You are on page 1of 11

Journal of Non-Crystalline Solids 358 (2012) 1897–1907

Contents lists available at SciVerse ScienceDirect

Journal of Non-Crystalline Solids


journal homepage: www.elsevier.com/ locate/ jnoncrysol

New observations on scratch deformations of soda lime silica glass


Payel Bandyopadhyay, Arjun Dey 1, Ashoke K. Mandal, Nitai Dey, Anoop K. Mukhopadhyay ⁎
CSIR-Central Glass and Ceramic Research Institute, Kolkata-700032, India

a r t i c l e i n f o a b s t r a c t

Article history: In spite of the wealth of literature, the role of the scratching speed in affecting the material removal mecha-
Received 21 March 2012 nism in soda lime silica (SLS) glass is yet to be comprehensively understood. Here we report the surface and
Received in revised form 18 May 2012 sub‐surface deformation mechanisms of SLS glass scratched under three different normal loads of 5, 10 and
Available online 16 June 2012
15 N at various speeds in the range of 100–1000 μm/s with a diamond indenter of ~ 200 μm tip radius. The
results show that at any given applied normal load, the width, depth, wear volume of the scratch grooves
Keywords:
Soda lime silica glass;
and wear rate of the SLS glass decreased with an inverse power law dependence on the applied scratching
Scratch; speed. The surface damage also reduced with the increase in scratching speed. A new, simple model was de-
Scratching speed; veloped to explain these observations. The significant contributions of the time of contact, the tensile stress
Deformation behind the indenter and the shear stress active just underneath the indenter in governing the material re-
moval mechanisms of the SLS glass were discussed.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction differences between the damage zones, glass composition, chemical


composition of lubricating liquid and presence or absence of water
The scratch tests provide a simple means to understand the defor- vapor [10–21]. The scratching speed, the angle, the radius as well as
mation and material removal mechanisms during dynamic contact in the shape of cutter tip along with the applied normal load affect the crit-
metals [1], polymers [2] and brittle solids like glass and ceramics ical condition for growth [22] of various sub-surface cracks in a wide va-
[3–7]. Of all these materials glass is the most brittle one. But today, riety of glasses when measured by different techniques [23–25]. We
the advanced applications of glass span from microwave safe crockery have recently reported the influence of normal load on the material re-
to mobile phones to laptops to advanced bio-glass for bone healing moval mechanisms and the degradation of the nanomechanical proper-
and repair to tablet screens to space shuttle components to lenses for ties inside the scratch groove of a soda lime silica (SLS) glass [26–29].
next generation radio telescopes. All these applications demand nano- Very recently [29] we reported the experimental observation of the in-
meter or even sub-nanometer scale surface finish of the glass based verse power law dependencies of width, depth, wear volume of scratch
components. That in turn demands very precise grinding and polishing grooves and the corresponding wear rate of SLS glass on the scratching
of glass which is not possible without a very clear cut picture of the speed (v) when the applied normal load was kept constant at a low
material removal mechanisms and their dependencies on the glass value of 5 N. Extending our previous work [29] here we report the de-
compositions as well machining parameters [6,7]. The single, double velopment of a new, simple model that explains the scratch behavior
or multiple pass scratch tests are therefore conducted to simulate the of SLS glass studied at three different normal loads when it is scratched
material removal mechanisms for a truly wide variety of glass, ceramics, with a blunt diamond indenter of ~200 μm tip radius at three different
composites, thin films and coatings [1–14]. Depending on load both de- scratching speeds.
formation and micro‐fracture can happen during scratching of glass [8]
and ceramics [9]. The other important factors for grinding and polishing 2. Materials and methods
of glass are force ratio, specific energy, critical load for Hertzian tensile
crack initiation, scratching speed, transition in material removal mech- The scratch tests were conducted on 25 × 25 × 1.40 mm polished
anisms, interaction among neighboring scratches and the characteristic soda lime silica glass slides (Blue Star, Kolkata, India) in air under
ambient (30° C) laboratory conditions at three constant applied nor-
mal loads (P) of 5, 10 and 15 N. A constant relative humidity of 30%
⁎ Corresponding author at: Central Glass and Ceramic Research Institute, 196, Raja was maintained during all the scratch tests in the laboratory. So, we
S.C. Mullick Road, Kolkata-32, India. Tel.: + 91 33 2473 3469/76/77/96; fax: + 91 33 expect the effect of humidity, if any, will be constant and hence was
2473 0957. not specifically taken into account while discussing the results. The
E-mail addresses: anoopmukherjee@cgcri.res.in, mukhopadhyay.anoop@gmail.com
(A.K. Mukhopadhyay).
temperature changes during the present scratch experiments were
1
Present Address: Thermal Systems Group, ISRO Satellite Centre, Vimanapura, Post, not measured. To the best of our knowledge there is no reported
Bangalore 560017, India. data on experimentally measured temperature data collected during

