Revised: 3 February 2019 / Accepted: 18 May 2019


# Australian Ceramic Society 2019

Abstract
The aim of this study is to determine the mechanical properties differences in between the air and tin surfaces of thin soda-lime
silicate float glasses that subjected to ion exchange using KNO3 salt bath. The ion exchange process was carried out at different
times and temperatures. Chemically tempered glasses were investigated by means of compressive stress (CS), microhardness
measurements, cracking probability, fractographic analysis, and flexural strength. The relationship among these properties was
also discussed. The Weibull distributions of the samples were determined for a better understanding of the strength results. The
AFM was used to determine the surface roughness. The weight of glass samples was increased gradually with increasing ion
exchange time and temperature due to the inter-diffusion of K+–Na+ ions. The fracture load of chemically strengthened glass with
ion exchange at 435 °C-8 h showed an increase of ~ 4.4 times that of untreated glass (raw glass), and it was selected as the
optimum process conditions. The number of broken pieces was increased by increasing flexural strength, and smaller pieces were
obtained with a great deal of branching. The air side always has greater compressive stress than the tin side. The maximum
hardness value was reached with ion exchange at 435 °C for 12 h on a tin surface was 8.25 GPa with an increase of ~ 18% with
respect to the raw glass. The crack resistance of chemically tempered glasses showed an increase in the range of 410–1290% and
241–1895% for air and tin surfaces, respectively. According to the AFM analysis, surface roughness of the samples after ion
exchange did not change dramatically. SEM-EDS analysis revealed that the surface potassium concentration and diffusion depth
increase with temperature.

Keywords Chemical tempering . Ion exchange . Soda-lime float glass . Mechanical strength . Indentation . Compressive stress

Introduction thermally tempered glass, and thermal tempering cannot be


used for strengthening complex-shaped or for low thermal
The Ultra-Thin Glass Market was estimated to be USD 9.73 expansion glasses [2–4]. All of these reasons will be led the
billion in 2017 and is foreseen to reach USD 16.99 billion by world’s leading glass companies to focus on improving the
2022, at a CAGR (Compound Annual Growth Rate) of 11.8% chemical tempering process (strengthening) of glass [5–7].
from 2017 to 2022 [1]. The decrease in glass thickness re- Increasing its mechanical strength by modifying its surface
duces the mechanical strength of the glass, and the thin glasses with a chemical tempering process, glass has started to be used
cannot theoretically reach the appropriate mechanical strength progressively in applications such as computers, tablets, mo-
through thermal tempering. In addition, any surface treatment bile phones and television screens, cover glasses used in solar
(such as cutting, drilling, or shaping) cannot be performed on cell panels, automotive glasses, and many other industrial uti-
lization [8].
In the chemical strengthening process by ion exchange,
* Yusuf Öztürk
ozturk.yusuf@tubitak.gov.tr
alkali ions on the glass surfaces are replaced with alkali ions
in the salt when a glass is immersed in the molten salt bath
1
TÜBİTAK, Marmara Research Center, Materials Institute, Gebze, under the glass transition temperature [9–11]. The radius of
Kocaeli, Turkey the ions diffused into the glass surface is greater than that of
2
Material Science and Engineering, Gebze Technical University, the ions coming out of the glass surface, thus the glass is
Gebze, Kocaeli, Turkey strengthened by creating compressive stress [12]. Besides
J Aust Ceram Soc

submerging in a molten salt bath, there are many different ion Experimental
exchange application techniques, such as exposing to salt va-
por, spraying the salt solution onto the glass surface, and coat- A ŞİŞECAM soda-lime silicate (SLS) float glass which
ing the surface of the glass by tempering the salt solution into has a thickness of 1.1 mm and 60 × 60 mm2 surface area
paste consistency [13, 14]. was used in this study. The chemical composition and the
Ion exchange method, which is generally used for the im- transition temperature BT g ^ of the glass are given in
provement of mechanical strength, also increases the thermal Table 1.
shock resistance, stiffness, and scratch resistance of glass and The glass samples were cleaned in ethanol and deionized
can change the optical, electrical, chemical, and physical prop- water for 5 min, respectively, by using an ultrasonic cleaner
erties of glass, such as refractive index, electrical conductivity, before the ion exchange process. The same cleaning process
and chemical composition [15–18]. was carried out after the ion exchange to remove salt residues
The glass bending strength is determined by the geometry, from the glass surface. The glass samples were preheated for
depth, distribution, and density of glass flaws, which consid- 1 h at 180 °C and then were placed in the molten KNO3 salt
erably reduce the theoretical strength of the glass [19, 20]. For solution. The schematic workflow diagram of the ion ex-
this reason, it is important to limit the formation and propaga- change process and all experimental conditions are given in
tion of cracks by creating compressive stress on the glass Fig. 1.
surface through ion exchange in order to increase the mechan- The tin surfaces of the glass samples were placed in
ical strength of the glass [21, 22]. contact with the bottom of the steel tank. The glass
There are many papers and patents on strengthening glass weight/salt weight ratio was kept constant at 1/5, so the
through ion exchange, and in these studies, the effects of ex- results did not influence by the change in the salt amount.
perimental parameters, such as process time and temperature, In this study, the Sigma Aldrich KNO3 salts were used
glass composition and thickness, and salt type and concentra- (purity > 99%), and the glass samples were immersed into
tion, were investigated [23, 24]. 100% the molten KNO3 salt bath in an electric furnace
Many flat glasses used for automotive and architectural (Protherm Chamber Furnace) at different temperatures
applications and large-sized glass products are produced (375 °C, 400 °C, 425 °C,435 °C, 450 °C, and 475 °C)
by the float glass production process, which was discov- for different times (4, 8, and 12 h).
ered by Pilkington [25]. In this production process, glass is Weight increases of the ultrasonically cleaned samples be-
produced by floating on the molten tin, and during produc- fore and after the ion exchange process were measured by a
tion, one surface of the glass is in contact with molten tin. digital scale weighing device with 1-mg measurement
Due to tin diffusion to one surface of the glass, chemical precision.
compositions of glass surfaces are differed from each other The flexural strength of chemically strengthened glass
and are called as air and tin surfaces [26]. The composition samples was measured by a ring-on-ring test with loading
and structural differences between the two sides can lead to and support diameters of 20 mm and 40 mm, respectively.
different compressive stress, depths of stress layers, and The biaxial flexural test fixture was measured according to
dissimilar diffusion characteristics, which can affect the the EN1288-5 norm. Tests of all glass samples were carried
properties of glasses [27–29]. out by applying loads on the air side and specimens were
As a result, the diffusion behavior of K+–Na+ ions on both taped on the tin side before the test to hold fragments
surfaces is very important for controlling the properties of together.
float glasses. However, the difference between the diffusion Afterwards, the fracture load distributions were studied
characteristics of the K+–Na+ ions on the tin and air surfaces using a conventional two-parameter Weibull approach [31].
of float glass is not still clear yet [30]. Moreover, the hardness, Vickers hardness measurements of the glass samples
compressive stress, and surface roughness change differences were performed before and after ion exchange treatment.
between the air and tin surfaces of the chemically tempered The indentations were created by using CSM Instrument
soda-lime silicate float glasses are still uncertain. Micro Hardness Tester with a loading capacity ranging
This paper clarifies the difference between the compressive
stresses (CS) and depths of stress layers (DOL) of air and tin
Table 1 Chemical composition and transition temperature (Tg) of
surfaces of float soda-lime silicate glasses and determines the
untreated (raw) soda-lime silicate glass
effect of process conditions to the final mechanical perfor-
mance of two surfaces. The effect of the ion exchange behav- XRF glass composition, wt% Tg (°C)
ior (strengthening process) of the tin and air surfaces of float
SiO2 Al2O3 Fe2O3 TiO2 CaO MgO Na2O K2O SO3
soda-lime silicate glass was examined systematically, and the
optimum results were obtained that are thought to provide an 72.33 0.99 0.09 0.054 8.43 4.40 13.45 0.06 0.20 560
important contribution to literature.
J Aust Ceram Soc