0022-3093/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.jnoncrysol.2012.05.041
1898 P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907

the actual experiments. Even if the same were available, the reliability
of the same could be questionable simply because it must have had to
be measured from a reasonable distance below the scratch groove by
insertion of an appropriate sensor e.g. a thermocouple. On the other
hand it is well known that the soda lime silica glass has a strain
point of ~ 511° C, annealing point of ~548° C and a softening point of
~ 715° C [30]. Therefore, it is expected that unless there is a sufficient
rise of local temperature there is little chance for visco-elastic defor-
mation to play any role in scratch damage evolution of the present
soda lime silica glass. According to 3D Finite Element Model of Scratch
temperature distribution, if a brass bonded diamond tool of 10 μm
size diamond grit is used to grind vitreous bonded silica at a grinding
speed of 20 m/s under 0.2 N load with a depth of cut of ~ 1 mm when
the friction coefficient is 0.2, the average temperature is ~128° C at
the contact region. Moreover, in the case of air cooling only 3.42% of
the total grinding energy enters the glass substrate while the rest
goes to the tool. Although no forced air cooling was done in the pre-
sent case and the glass is a soda lime silica glass based on the litera- Fig. 1. Schematic illustration of the bonded interface technique to estimate the sub-
ture data a conservative estimate of 130° C can be assumed for the surface deformation.
contact zone temperature [31]. On the other hand, according to
Tomozawa et al. the fictive temperature on the fracture surface of
silica glass was measured to be nearly 1120° C [32]. Further, it has grinding wheel was used to make both the top and bottom surfaces flat
been shown that in soda lime silicate glass the microhardness mea- and parallel to each other. These two surfaces were actually perpendic-
sured by a sharp contact (~ 1.5 μm tip radius) in the close vicinity of ular to the bonded interface. Next the top surface was polished down to
a large indentation was lower than that evaluated at region far re- 0.25 μm diamond paste to achieve a mirror finish. Then the assembly is
moved from the same large indentation due to local increase in fictive taken out of the polymer back-up material. This makes the bonded in-
temperature (~1050–1250° C) that degraded the local hardness [33]. terface ready for further experimentation as shown in Fig. 1. Now
Thus from the literature data it is really difficult to make any certain when a scratch is made on the top surface of this bonded interface,
guess about the temperature in the scratch groove of the present the sub-surface comes on the larger rectangular polished area part as
SLS glass. In the absence of comprehensive experimental data there- depicted in Fig. 1. After conducting the desired scratch experiments on
fore it may be plausible to suggest that the temperature in the contact the bonded interfaces at pre-determined values of loads and speeds,
zone between the blunt (~200 μm tip radius) indenter and the SLS the adhesive was dissolved in acetone and the sample surfaces were uti-
glass during scratching was ~130° C [31]. If this suggestion is correct lized for the experimental observations of the sub-surface deformation
then such a temperature would be too low in comparison to strain and damage processes. The damage evolution processes during the
point of ~511°C, annealing point of ~548° C and a softening point of scratch tests in both surface and sub-surface regions were studied
~715° C for the present SLS glass to deform by visco-elastic deformation. using optical (GX51, Olympus, USA) and field-emission scanning elec-
On the other hand if the visco-elastic deformation is absent it is highly tron microscopy (FESEM, Supra VP35, Carl Zeiss, Germany).
unlikely that any physical condition or the chemical composition
would change during the scratch experiments. Therefore, it appears rea- 3. Results
sonable to assume that such factors did not influence the present results
which depict new observations in the evolution of velocity dependent For any given applied load, the scratch width, w (Fig. 2(a)), depth, d
damage phenomena during scratching of the SLS glass. In each case (Fig. 2(b)) wear volume, Wv (Fig. 2(c)) and wear rate, WR (Fig. 2(d)) all
the scratching speeds (v) were 100, 500 and 1000 μm/s. A scratch tester had inverse power law dependencies on the scratching speed. The
(Model TR-102-M3, Ducom, Bangalore, India) equipped with a Rock- corresponding low magnification optical photomicrographs are
well C diamond indenter with ~200 μm tip radius was used for this pur- shown in Fig. 3. It is evident that for any given applied load, the
pose. Further details may be found in [26–29]. A compound force scratch width was highest (Fig. 3(a), (c), (e)) at lowest scratching
transducer is coupled with the scratch tester which can measure both speed (e.g. 100 μm/s) and lowest (Fig. 3(b), (d), (f)) at the highest
normal load (P) and the tangential force (F). It had a load cell with a res- speed (e.g. 1000 μm/s). The number of Hertzian tensile cracks and
olution of ±0.01 N. The stroke length for all scratches was 3 mm. The the microdamage inside the scratch grooves reduced with scratching
scratch offset was set at 0.5 mm for all the scratches. The values of the speed (Fig. 3). The corresponding SEM [26,27,29] photomicrographs
coefficient of friction (μ) were estimated using the relation μ = P/F. for an applied load of 5 N showed the genesis and the mode of prop-
The depth and width data of the scratches was measured by a pro- agation of the edge cracks in tracks made at 100 (Fig. 4(a)) and the
filometer (Form Talysurf 120, Taylor Hobson, UK). The sub-surface de- damage in the scratch track much less for 1000 μm/s (Fig. 4(b))
formation was studied using the well known bonded interface scratching speeds. The interactions among the Hertzian tensile
technique [34,35] as shown in Fig. 1. For this purpose two pieces of pro- cracks as well as the damage inside the scratch grooves were lower for
spective soda lime silica glass samples were individually prepared by the higher velocities. At higher applied loads of 10 (Fig. 4(c), (d)) and
polishing down to 0.25 μm diamond paste surface finish on their larger 15 N (Fig. 4(e), (f)) micro-chip formation, as well as micro-wear debris
rectangular area sides as shown in Fig. 1. These opposing surfaces were formation occurred and the degree of their occurrences were more sig-
then bonded together using a thin layer of a commercially available nificant at lower scratching speeds (e.g. 100 μm/s, Fig. 4(c), (e)) than at
common adhesive that is dissolvable in acetone. The bonding of the op- higher scratching speeds (e.g. 1000 μm/s, Fig. 4(d), (f)). To have a better
posite polished faces was thus achieved by placing the two of them face view of the surface deformation features in depth the FE-SEM photomi-
to face with each other under a small clamping pressure using a small crographs for surface deformations of scratch grooves created at speeds
vise mounted on a laboratory work table. Then this assembly was of 100 and 1000 μm/s are shown in order in Fig. 5(a)–(f) for applied nor-
mounted in a steel mould using commercially available epoxy and hard- mal loads of 5, 10 and 15 N respectively. These higher magnification
ener as is the standard practice. After removal from the mold the assem- photomicrographs show a general feature that the areas entrapped
bly was cured for 24 h at room temperature. Initially, a 10 μm diamond in between the Hertzian tensile ring cracks were heavily plastically
P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907 1899