Fig. 1 Schematic workflow diagram of the ion exchange process and all experimental conditions

from 1 g (0.0098 N) to 10,000 g (98 N) for 7-s dwell time. Japan). Five different measurements of both the air and tin
Observation of the indentations and measurement of diag- surfaces were performed for each glass sample. The average
onal lengths were made using an optical microscope CS and DOL values and standard deviation calculations were
(Nikon MB 10) connected to the Vickers Micro Hardness obtained to verify the reproducibility of the measurements.
Tester. The air and tin sides were examined separately. Phase characterizations of samples were carried out by
Microhardness measurements were performed at 0.49 N. using Shimadzu™ XRD-6000 X-ray diffractometer in the
In addition, the Vickers Indentation method was used to 2θ range of 2–70° with the step size of 0.02° using a Cu
evaluate the crack resistance and the crack formation proba- radiation tube (wavelength 1.5405 Å).
bility in glass. The percentage of crack formation probability After the ring-on-ring test, some fragments were select-
was obtained by dividing the number of corners where radial ed from the clean-flat surfaces of broken samples and then
cracks are formed by the total number of the corners indenta- coated by sputtering Platinum (Pt). The K/Na ratios were
tions. The applied load was elevated incrementally, and twen- determined from sodium and potassium concentration pro-
ty Vickers hardness indentations were performed for each ap- files with a point analysis at 1-micron intervals (about
plied load (total 80 corners). From the crack initiation proba- 30-μm long) by using the JEOL JSM-6510-LV scanning
bility (%) versus the indentation load (N) graph, the indenta- electron microscopy (SEM) and Oxford Instruments Inca
tion load corresponding to 50% crack initiation probability Energy Dispersive Spectroscopy (EDS). All the analyses
was determined as Bcrack resistance.^ Both the air and tin were performed for each glass sample on the air and tin
sides were examined separately [18, 32]. sides separately to investigate the different diffusion per-
The AFM was used to determine the surface roughness. formance of the two sides.
The AFM measurements were carried out using a model All data are presented as mean ± standard deviation(S).
Quesant Ambios AFM Q-Scope (Nano Scope Tapping STDEV.S uses Eq. 1:
Mode TM). The per scans of a surface region of 40 × 40 μm vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

u  2
were investigated. The measurements were repeated for each u ∑ x−x
t
sample on different spots of the surface. ð1Þ
After the ion exchange process, compressive stress (CS), ðn−1Þ
depths of stress layers (DOL), and central tension (CT) of the
samples were determined by birefringence stress analysis where x is the sample mean average (data 1, data 2,…), and n
(FSM-6000LE model surface stress meter, ORIHARA, is the sample size.
J Aust Ceram Soc

Results and discussions where Ma, Mb is the molecular mass of the ion-exchanged ions
(g/mol); Cs is the surface concentration per unit volume, and t
Weight increase results and evaluation is process time.
Weight increase which has an important role in the fast
Since potassium is a heavier alkali ion than sodium, the weight detection of possible errors, especially because the ability to
increase occurred as a result of the inter-diffusion of Na+ ions quickly and easily measure for industrial applications follow-
in the glass with K+ in the molten salt bath. The weight in- ing the determination of the optimum process parameter is
crease of the samples is given in Fig. 2 as a function of tem- necessary.
perature and time.
The weight of glass samples increased gradually with in-
creasing process temperatures (Fig. 2a) and times (Fig. 2b). A Strengthening (failure load) results and evaluation
total of 25 °C increment at ion exchange temperature caused
weight gain of about 2.5 mg, but the weight gain rate was In order to analyze the strengthening provided to the glass
higher at 450 °C (Fig. 2a). The analyses, as stated by Gy [2] surface by ion exchange, the ring-on-ring test was preferred
and Erdem et al. [18], showed that the weight of the samples because, in this test set up, a large tensile stress is not imparted
increased with rising temperature and time. Also, according to to the processed edges of the test sample. Besides, the Weibull
Gy [2], the measurement of the weight increase in glass de- distribution was used to understand the strength values better
pending on the temperature and time provides a quick and and to represent the statistical approach of the depth of distri-
useful way to define a constant inter-diffusion coefficient. bution of the surface defects (cracks). Weibull modulus (m),
The principle of this method is based on the weight change characteristic strengths (σ0), and bending strength values are
in glass samples as a result of the ion exchange process since presented in Table 2, and the graphs are shown in Figs. 3, 4,
the ionic weight of potassium (39,0983 u) is greater than the and 5.
sodium atomic weight (22,9897 u) [2, 33]. After the ion ex- Systematic experimental studies have been carried out on
change, the total mol amount (Q) of incoming ions can be the 12-h constant process time to determine the effect of the
calculated from the weight increase in the glass specimens ion exchange temperature on the fracture load (Fig. 3a). After
(ΔW) and incoming ions per unit surface area of this glass ion exchange, the increase in mechanical strength can be
(Sıx) (Eq. 2). This relation will allow the determination of clearly observed at all temperature and time parameters (Fig.
the mass average diffusion coefficient, Da (cm2/s) (Eq. 3) 3a, b). The results obtained showed that the fracture load in-
[33, 34]. Thus, it is possible to reconstitute the concentration creased continuously with increasing temperature in the range
depth profile using the value of the mass average diffusion of 375–435 °C, and the fracture load decreased at tempera-
coefficient and assess the depth of the ion exchange layer. tures above 435 °C. This result can be explained by the fact
that the glass shows a stress relaxation at temperatures above
ΔW
Q¼ ð2Þ 435 °C. The decrease in strength is attributed to the stress
S ðMa−M bÞ relaxation behavior of glass in the case of long ion exchange
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi times at the high temperatures. Glass network relaxation oc-
Da  t curs when larger ions are introduced in the glass, as the net-
Q ¼ 2  Cs ð3Þ
π work restructures so that new places for the larger ions are

16 16
(a) (b) 425˚C
14 14 435˚C
450˚C
Weight increase, mg

Weight increase, mg

12 12

10 10

8 8

6 6
4 4
2 2
0 0
350 375 400 425 450 475 500 0 4 8 12 16
Temperature, °C Time, h
Fig. 2 Weight increase of ion-exchanged samples for a varying temperatures (for 12 h) and b varying times
Table 2 All experimental data with standard deviation values
J Aust Ceram Soc

Sample Weight increase, Failure load, m (Weibull modulus) σ0, N % load of failure Surface Compressive stress, Depths of stress layer, Central tension, Microhardness, Crack resistance,
mg N possibility, N MPa μm MPa GPa N