Fig. 2. Inverse power law dependencies of (a) width (w), (b) depth (d), (c) wear volume (WV) of the scratch groove in the SLS glass and (d) wear rate (WR) on the scratching speeds (v).

deformed showing signatures of extensive shear induced deformations little lifted out. At higher load a significant role of local comminution pro-
and micro‐fractures, Fig. 5(c)–(f). The photomicrographs in Fig. 5(a), cess of the thin broken out pieces (i.e. micro-wear chips) of surface re-
(b) showed that for a given applied normal load of 5 N, apparently the gion was also evident, Fig. 5(e), because this process leads to the
initiation of edge cracks from the existing tensile ring cracks were formation of the micro‐wear debris. At a higher scratching speed of
more evident in the case of scratch made at 100 μm/s (Fig. 5(a)) than 1000 μm/s the additional feature present was that (Fig. 5(f)) the region
at 1000 μm/s (Fig. 5(b)). However, at higher scratching speeds more of near the boundary of the Hertzian tensile crack had suffered extensive
intra-tensile ring cracks were evident (Fig. 5(b)). At higher applied nor- local micro fracture and comminution of the glass particles.
mal load e.g. 10 N an additional feature of a large number of shear bands Fig. 6(a)–(f) shows the low magnification FESEM photomicro-
were present in the surface of the scratch grooves especially in the re- graphs of the sub-surface region of the scratch grooves created at 5
gion between the Hertzian tensile ring cracks (Fig. 5(c)). Interestingly, (Fig. 6(a), (b)), 10 (Fig. 6(c), (d)) and 15 N (Fig. 6(e), (f)) at 100
at lower scratching speed of 100 μm/s the shear bands were mutually and 500 μm/s respectively. There were two regions in the sub-
perpendicular to each other, Fig. 5(c). However at a higher scratching surface damage zone, the main scratch grooves and the damage
speed of 1000 μm/s the shear bands were oriented at small angles area beyond the scratch groove. All the photomicrographs generally
with respect to the direction of scratching and also with respect to showed that the main scratch grooves were full of crushed material
each other, Fig. 5(d). The extent of micro-cracking was more at lower and large numbers of shear band are clearly visible beyond the
scratching speed (Fig. 5(c)) than at higher scratching speed (Fig. 5(d)). main scratch grooves. In addition to shear deformation, microcracks
Generally, the shear deformation band formation occurred within the were also visible for the higher scratching speeds. The corresponding
space bounded the two consecutive Hetrzian tensile ring cracks but high magnification photomicrographs are shown in Fig. 7. These pho-
did not cross the boundary of the ring cracks, Fig. 5(d). So it appeared tomicrographs (Fig. 7) showed two generic features in the sub-
as if the shear deformation zones were arrested at the boundary of the surface deformation of the present SLS glass. Firstly, there was huge
Hertzian tensile ring cracks (Fig. 5(d)–(f)). At still higher load of 15 N shear deformation active at multiple planes in the sub-surface regions
the fracture events are more than the shear deformation (Fig. 5(e), of the scratch grooves wherein the shear induced deformation and/or
(f)). At a lower scratching speed of 100 μm/s (Fig. 5(e)) now there micro‐fracture regions were oriented in multiple directions including
were several pieces of thin glass slabs in the surface of the scratch groove. occasionally orthogonal ones. At a given applied normal load, this had
These thin portions appeared to be almost sheared out but still attached happened for all scratching speeds (Fig. 7(a), (b)). The sub-surface
to the bulk material. In these sheared out portions there were clear pres- shear induced damage signatures were more prominent at a higher
ence of shear deformation lines nearly parallel to the boundary of the scratching speed than at a lower scratching speed. Similarly, for a
Hertzian tensile crack. A large number of nearly parallel shear deforma- given scratching speed, the occurrence of similar phenomenon
tion bands intersected them nearly orthogonally. When one of two such could be noticed for all applied normal loads (Fig. 7(a), (c), (e)).
heavily sheared portions tried to come over the other the registry was But, the damage evolution was more acute at a higher applied normal
not perfect and such regions apparently became favourite zones for load than at a lower applied normal load. Secondly, shear induced
shear induced micro-crack formation probably during the unloading micro-cracks generated in the sub-surface region of the SLS glass
cycle. Near the outer periphery of the Hetrzian tensile ring crack many (Fig. 7(a), (d)). These micro-cracks lead to micro-wear chip formation
microfactured regions were evident (Fig. 5(e)) which appeared to be a when the portion gets detached from all sides and further comminution
1900 P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907

Fig. 3. Typical lower magnification optical photomicrographs of the scratch grooves respectively at the lowest (100 μm/s ) and highest (1000 μm/s) scratching speeds for applied
normal loads of 5 (a, b), 10 (c, d) and 15 N (e, f) in correspondence. Black solid arrow in (a) indicates scratching direction for all cases.

of such micro-wear chips lead to formation of micro and/or nano-wear the lower is the time of contact between the scratching indenter
debris formation (Fig. 7(a) and (c)–(f), [26–29]). and the glass surface, i.e.

 
4. Discussions 1
t∞ ð2Þ
v
The inverse power law dependencies of width, w (Fig. 2(a)),
depth, d (Fig. 2(b)), wear volume, WV (Fig. 2(c)) and wear rate, WR,
(Fig. 2(d)) can be explained by a simple model presented here. Let It also is evident that the grand probability ζ for occurrence of all
the events will be proportional to “t”, the time of contact. So, we have:
the number of events (e.g. nanoscale plasticity, or physical damage
evolution or micro-cracks) induced by the scratching experiment at  
an applied normal load, P be N. If the probability of a single event N 1
h ∞t∞ ð3Þ
be “h”, then according to the rule of probability [36], the grand prob- v
ability ζ of all the N events occurring simultaneously but independent
of each other, is given by: which gives,

N −ðα Þ
ζ ¼ ∏ hi ¼ h
N
ð1Þ h∞v ð4Þ
i¼1
where, α = 1/N. But both scratch width (w) and scratch depth (d)
Let the time of contact between the scratching indenter and the are proportional to “h” because the more the probability of occur-
SLS glass be “t”. It follows that the higher the scratching speed (v) rence of the micro-cracks and their interaction, the higher will be
P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907 1901

Fig. 4. Typical lower magnification SEM photomicrographs of the surface of the scratch grooves respectively at the lowest (100 μm/s) and highest (1000 μm/s) scratching speeds for
applied normal loads of 5 (a, b), 10 (c, d) and 15 N (e, f) in correspondence. Black solid arrow in (a) indicates scratching direction for all cases. The hollow white, yellow, red, black
arrows indicate respectively the Hertzian tensile cracks, the intra Hertzian tensile crack, micro/nano wear debris, edge cracks in the scratch track and the black dotted circle indi-
cates the interaction and hence most probable zone of chipping.

the width (w) and depth (d) of the scratch grooves. In other words, which means that:
we can write in a generalized sense:
−ðγ Þ
  d ¼ Bv ð10Þ
N 1
w ∞t ∞ ð5Þ
v
where B is the pre-exponential factor and γ is the power law expo-
nent. Further, it must be recognized that here both β = N1− 1 and
or,
γ = N2− 1 are constants similar to α. Here, N1 stands for the sum
total effect of the number of micro-cracks on the surface of the
−ðβ Þ
w∞ v ð6Þ scratch groove and their interactions while N2 stands for the sum
total effect of the number of micro-cracks in the sub-surface of the
which means scratch groove and their interactions. Since both N1 and N2 are inte-
gers, both β and γ will have fractional values. Further, since both
w ¼ Av
−ðβÞ
ð7Þ depth and width of the scratch groove exhibit inverse power law de-
pendencies on the scratching speed (v) from Eqs. (7) and (9) above;
assuming the groove cross‐section as approximately triangular, the
where, A is the pre-exponential factor and β is the power law expo-
wear volume of the scratch groove (WV) of length (l) can be approx-
nent. Similarly, we can write:
imately written as:
 
N 1
d ∞t ∞ ð8Þ WV e 0:5wld ð11Þ
v

or, It follows from Eqs. (6) and (9) that

−ðγ Þ −δ
d∞ v ð9Þ WV ∞ ðvÞ ð12Þ
1902 P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907

Fig. 5. Typical high magnification, high resolution FESEM photomicrographs of the surface of the scratch grooves respectively at the lowest (100 μm/s) and highest (1000 μm/s)
scratching speeds for applied normal loads of 5 (a, b), 10 (c, d) and 15 N (e, f) in correspondence. The hollow black, white and the red arrows indicate respectively the micro-
cracks, the shear bands and the micro/nano wear debris in the scratch track.

which means that: γ. Further, the wear rate WR for a given applied normal load P,
scratching speed v and scratch length (l) is given by:
−δ
WV ¼ C ðvÞ ð13Þ

where C is the pre-exponential factor and δ is the power exponent Wv


WR ¼ ð14Þ
which should have a fractional value as explained before for β and P:l
P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907 1903

Fig. 6. Typical low magnification, high resolution FESEM photomicrographs of the sub-surface region of the scratch grooves respectively at the lowest (100 μm/s) and intermediate
(500 μm/s) scratching speeds for applied normal loads of 5 (a, b), 10 (c, d) and 15 N (e, f) in correspondence. The hollow black and white arrows indicate respectively the shear
bands, micro-cracks and the red dotted lines indicate the main scratch groove.