%1 %0.0001

Raw – 501.51 ± 269.67 1.88 597.45 52.02 0.39 Air – – – 7.02 ± 0.26 1.08
Tin – – – 7.26 ± 0.48 0.68
375–12 0.67 ± 0.60 1314.80 ± 397.00 3.40 1470.80 380.05 25.26 Air 807.09 ± 51.23 4.96 ± 1.18 18.51 ± 10.79 7.35 ± 0.32 14.51
Tin 648.02 ± 42.83 4.59 ± 0.87 6.64 ± 0.45 7.30 ± 0.30 10.07
400–12 2.73 ± 0.79 1467.54 ± 474.05 3.14 1655.63 382.54 20.32 Air 653.02 ± 3.28 6.76 ± 0.05 5.45 ± 0.13 7.13 ± 0.28 13.58
Tin 621.56 ± 5.74 5.55 ± 0.25 4.58 ± 0.20 7.33 ± 0.36 3.76
425–12 5.00 ± 0.60 1532.52 ± 246.96 7.26 1692.95 898.37 252.45 Air 567.51 ± 9.39 12.24 ± 0.36 7.67 ± 0.16 7.73 ± 0.45 13.67
Tin 556.59 ± 10.87 11.40 ± 0.28 6.64 ± 0.07 7.65 ± 0.31 6.37
435–12 6.50 ± 1.00 1784.30 ± 249.46 5.83 1924.99 874.03 179.91 Air 544.15 ± 35.91 13.46 ± 1.56 8.52 ± 0.77 8.12 ± 0.39 12.95
Tin 494.32 ± 34.35 12.86 ± 0.98 6.83 ± 0.63 8.25 ± 0.18 12.95
450–12 7.42 ± 0.67 1307.69 ± 133.51 11.20 1362.85 903.93 397.12 Air 469.45 ± 27.35 17.89 ± 0.21 11.45 ± 4.18 7.57 ± 0.21 10.90
Tin 456.05 ± 6.72 14.97 ± 0.8 7.80 ± 0.12 7.36 ± 0.44 6.30
475–12 12.75 ± 0.97 1148.58 ± 391.49 3.90 1400.86 430.09 40.38 Air 373.07 ± 51.79 23.62 ± 1.57 11.24 ± 3.21 7.06 ± 0.26 5.51
Tin 373.10 ± 4.88 23.40 ± 0.54 10.39 ± 0.17 7.16 ± 0.47 2.32
425–4 3.00 ± 0.85 1543.00 ± 361.00 4.90 1682.62 658.06 100.34 Air 617.59 ± 6.94 6.82 ± 0.10 4.87 ± 0.11 7.35 ± 0.25 15.02
Tin 601.87 ± 15.92 5.06 ± 0.46 4.23 ± 0.58 7.77 ± 0.34 13.57
425–8 4.08 ± 0.52 1529.00 ± 447 4.51 1692.08 610.16 79.07 Air 595.26 ± 17.00 9.67 ± 0.32 6.41 ± 0.30 7.57 ± 0.86 11.76
Tin 565.59 ± 23.67 7.84 ± 0.82 5.67 ± 0.23 7.84 ± 0.51 8.66
435–4 3.33 ± 1.30 1493.30 ± 197.17 3.29 1672.38 413.14 25.10 Air 603.96 ± 19.26 7.16 ± 0.62 5.27 ± 0.16 7.77 ± 0.21 11.93
Tin 573.44 ± 13.41 5.38 ± 0.83 4.43 ± 0.26 7.50 ± 0.27 9.64
435–8 5.16 ± 0.93 2226.50 ± 249.46 10.63 2332.13 1512.90 635.79 Air 579.64 ± 13.97 10.88 ± 1.13 6.98 ± 0.37 7.59 ± 0.30 11.08
Tin 535.64 ± 32.99 9.63 ± 1.45 5.83 ± 0.51 7.16 ± 0.12 10.71
450–4 4.42 ± 0.67 1259.54 ± 370.89 3.33 1389.43 349.21 21.96 Air 502.22 ± 25.41 11.10 ± 1.97 5.80 ± 0.63 7.62 ± 0.24 4.76
Tin 498.50 ± 22.54 9.86 ± 1.54 5.77 ± 0.37 7.40 ± 0.28 3.75
450–8 6.50 ± 0.80 1357.84 ± 155.36 10.23 1415.35 902.60 366.54 Air 488.16 ± 11.86 14.69 ± 0.86 8.11 ± 0.03 7.58 ± 0.33 7.42
Tin 486.40 ± 18.36 13.55 ± 0.32 7.14 ± 0.04 7.19 ± 0.41 3.84
J Aust Ceram Soc

2800 2800
(a) (b)
2400 2400 425˚C
2000 2000 435˚C
Failure load, N

Failure load, N
450˚C
1600 1600

1200 1200

800 800

400 400

0 0
Raw 375 400 425 435 450 475 Raw 4 8 12
Temperature, ˚C Time, h
Fig. 3 Failure loads of raw and ion-exchanged glasses for a increasing temperature (for 12 h) and b increasing time

created [18, 35, 36]. Similarly, Varshneya [37] reported that was evaluated as the optimum temperature, since the fracture
elastic deformation occurs when a smaller ion is substituted by load of chemically strengthened glass showed an increase of
a larger ion, and subsequent plastic deformation occurs due to ~ 3.5 times (1784.30 N) that of raw glass (Fig. 3a).
the readjustment of non-bridging oxygens or displacement The results of the simultaneous examination of ion ex-
and bond bending of bridging oxygens. A total of 435 °C change temperature and the effect of time are given in Fig.

2 12
(a) (b)
1
10
m (Weibull Modulus)

0
8
lnln [1 / (1-F)]

-1
Raw 6
-2 375˚C
400˚C 4
-3 425˚C
435˚C
-4 2
450˚C
475˚C
-5 0
4.00 5.00 6.00 7.00 8.00 9.00 Raw 375 400 425 435 450 475
lnσ (N) Temperature, ˚C
2700
(c)
2400
2100
1800
σ0 (N)

1500
1200
900
600
300
0
Raw 375 400 425 435 450 475
Temperature, ˚C
Fig. 4 Effect of increasing temperature on a the two-parameter Weibull distribution fitted, b m, and c σ0 of ion-exchanged glasses
J Aust Ceram Soc

2
12
(a) (b)
1
10 425°C

m ( Weibull modulus)
435°C
0 450°C
8
lnln [1 / (1-F)]

-1 Raw
425-4 6
-2 425-8
425-12
435-4 4
-3 435-8
435-12
-4 450-4 2
450-8
450-12
-5 0
4.00 5.00 6.00 7.00 8.00 9.00 Raw 4 8 12
lnσ (N) Time, h

2700
(c)
2400 425°C
2100 435°C
450°C
1800
σ0 (N)

1500
1200
900
600
300
0
Raw 4 8 12
Time, h
Fig. 5 Effect of increasing process time on a the two-parameter Weibull distribution fitted, b m, and c σ0 of ion-exchanged glasses

2b. For 425 °C and 450 °C, it is understood that increasing the where F (σ, V) is the probability of fracture, V is the volume of
ion exchange time from 4 to 12 h did not cause a significant the material, V0 is the normalization volume, and σ is the
change in the fracture strength. But at 435 °C, while a rela- homogeneously applied stress. σ0 is a scaling parameter that
tively more pronounced effect was observed, ~ 2227 N, the is called the characteristic strength identified as the stress at
highest fracture load reached at 8 h of ion exchange. fracture probability of 63.2%. The Weibull modulus, m, is an
The effect of the ion exchange process on the standard indication of the strength distribution. Since the samples are
deviation values is not clear, as it provides a significant in- equal in size, V = V0 is assumed, and the exponential Weibull
crease in the strength of the glass. Additional defects and modulus was transformed into the linear form by taking the
irregular distribution of the ion-exchanged layer can be con- natural logarithm twice (Eq. 5):
sidered as the factors affect negatively to the standard devia-  
tion especially during the preparation of the samples. The 1
lnln ¼ lnV þ mlnðσ−σ0Þ−mlnσ0 ð5Þ
lowest standard deviation is obtained for glasses ion ex- 1− F
changed at 450 °C for 12 h.
Strength measurements were applied to 10 specimens for In such a procedure, a fracture probability (F) is required
each condition, and fracture load distributions were calculated for each test sample, and Eq. 6 is used:
using the conventional, two-parameter Weibull distribution
(Eq. 4): i−0:5
F¼ ð6Þ
     N
V σ m
F ðσ; V Þ ¼ 1−exp − ð4Þ
V0 σ0 where N is the total number of tested samples, that i is the
sample sequence in order of increasing stress [38].
J Aust Ceram Soc

When Fig. 4a–c and Table 2 are examined, it is seen that the close to the highest value. At the same time, it was determined
Weibull modulus (m) and the characteristic strength (σ0) that the optimum process condition was 435 °C-8 h because of
values of ion-exchanged glasses increased with the process both lower temperature and time requirements, and also, it has
temperature. No direct relationship was obtained between a higher characteristic fracture load. The highest σ0 value with
the ion exchange temperature and the Weibull modulus. a fracture load value of 2332.13 N is observed for the glass
When the characteristic strength values (σ0) are taken into treated for 435 °C-8 h.
consideration, the bending strength is similar to σ0 measure- The 425 °C-8 h and 435 °C-8 h samples of ion exchanged
ments. The Weibull modulus increased with temperature in- showed 63.2% a probability of fracture with σ0 values
crease from 375 to 425 °C, while it decreased at 435 °C, where 1692.08 and 2332.13 N, respectively. Also, the highest values
the maximum fracture load was obtained, and increased again were obtained under 450 °C-8 h and 435 °C-8 h process con-
at 450 °C (Fig. 3b). The highest Weibull module was reached ditions for 1% and 1 ppm load of failure probabilities.
at 450 °C as 11.20. Similarly, it has been understood that there In order to confirm the effect of the annealing (heating) of
is no relation between the fracture load and the Weibull mod- the raw glass on the mechanical properties, after the raw sam-
ule. These results give priority to the glass which ion- ples were only heated, the fracture load values were examined.
exchanged at 450 °C in terms of reproducibility. The charac- The heating process was performed at 435 °C-8 h, which was
teristic fracture load increased steadily with increasing tem- determined as the optimum process parameter. Only the heat-
perature, and after reaching the maximum point of ~ ed samples (without the presence of KNO3), the raw glass, and
1924.99 at 435 °C, it tended to fall at temperatures above the 435-8 samples are compared in Table 3. A significant rise
435 °C due to stress relaxation (Fig. 4c). With this aspect, it in the failure load values of samples of only the heating pro-
showed a similar behavior to the relationship between fracture cess was not observed. In addition, a slight decrease was cal-
load and temperature. Depending on the application area, it culated in the Weibull modulus value. As a result, no ion
may be desirable to determine under which load the specific exchange process was carried out, so a close value was ob-
rupture rates can be achieved [39]. From this point of view, the tained with the raw samples. Furthermore, heating samples are
fracture load values for 1% and 1 ppm fracture probabilities significantly lower than the mechanical properties of the ion-
are calculated (Table 2) and the highest results were obtained exchanged 100K-435-8 sample. For these reasons, only the
for ion-exchanged glass at 450 °C as 903.93 and 397.12 N, heating process was not performed at the other temperature
respectively, for 1% and 1 ppm fracture probabilities. parameters for raw glasses.
It is estimated that the mechanical strength values of
400 °C are scattered and distant from each other that shows Fractographic results and evaluation
the defect rate in these glasses is higher than those at other
temperatures. At 450 °C, it can be assumed that the defect It is known that there is a relationship between glass fracture
ratio is homogeneous and less than other temperatures, and behavior and glass strength. The fractured specimens were
this glass is reliable for certain strength value ranges. photographed and examined for fracture behavior after the
Moreover, these values are in good agreement with the stan- ring-on-ring test. The fracture fragments between the outer
dard deviations of strength measurements of ion-exchanged ring and the glass edges were counted as shown in Fig. 6
glasses [40]. and Table 2, and the relationship between the number of frag-
From Fig. 5b, when the temperature and the time were ments after the fracture test and the fracture strength was
evaluated together, the Weibull module increased with in- established.
creasing ion exchange temperature except for 435 °C-12 h Fragmentation is a dynamic process in which rapid crack
process conditions. It can be concluded that Weibull modulus branching or multiple crack initiation and propagation occurs
(m) were high due to the better filling of surface defects in during the fracture of a highly strained material [42]. During
glass (e.g., flaw, crack) by increasing ion exchange tempera- the bending, it was seen that in the central region where the
ture [41]. stress was applied, the fracture pattern called the inner ring
The highest Weibull modulus is obtained for 450 °C-12 h was formed, and the fracture pieces expanded towards the
process. The Weibull module of 435 °C-8 h process is also edges of the sample. Since the number of cracks is a function

Table 3 Raw, only heated raw samples, and 435-8 samples

Failure load m (Weibull modulus): σ0 (N)

Raw 501.51 ± 269.67 1.88 597.45


Only heated raw samples (435 °C-8 h) 580.20 ± 293.79 1.20 829.66
435-8 2226.50 ± 249.46 10.63 2332.13
J Aust Ceram Soc

Fig. 6 Images of fracture pattern (a) (b)


for ion-exchanged glass with dif-
ferent exchange times and tem- Number of
peratures. a Representative sam- Sample code
fractured part
ple. b Number of fracture part
Raw 42
375-12 65
400-12 78
425-12 82
435-12 97
450-12 67
475-12 69
425-4 95
425-8 78
435-4 71
435-8 112
450-4 65
450-8 63

of the magnitude of elastic energy that is released at the mo- pieces increased in parallel with the increase in strength so in
ment of fracturing, the total number of cracks is less in a glass accordance with the literature.
of low fracture strength. Conversely, the number of cracks
increases as more elastic energy is necessarily applied to crack Compressive stress and depths of stress layer results
the strengthened glasses [41, 42]. and evaluation
The highest number of broken pieces of glass showing the
highest fracture strength value was reached at 435 °C-8 h (112 Surface compressive stress (CS), depths of stress layer
fragments) (Fig. 6b). It is also apparent that the ion exchange (DOL), and center tensile stress (CT) on the air and tin
process carried out at a constant temperature of 450 °C in a sides resulting from ion exchange process were deter-
100% KNO3 salt bath does not have a significant effect on the mined by the FSM-6000 LE instrument. The measure-
number of fragments in a similar manner to that of the time. ments were taken from five different points for each of
The number of broken parts of the reference glass was also the air and tin surfaces of the samples, and the results are
determined to be lower than that of the glasses obtained from given in Table 2 and Fig. 7. In the process conditions,
the other ion exchange processes, similar to the mechanical during which the 100% KNO3 melt salt solution and the
strength value. 12 h ion exchange time were kept constant, it was seen
The role of internal central tension, which causes fragmen- that as the ion exchange temperature increased, the com-
tation, is well known for thermally tempered glasses, and in pressive stresses in the glass air and tin surfaces de-
these studies, it was assumed that internal central tension pre- creased, and the depths of stress layer increased (Fig.
dominates influence in fragmentation and its value was con- 7a, b). By increasing the ion exchange time and tempera-
nected to the fragment numbers [43, 44]. Hill and Donald [45] ture, the amount of K+ ions that were diffused to the glass
studied on ion exchange properties of a lithium aluminosili- surface continuously increased; thereby, the DOL was in-
cate glass, and they reported that the higher strength samples creased. Increasing the time and temperature of the ion
had smaller fragments after biaxial bending loading in frag- exchange process can generate stress relaxation and, thus,
mentation of the glass. Bouyne and Gaume [46] predicted that decreases compressive stress [28, 30, 48, 49]. The DOL
to make identical fragmentation levels, the tensile stress in of the samples is a minimum of 4.6 μm, and the CS
chemically tempered glass could be considerably less than in values range from ~ 373 to 807 MPa. The depth of the
thermally tempered glass. stress layer, with every 25 °C increase in temperature, was
In our study, similarly, the number of broken pieces in- found to provide regularly a ~ 6-μm depth increase on the
creased by increasing the flexural strength, and smaller pieces surface of the air. The general tendency was similar on tin
were obtained by branching. As a possible reason for this, the surfaces.
effect of the sum of the center tensile stress and the surface The depth of stress layer was always found to be larger on
compressive stresses were interpreted. It has a significant ef- the air side than the tin side. Since tin is known to increase the
fect on mechanical strength that the glass sample already had network connectivity of the glass, and presence of tin in glass
pre-existing defects and occurred during sample preparation blocked the K+ diffusion and, thus, it may cause a decrease in
[41, 42, 47]. All these results show that the number of broken depth of stress layer. It was also reported that compressive
J Aust Ceram Soc