Therefore, it follows from Eq. (12) that which means that:


−ε −ε
WR ∞ ðvÞ ð15Þ WR ¼ DðvÞ ð16Þ

Fig. 7. Typical high magnification, high resolution FESEM photomicrographs of the sub-surface region of the scratch grooves respectively at the lowest (100 μm/s) and intermediate
(500 μm/s) scratching speeds for applied normal loads of 5 (a, b), 10 (c, d) and 15 N (e, f) in correspondence. The hollow black, white and the red arrows indicate respectively the
micro-cracks, the shear bands and the micro/nano wear debris in the sub-surface region of the scratch grooves.
1904 P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907

Here, D is the pre-exponential factor and ε is the power law expo- of sub-surface defects whose influence therefore fail to be exhibited as
nent which should have a fractional value as explained before for β, γ prominently as in the case of the low load (e.g. 5 N) scratch situation.
and δ; because for a given experiment “P” and “l” are constants. It may As far as the pre-exponential factors are concerned the data pres-
be noted that both the Eqs. (11) and (13) predict inverse power law ented in Table 1 show certain interesting features. For instance for a
dependencies of WV and WR on the scratching speed “v”. The trends given wear parameter e.g. width, depth, wear volume and wear
of the data presented in Fig. 2 matched with those predicted by rate, the pre-exponential factor always increased with load with a
Eqs. (7), (9), (11) and (13) proposed in the current model. Thus, steep gradient. This trend was opposite to what was found for the
the present work not only matches with those reported in [15,17] power law exponents as discussed above. On the other hand for a
but also provides a model which can explain the present and earlier given applied normal load, pre-exponential factor always decreased
[15,17] observations for glasses of different varieties as far as the de- systematically ranging from scratch width to scratch depth to wear
pendencies of scratch parameters on the scratching speeds are con- volume with a steep gradient (Table 1). This trend was also just the
cerned. The values of all the pre-exponential factors and the opposite of what was found for the power law exponents as men-
corresponding exponents as obtained by fitting of the present exper- tioned above.
imental data to the corresponding Eqs. (7), (10), (16), and (13) are These data were also consistent with the corresponding low mag-
given in Table 1 as a ready reference for the various combinations of nification optical photomicrographs (Fig. 3(a)–(f)). The critical load
the applied normal loads (P) and the scratching speeds (v). The requirement for Hertzian tensile cracks formation calculated follow-
data should be discussed in view of the fact that at lower speed ing [16,37] and using appropriate experimental data [26–29] was typ-
there is more “time” to meet “defects” in the brittle glass material. ically low at 0.0012–0.012 N. But yet the number of Hertzian tensile
The data presented in Table 1 prove that for a given wear parameter cracks reduced with scratching speed (Fig. 3). This happened because
e.g. width, depth, wear volume and wear rate, the power law exponent the time of contact (e.g. ca. 30 s at 100 μm/s versus 3 s at 1000 μm/s)
always decreased with load with a steep gradient. This data therefore also reduces with the scratching speed [29] following the inverse power
means a reduced degree of dependence of the corresponding wear pa- law as suggested by the model described above. To explain the presence
rameter on the applied normal load. This happens because the zone of of large number of shear induced deformation features in the surface
influence at both surface and sub-surface regions is smaller for a and sub-surface regions of the scratch grooves (Figs. 5–7) the maximum
lower and larger for a higher applied load. So it is highly likely that shear stress (τmax) was calculated following [38]
the number of microstructural existing defects whose effects are aver-
!1=3
aged out will be large in the case of the higher zone of influence than 16Peff E2r
in the case of lower sized zone of influence. This will surely affect the ex- τmax ¼ 0:445 ð17Þ
9π2 r2
tant and the severity of interaction between the already existing micro-
structural defects and the scratch process induced defects in the present
SLS glass. Further, it was very interesting to note that in general for a where
given applied normal load, the power law exponents increased system- hn  o n  oi−1
2 2
atically ranging from scratch width to scratch depth to wear volume Er ¼ 1−νs =Es þ 1−νi =Ei ð18Þ
with a moderate gradient. It is evident from this data that as far as the
degree of the power law dependencies of the various wear parameters The effective load Peff is given by [39,40]
on the applied normal load are concerned, for a given applied normal
load the sub-surface related wear parameters assume significance gen-  
2 0:5
erally higher than those of the surface related wear parameter. This Peff ¼ P 1 þ μ ð19Þ
happens because the shear stress increases generally up to the maxi-
mum depth of magnitude equal to half the static contact radius follow- where μ is the experimentally measured friction coefficient and r is the
ing [16,37,38]. In addition, the magnitudes of these exponents were the radius of the indenter.
highest for the lowest applied normal load of 5 N and the lowest for the The maximum shear stress data are plotted as a function of the
highest applied normal load of 15 N. This happens most likely because scratching speed, v in Fig. 8. Its magnitude (τ ~ 10–30 GPa, Fig. 5) was
the probability of interaction between the already existing sub-surface much higher than the theoretical shear strength (τtheor ~ 3–6 GPa) of
defects in the glass and the scratch induced damages was the highest SLS glass [41,42]. This was why a huge number of shear bands and
in the case of the applied normal load of 5 N which has the smallest shear induced micro-damaged localized zones were observed on the
zone of influence. However, the probability of interaction between the al- surface of the scratch grooves and much more so in the sub-surface re-
ready existing sub-surface defects in the glass and the scratch induced gion beyond the scratch grooves in the present SLS glass, Figs. 6 and 7.
damages was the lowest in the case of the applied normal load of 15 N
which has the largest zone of influence and thereby might have already
consumed in the scratch groove of higher depth and width a large number