Fig. 7 CS (compressive stress) 1000 30


Atm.
and DOL (depth of stress layer) (a) Atm.
Tin
(b) Tin
900
on air and tin sides of ion- 25

Compressive stress, MPa


exchanged specimens. a CS at 800
different exchange temperatures 20
700
for 12 h. b DOL at different ex-
change temperatures for 12 h. c 600 15
CS at 425–435–450 °C for dif-
ferent exchange times. d DOL at 500
10
425–435–450 °C for different 400
exchange times
5
300

200 0
375 400 425 435 450 475 375 400 425 435 450 475
Temperature ,˚C Temperature ,˚C

1000 30
Atm-425˚C
900
(c) Tin-425˚C (d) Atm-425˚C
Tin-425˚C
Atm-435˚C 25 Atm-435˚C
Compressive stress, MPa

Tin-435˚C Tin-435˚C
800 Atm-450˚C Atm-450˚C
Tin-450˚C 20 Tin-450˚C
700

600 15

500
10
400
5
300

200 0
4 8 12 4 8 12
Time, h Time, h

stress is generally greater on the tin side than on the air side for Microhardness results and evaluation
aluminosilicate (ALS) float glasses [28, 50]. Unlike this, our
results also show that the air side has greater compressive The results of the microhardness measurements are given in
stress than the tin side for soda-lime silica glass. It was report- Table 2, and the comparative graphs are shown in Fig. 8. After
ed that the tin is diffused from the surface towards the interior all ion exchange process conditions, the hardness of air sur-
by the ion exchange process. Due to the ion exchange process, faces of glasses was increased. In the ion exchange experi-
tin vacancies formed on the surface by diffusion of tin towards ments at different temperatures for the 12-h time, the hardness
to inner layers, and this may cause lower compressive stress values increased until 435 °C and then gradually decreased
than the air side. SLS glass has a lower density than ALS glass with increasing temperature for both surfaces (Fig. 8a). The
and has more non-bridging oxygen [51]. As a matter of fact, it maximum hardness value was obtained as 8.25 GPa with an
is thought that the different network structure of both glasses increase of ~ 18% with respect to the raw glass on the tin
can cause the air surface to have higher compressive stress surface at 12 h-435 °C process conditions (Fig. 8b). When
values than the tin surface. In the literature, the compressive examining the difference between air and tin surfaces, it was
stress behavior of SLS glasses on the air and tin surfaces has seen that the increase in the air surface is greater, but there is
not been examined up to now, and from this aspect, our study no clear relationship between the hardness values of the two
also carries its own authenticity. surfaces. Also, hardness values generally increased with in-
As to ensure a reliable strength, the depth of a compressive creasing ion exchange time on the air surfaces (Fig. 8c), but
layer must generally be greater than the size of typical surface more complicated results were obtained on the tin surfaces
flaws [36]. While there is no clear relationship between com- (Fig. 8d). It is noteworthy that ion exchange enhanced the
pressive stress and fracture load, it is considered that the opti- hardness.
mum depths of stress layer are about 9.5–11 mm depending on The non-bridging oxygen (NBO) size in glass is closely
fracture loads (for 425 °C-8 h and 435 °C-8 h). Because the associated with the contending shear and densification defor-
reproducibility of these experiments (425 °C-8 h and 435 °C- mation mechanisms, and SLS glass contains plenty of NBOs,
8 h) has been tested and similar fracture load values, CS, which results in increased subsurface shear damage as inden-
DOL, and CT have been obtained. tation load increases [22, 52].
J Aust Ceram Soc

Fig. 8 Effect of increasing ion 9 9


exchange temperature on (a) (b)
hardness; a air, b tin surfaces.
Effect of increasing ion exchange

Hardness, GPa

Hardness, GPa
time on hardness; c air, d tin 8 8
surfaces

7 7

6 6
Ref. 375 400 425 435 450 475 Ref. 375 400 425 435 450 475
Temperature, ˚C Temperature, ˚C

12.00 12.00
425°C (c) 425°C (d)
435°C 435°C
10.00 450°C
10.00
450°C
Hardness, GPa

Hardness, GPa
8.00 8.00

6.00 6.00

4.00 4.00

2.00 2.00

0.00 0.00
Raw 4 8 12 Raw 4 8 12
Time, h Time, h

When the effect of the glass composition on the hardness Another reason for the increase in hardness is the increase
was examined, it is reported that the hardness of the Na- in density because of the filling of larger K+ ions into the glass
containing glass is higher than the K-containing glass with a matrix. However, it is evident that ion exchange in 100%
similar composition [4, 32]. Filling potassium into the glass KNO3 molten bath enhances the resistance of our glass versus
surface with ion exchange produces a compressive stress radial/median crack formation during Vickers indentation.
against the penetration of the Vickers pyramidal tip and also The most important factors determining the strength of the
can increase the elastic recovery when the load is withdrawn glass are seen as crack initiation and crack propagation. It is
[9, 53]. For this reason, the increase of hardness in glass known that the glass exhibits high theoretical strength, and
strengthened with ion exchange is not due to the composition- when a crack occurs on the glass surface, the stress concen-
al change but related to the filling effect induced by the pass- tration at the crack tip accumulates, and the catastrophic frac-
ing of K+ ion in place of the Na+ ion that was caused by ture occurs at strengths much lower than the theoretical
increased compressive stresses and density. strength. Therefore, Bcrack resistance analysis^ is proposed
Kolluru et al. [54] reported that tin on the tin surface can as a method of evaluating crack initiation and crack propaga-
increase the network connectivity of the glass. This in- tion (BCracking resistance results and evaluation^ section)
crease in network connectivity will prevent the diffusion [56–58].
of K+ ions and reduce the depth of the stress layer [28]. It is
worthy to note that the hardness of the raw glass tin surface Cracking resistance results and evaluation
is higher than air side, and the hardness of the air surfaces
can be higher after ion exchange. As mentioned above, the In this section, in the interest of clear assessments on the
compressive stress on the air surface after ion exchange is hardness alterations among ion-exchanged samples, further
always greater than on the tin surface, and the compressive comparisons of crack probability were evaluated. Figure 9
stress can lead to a higher hardness (Fig. 7a, c). For this shows the cracking probabilities of ion-exchanged specimens
reason, the larger CS on the air surface may be one of the for air and tin surfaces, and the cracking resistance data are
reasons for the hardness being greater than the tin surface. given in Table 2. When ion-exchanged samples compared
On the other hand, perhaps one of the most significant with the raw glass at all temperature and time conditions, the
causes is the glass chemical composition differences be- possibility of cracking formation on both surfaces was deter-
tween the air and tin surfaces [50, 55]. mined to occur at higher loads and the crack resistance
J Aust Ceram Soc

100 100
(a) (b)
80 80

Cracking probability, %
Cracking probability, %

60 60

40 40 Raw
Raw
375˚C 375˚C
400˚C 400˚C
425˚C 20 425˚C
20 435˚C
435˚C
450˚C 450˚C
475˚C 475˚C
0 0
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Load, N Load, N
100 100
(c) (d)
Cracking probability, %

80

Cracking probability, %
80

60 60
Raw Raw
425-4 425-4
425-8 425-8
40 425-12 40 425-12
435-4 435-4
435-8 435-8
20 435-12 20 435-12
450-4 450-4
450-8 450-8
450-12 450-12
0 0
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Load, N Load, N
Fig. 9 Cracking behavior of ion-exchanged glasses after Vickers indentation-cracking possibilities (%) the effects of a temperature on air surface, b
temperature on the tin surface, c time (hour) on air surface, d time (hour) on the tin surface

increased. On the air surface, the cracking probabilities in- Indentation crack formation behavior depends on pre-
creased with increasing temperature, and a more complicated existing defects on the glass surface and residual stresses in
layout was exhibited on tin surfaces (Fig. 9a, b). There were the glass [22]. The purpose of compressive stress created on
no apparent differences in the microhardness results of the air the glass surface is to beat residual tensile stresses induced by
and tin surface (BMicrohardness results and evaluation^ sec- the Vickers indenter and to discontinue crack initiation and
tion), but the cracking resistance results clearly showed the propagation [60, 61]. As mentioned in previous studies, the
superiority of the air surface over the tin surface for all process ion-exchanged generated compression stress on the glass sur-
conditions (Table 2). face for all specimens’ crack resistance and therefore the crack
When we look at the process times, it is generally de- initiation loads were developed, which improves the mechan-
termined that the cracking probability decreased with in- ical behavior [18, 62].
creasing time, and correspondingly, the cracking resistance Kese et al. [15] and Koike et al. [32] investigated the effect
increased (Fig. 9c, d). An improvement of about 13.9 times of stress on hardness in soda-lime glass by hardness measure-
on the air surface of the chemical-tempered samples and ment during the bending strength test and deduced that stress
19.95 times on the tin surfaces was calculated with respect slightly affected hardness, and in their analysis results, the
to raw glass. hardness was not much improved by the compressive stress.
The Vickers indentation-cracking test is used to measure As a result, surface compression has a small effect on the
resistance to crack formation in a rigid configuration under measured crack resistance (and hardness) in the chemically
controlled/repeatable test conditions [59]. It has been reported tempered glass. As can be seen in other studies, it could be
that when K+ replaces Na+, the network connectivity increases evaluated that viscoelastic relaxation in the glass structure
due to the increase in density, which also supports network occurs near the glass transition temperature at high tempera-
rearrangement [32]. tures, which affects mechanical properties such as hardness
J Aust Ceram Soc

Fig. 10 AFM image of an ion exchange process. a Raw glass. b 435 °C-8 h

and cracking resistance [45, 63]. In addition, when we com- exchanged glasses was smooth and no buckling pattern was
pare our crack resistance values with our mechanical strength found. Ra values range from 0.28 to 1.18 nm.
(fracture load) values, it is observed that the difference ratio in An improvement in surface roughness was observed for all
cracking resistance between air and tin surfaces of our ion- process conditions (Figs. 11, 12). In the 12-h studies, the sur-
exchanged glasses is lower in high mechanical strength. This face roughness increased on air and tin surfaces of glass with
is an important point for both surfaces in terms of mechanical increasing temperature (Figs. 11a, 12a). In studies carried out
strength and cracking resistance analysis. at 425 °C and 435 °C with 4, 8, and 12-h waiting times, it was
However, there was no clear relationship among the com- observed that generally, the surface roughness increased with
pressive stress, hardness and cracking resistance values, and increasing waiting time (Fig. 11a, b). It was also observed
this requires further investigation. Generally, glass is impor- after ion exchange studies that the surface of the tin is always
tant for the development of the ion exchange process to create rougher than the air [64], because it is known from previous
a much better surface, since the micron and submicron flaws studies that tin increases surface roughness. Since our me-
(crack, defect) that occur during and after production signifi- chanical strength tests are made only on the air surface, it is
cantly affect the mechanical strength of the glass material. not possible to say exactly what the effect of surface rough-
Production conditions and glass composition are also impor- ness on mechanical strength is.
tant influences on ion exchange research. As the diameter of each potassium ion is larger than the
diameter of the sodium ion, pre-existing microcracks on the
glass surface are filled by potassium ions and a smoother
AFM (surface topography) results and explanation surface is obtained [65]. In this study, the ion exchange depths
of stress layer (DOL) range from 4.59 to 23.62 μm, and it is
Figure 10 illustrates the surface topography images of AFM to understood that potassium ions diffuse deeper from the sur-
raw and 435 °C-8 h process and Figs. 11 and 12 show the face with increasing temperature and time. Considering in this
AFM numerical results. The surface of the raw and ion- manner, it can be said that the increase in surface roughness

1.0 1.0
(a) (b) 425
435
0.8 0.8 450

0.6 0.6
Ra, nm

Ra, nm

0.4 0.4

0.2 0.2

0.0 0.0
Ref. 375 400 425 435 450 475 Ref. 4 8 12
Temperature, ˚C
Time, h
Fig. 11 On the surface roughness of the air surface. a The effect of the ion exchange temperature at a fixed time of 12 h. b The effect of the ion exchange
time at 425 °C, 435 °C, and 450 °C
J Aust Ceram Soc

1.5 1.5
(a) (b) 425
435
1.2 1.2 450

0.9 0.9
Ra, nm

Ra, nm
0.6 0.6

0.3 0.3

0 0.0
Ref. 375 400 425 435 450 475 Ref. 4 8 12
Temperature, ˚C Time, h
Fig. 12 On the surface roughness of the tin surface. a The effect of the ion exchange temperature at a fixed time of 12 h. b The effect of the ion exchange
time at 425 °C, 435 °C, and 450 °C

due to the increase in temperature may be caused by reason of a high-stress profile and compressive stress during ion ex-
the fact that the potassium ion is not deeply filled with the change [3, 18, 66].
defects (or microcracks) in the surface. At lower temperatures, Figure 13 shows the K/Na concentration profiles ratio on
the ion exchange process had a surface roughness reducing the the air and tin sides of the ion-exchanged glasses at 12 h for
effect, since potassium ions may remain on the surface layer different exchange temperature.
and can result in the better filling of surface defects The surface potassium concentration and diffusion depth
(microcracks). It should also be noted that the changes in increase with temperature. In all samples, while the potassium
surface roughness values of the ion exchange process are very ion concentration decreases continuously from the surface to
small ranges. the depth, the sodium ion concentration increases. The potas-
sium ion concentration on the air side is always greater than
K/Na concentration profile and explanation that on the tin side, and these are parallel to the depths of stress
layer (DOL) results. As the depth is progressed, irregular for-
Developing mechanical properties in glasses is directly related mations occur due to the glass structure and SEM-EDS point
to the amount of compressive stress on the surface and the analysis error margin (Fig. 13).
depth profile, which is linked to the ion concentration amount During ion exchange, it is known that structural relaxation
entering into the surface and the ionic diffusion depth [4, 11, occurs at high temperatures close to the glass transition tem-
18, 21]. In addition, the other important parameter is the effect perature and for an extended period of time, and a higher
combination of stress and structural relaxation in glasses. It is diffusion rate was expected with increasing temperature/time.
important to minimize the behavior of the glass while creating In the SEM-EDS K/Na ratio depth profile analysis, the

45 40
Air Tin
40 35
375-12 375-12
35 400-12 30 400-12
425-12 425-12
% K/Na rao

30
% K/Na rao

435-12 25 435-12
25 450-12 450-12
475-12 20 475-12
20
15
15
10 10

5 5
0 0
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
Depth, μm Depth, μm
Fig. 13 K/Na concentration profile for different temperature
J Aust Ceram Soc

Fig. 14 X-ray diffraction patterns 2000


of raw and 475 °C-12 h ion- Raw
exchanged glass samples 1800
1600 475-12

Intensity (counts)
1400
1200
1000
800
600
400
200
0
10 20 30 40 50 60 70
2 theta (degree)

diffused potassium ion to the sample increased with tempera- varied from 0.007 to 0.144%, respectively. The results of the
ture (coming out of Na ions increased) and progressed with ring-on-ring test showed that the mechanical strength was best
time and this is also shown in the weight increase measure- at 435 °C-8 h (4.4 times) and that the mechanical strength of
ments (Fig. 13). the chemical-tempered soda-lime silica glass decreased due to
With regard to the effect of the temperature, a deeper stress relaxation at high temperatures and long process times.
profile obtained at a higher surface potassium concentra- The statistical Weibull module was found to be a reliable glass
tion or at higher temperatures does not always mean better of chemical-tempered glass at 450 °C, but the highest fracture
mechanical performance [18, 30, 66]. This is due to the strength (63.2% fracture probability) was achieved at 435 °C-
glass relaxation behavior which overcomes the stress 8 h. As the mechanical strength increased, the accumulated
forming process at temperatures close to the glass transi- elastic energy in the chemically tempered glass increased; the
tion temperature during ion exchange [3, 4, 18]. Therefore, samples, consequently, were broken into too many pieces.
the very high diffusion rate remains insignificant and the Compressive stress decreased with increasing temperature,
concentration profile and the compressive stress profiles and the air side always had greater compressive stress than
do not coincide. the tin side for soda-lime silica glass. These results also im-
Moreover, XRD analyses were performed to verify that proved the mechanical strength of the glass. Chemical tem-
there was no crystallization at the raw and the highest temper- pering improved the hardness of float soda-lime silicate
ature ion-exchanged glass (475-12). XRD patterns of untreat- glasses to about 18%. For the chemically tempered glass sam-
ed glass, 475 °C for 12 h ion-exchanged glass are shown in ples, it was determined that the crack resistance increased by
Fig. 14. 13.9 times at the air side and 19.95 times at the tin side. The
The raw glass and ion-exchanged sample were determined AFM surface roughness increased with increasing tempera-
no crystalline phase, high-temperature ion-exchanged glass ture and time; lower values were also obtained according to
(475 °C-12 h) still being amorphous as the raw sample (Fig. raw glass. SEM-EDS concentration profile ratios showed that
14). the potassium ion concentration decreases continuously from
the surface to the depth, but the sodium ion concentration
increases.
Conclusion
Acknowledgments Many thanks to Bilal ALCAN for his contributions to
the experimental studies from the TUBITAK Marmara Research Center
The mechanical behavior of chemical-tempered soda-lime sil-
Materials Institute. The authors are thankful to ŞİŞECAM Science and
icate glass using the ring-on-ring test and Vickers indentation Technology Center.
has been investigated, and the relation between compressive
stress and depth of stress layer of mechanical behavior has Funding information The authors of this study are grateful to the
been evaluated. The weight of glass samples increased grad- Scientific & Technological Research Council of Turkey (TUBITAK)
for the financial support under the project entitled and numbered as
ually with time and temperature increase due to the inter-
BThird-Generation Glass Toughening: Ultra Strong Sheet Glass For
diffusion of K+–Na+ ions. For the 375 °C and 475 °C ion Large-Area Applications Through Super-Efficient Side-Selective Ion
exchange temperature ranges, the rate of weight increase Exchange-215M127.^
J Aust Ceram Soc

References 22. Gross, T.M., Price, J.J.: Vickers indentation cracking of ion-
exchanged glasses: quasi-static vs. dynamic contact. Front. Mater.
4, 4 (2017)
1. Ultra-Thin Glass Market - Forecast to 2022, Market Report. https://
23. Sklyarevich, V., Mykhaylo, S.: Method for the chemical strength-
www.researchandexperts.com/report/01L7-CH5730/ultra-thin-
ening of glass, U.S. Patent 9,505,654, issued November 29, 2016
glass-market-forecast-to-2022/ (2017). Accessed 22 December
2017 24. Talimian, A., Sglavo, V.M.: Ion-exchange strengthening of borosil-
icate glass: influence of salt impurities and treatment temperature. J.
2. Gy, R.: Ion exchange for glass strengthening. Mater. Sci. Eng. B.
Non-Cryst. Solids. 456, 12–21 (2017)
149(2), 159–165 (2008)
25. Pilkington, L.A.B.: Manufacture of flat glass. U.S. Patent No. 3,
3. Karlsson, S., Jonson, B., Stålhandske, C.: The technology of chem-
222,154. 7 Dec. 1965
ical glass strengthening–a review. Glass Technol. Eur. J. Glass Sci.
26. Pilkington, L.A.B.: Review lecture. the float glass process. Proc. R.
Technol. Part A. 51(2), 41–54 (2010)
Soc. London Ser. A Math. Phys. Sci. 314(1516), 1–25 (1969)
4. Varshneya, A.K.: Chemical strengthening of glass: lessons learned
27. Goodman, O., Derby, B.: The mechanical properties of float glass
and yet to be learned. Int. J. Appl. Glas. Sci. 1(2), 131–142 (2010)
surfaces measured by nanoindentation and acoustic microscopy.
5. (Anonymous): Corning’ Gorilla Glass brings King Kong glass Acta Mater. 59(4), 1790–1799 (2011)
strength to high-tech toys. Am. Ceram. Soc. Bull. 89, 41–42 (2010) 28. Jiang, L., Guo, X., Li, X., Li, L., Zhang, G., Yan, Y.: Different K+–
6. Geithe, A.: Touching allowed. Schott Technol. Mag. Sol. 28–29 Na+ inter-diffusion kinetics between the air side and tin side of an
(2011) ion-exchanged float aluminosilicate glass. Appl. Surf. Sci. 265,
7. Nippon Electric Glass has started mass-producing specialty glass 889–894 (2013)
for chemical strengthening, which is used for cover glass of smart 29. Shen, J., Green, D.J., Pantano, C.G.: Control of concentration pro-
phones and mobile devices, Press release. http://www.neg.co.jp/ files in two step ion exchanged glasses. Phys. Chem. Glasses. 44(4),
uploads/sites/2/news_20110418_en.pdf (2011). Accessed 24 April 284–292 (2003)
2018 30. Li, X., Jiang, L., Wang, Y., Mohagheghian, I., Dear, J.P., Li, L., Yan,
8. Jacoby, M.: New applications for glass emerge. Chem. Eng. News. Y.: Correlation between K+-Na+ diffusion coefficient and flexural
90(25), 34–36 (2012) strength of chemically tempered aluminosilicate glass. J. Non-
9. Varshneya, A.K.: Fundamentals of inorganic glasses. Elsevier Cryst. Solids. 471, 72–81 (2017)
(2013) 31. Wachtman, J.B., Roger Cannon, W., John Matthewson, M.:
10. Leboeuf, V., Blondeau, J.-P., De Sousa Meneses, D., Véron, O.: Mechanical properties of ceramics. John Wiley & Sons (2009)
Potassium ionic exchange in glasses for mechanical property im- 32. Koike, A., Akiba, S., Sakagami, T., Hayashi, K., Ito, S.: Difference
provement. J. Non-Cryst. Solids. 377, 60–65 (2013) of cracking behavior due to Vickers indentation between physically
11. Tyagi, V., Varshneya, A.K.: Measurement of progressive stress and chemically tempered glasses. J. Non-Cryst. Solids. 358(24),
buildup during ion exchange in alkali aluminosilicate glass. J. 3438–3444 (2012)
Non-Cryst. Solids. 238(3), 186–192 (1998) 33. Bartholomew, R.F., Garfinkel, H.M.: Chemical strengthening of
12. Garza-Méndez, F.J., Hinojosa-Rivera, M., Gómez, I., Sánchez, glass. Elast. Strength Glas. Glass Sci. Technol. 5, 217 (2012)
E.M.: Scaling properties of fracture surfaces on glass strengthened 34. Macrelli, G., Isoclima, S. P.A.: Glass chemical strengthening by ion
by ionic exchange. Appl. Surf. Sci. 254(5), 1471–1474 (2007) exchange, (2015)
13. Karlsson, Stefan, Sharafat Ali, and Michael Strand. Chemical 35. Ingram, M.D., Davidson, J.E., Coats, A.M., Kamitsos, E.I.,
strengthening of flat glass by vapour deposition and in-line alkali Kapoutsis, J.A.: Origins of anomalous mixed-alkali effects in ion-
metal ion exchange, (2014) exchanged glasses. Glass Sci. Technol. Glastechnische Berichte.
14. Karlsson, S., Jonson, B., Wondraczek, L.: Copper, silver, rubidium 73(4), 89–104 (2000)
and caesium ion exchange in soda–lime–silica float glass by direct 36. Donald, I.W.: Methods for improving the mechanical properties of
deposition and in line melting of salt pastes. Glass Techno. Eur. J. oxide glasses. J. Mater. Sci. 24(12), 4177–4208 (1989)
Glass Sci. Technol. Part A. 53(1), 1–7 (2012) 37. Varshneya, A.K.: Ion exchange: physical properties of ion-
15. Kese, K.O., Li, Z.C., Bergman, B.: Influence of residual stress on exchanged and melt-processed glasses differ. Glas. Res. 10, 21–
elastic modulus and hardness of soda-lime glass measured by nano- 26 (2001)
indentation. J. Mater. Res. 19, 3109–3119 (2004) 38. Weibull, W.: A statistical distribution function of wide applicability.
16. Yano, T., Nagano, T., Lee, J., Shibata, S., Yamane, M.: Cation site J. Appl. Mech. 18(3), 293–297 (1951)
occupation by Ag+/Na+ ion-exchange in R2O–Al2O3–SiO2 39. Le Bourhis, E.: Glass: mechanics and technology. John Wiley &
glasses. J. Non-Cryst. Solids. 270(1–3), 163–171 (2000) Sons (2014)
17. Braunger, M.L., Escanhoela, C.A., Ziemath, E.C.: Electrical con- 40. Kurkjian, C.R., Gupta, P.K., Brow, R.K.: The strength of silicate
ductivity of Ag–Na ion exchanged soda-lime glass. Solid State glasses: what do we know, what do we need to know? Int. J. Appl.
Ionics. 265, 55–60 (2014) Glas. Sci. 1(1), 27–37 (2010)
18. Erdem, I., Guldiren, D., Aydin, S.: Chemical tempering of soda lime 41. Pepi, J.W.: Strength properties of glass and ceramics. Spie Press
silicate glasses by ion exchange process for the improvement of (2014)
surface and bulk mechanical strength. J. Non-Cryst. Solids. 473, 42. Tandon, R., Jill Glass, S.: Fracture initiation and fragmentation in
170–178 (2017) chemically tempered glass. J. Eur. Ceram. Soc. 35(1), 285–295
19. Haldimann, Matthias, Fracture strength of structural glass elements, (2015)
(2006) 43. Barsom, J.M.: Fracture of tempered glass. J. Am. Ceram. Soc.
51(2), 75–78 (1968)
20. Shelby, J.E.: Introduction to glass science and technology. Royal
Society of Chemistry (2005) 44. Akeyoshi, K., Kanai, E.: Mechanical properties of tempered glass.
In: Proceedings of the 7th International Glass Congress, vol. 14, pp.
21. Jannotti, P., Subhash, G., Ifju, P., Kreski, P.K., Varshneya, A.K.:
80–85 (1965)
Influence of ultra-high residual compressive stress on the static
45. Hill, M.J.C., Donald, I.W.: Stress profile characteristics and me-
and dynamic indentation response of a chemically strengthened
chanical behaviour of chemically strengthened lithium magnesium
glass. J. Eur. Ceram. Soc. 32(8), 1551–1559 (2012)
aluminosilicate glasses. Glass Technol. 30(4), 123–127 (1989)
J Aust Ceram Soc

46. Bouyne, E., Gaume, O.: Fragmentation of thin chemically tempered 58. Wada, M., Furukawa, H., Fujita, K.: Crack resistance of glass on
glass plates. Glass Technol. C. 43, 300 (2002) Vickers indentation. In: Proc. Int. Congr. Glass 10th, vol. 11, pp.
47. Quinn, G.D., Quinn, G.D.: Fractography of ceramics and glasses. 39–46 (1974)
National Institute of Standards and Technology, Washington, DC 59. Price, J.J., Scott Glaesemann, G., Clark, D.A., Gross, T.M.,
(2007) Barefoot, K.L.: 69.3: a mechanics framework for ion-exchanged
48. Varshneya, A.K.: The physics of chemical strengthening of glass: cover glass with a deep compression layer. In: SID Symposium
room for a new view. J. Non-Cryst. Solids. 356(44–49), 2289–2294 Digest of Technical Papers, Vol. 40, No. 1, pp. 1049–1051.
(2010) Blackwell Publishing Ltd (2009)
49. Shen, J., Green, D.J.: Prediction of stress profiles in ion exchanged 60. Tandon, R., Green, D.J., Cook, R.F.: Surface stress effects on in-
glasses. J. Non-Cryst. Solids. 344(1–2), 79–87 (2004) dentation fracture sequences. J. Am. Ceram. Soc. 73(9), 2619–2627
50. Li, X., Jiang, L., Zhang, X., Yan, Y.: Influence of residual compres- (1990)
sive stress on nanoindentation response of ion-exchanged alumino- 61. Morris, D.J., Myers, S.B., Cook, R.F.: Indentation crack initiation in
silicate float glass on air and tin sides. J. Non-Cryst. Solids. 385, 1– ion-exchanged aluminosilicate glass. J. Mater. Sci. 39(7), 2399–
8 (2014) 2410 (2004)
51. Ziemath, E.C., Saggioro, B.Z., Fossa, J.S.: Physical properties of 62. Sane, A.Y., Cooper, A.R.: Stress buildup and relaxation during ion
silicate glasses doped with SnO2. J. Non-Cryst. Solids. 351(52–54), exchange strengthening of glass. J. Am. Ceram. Soc. 70(2), 86–89
3870–3878 (2005) (1987)
52. Hagan, J.T., Van Der Zwaag, S.: Plastic processes in a range of
63. Shaisha, E.E., Cooper, A.R.: Residual stress in singly and doubly
soda-lime-silica glasses. J. Non-Cryst. Solids. 64(1–2), 249–268
ion-exchanged glass. J. Am. Ceram. Soc. 64(1), 34–36 (1981)
(1984)
53. Talimian, A., Sglavo, V.M.: Can annealing improve the chemical 64. Moseler, D., Heide, G., Frischat, G.H.: Atomic force microscope
strengthening of thin borosilicate glass? J. Non-Cryst. Solids. 465, study of the topography of float glasses and a model to explain the
1–7 (2017) bloom effect. N. p., Germany (2002) Web
54. Kolluru, P.V., Green, D.J., Pantano, C.G., Muhlstein, C.L.: Effects 65. Wang, R.-B.: Method for strengthening glass substrate and article
of surface chemistry on the nanomechanical properties of commer- manufactured by the same., U.S. Patent Application 13/928,519,
cial float glass. J. Am. Ceram. Soc. 93(3), 838–847 (2010) filed July 3, 2014
55. Hand, R.J., Tadjiev, D.R.: Mechanical properties of silicate glasses 66. Shen, J., et al.: Stress relaxation of a soda lime silicate glass below
as a function of composition. J. Non-Cryst. Solids. 356(44–49), the glass transition temperature. J. Non-Cryst. Solids. 324(3), 277–
2417–2423 (2010) 288 (2003)
56. Kato, Y., Yamazaki, H., Yoshida, S., Matsuoka, J.: Effect of densi-
fication on crack initiation under Vickers indentation test. J. Non- Publisher’s note Springer Nature remains neutral with regard to
Cryst. Solids. 356(35–36), 1768–1773 (2010) jurisdictional claims in published maps and institutional affiliations.
57. Yoshida, S., Hidaka, A., Matsuoka, J.: Crack initiation behavior of
sodium aluminosilicate glasses. J. Non-Cryst. Solids. 344(1–2), 37–
43 (2004)

View publication stats

You might also like