Table 1
Values of the pre-exponential factors and the power law exponents for the SLS glass.

Load(N) 5 10 15

Values of the power law exponents (β, γ, δ, ε) and the pre-exponential Factor (A, B,
C, D)
Width β = 0.11, β = 0.07, A = 23.34 β = 0.02, A = 23.75
A = 17.135
Depth γ = 0.29, γ = 0.09, B = 0.226 γ = 0.03, B = 0.374
B = 0.0754
Wv δ = 0.4, δ = 0.16, δ = 0.05,
C = 2 × 10− 15 C = 2.4 × 10− 14 C = 1.9953 × 10− 14
Fig. 8. The variations of maximum shear stress, τ active just underneath the scratching
wr ε = 0.4, ε = 0.16, ε = 0.05,
indenter as a function of scratching speeds at an applied normal load of 15 N. The insets
D = 1 × 10− 13 D = 8.125 × 10− 13 D = 4.434 × 10− 13
show the similar data for applied normal loads of 5 and 10 N.
P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907 1905

applied normal load with the scratching speed. This data explains
why for a given scratching speed the sub-surface brittle fractures
and consequent damage evolutions aggravated as the applied normal
load was increased, Figs. 6 and 7(a), (c), (e) and (b), (d), (f). As the
magnitude of the tensile stress (~0.75 to 3 GPa) was already much
higher than the experimentally measured three point bend strength
(~98.11 ± 9.22 MPa) of the SLS glass [26], the occurrence of substan-
tial brittle fractures and consequent damage evolutions (Figs. 5–7)
was expected.
Depending upon the damage evolution process in the surface and
sub-surface regions two schematic sketches are given in Figs. 10
and 11 to illustrate the different modes of deformation and damage
in the present SLS glass. As depicted in Fig. 10, the general feature
was that the interactions among the Hertzian tensile cracks as well
Fig. 9. The variations of the maximum tensile stress, σ acting at the wake of the scratching as the damage inside the scratch grooves were lower for the higher
indenter as a function of the scratching speeds for different applied normal loads.
velocities. Further, the main characteristic features were in and out
plane shear deformation, shear induced localized micro‐fracture,
For a given applied load the estimated magnitude of maximum micro-chip formation, edge crack formation, intra Hertzian tensile
shear stress increased with the scratching speed, Fig. 8. That is the crack formation, in and out of plane microcracking as well as micro-
reason why, for a given applied normal load the sub-surface shear in- wear debris formation. The edge cracks which generated from the
duced damage signatures were more prominent at a higher Hertzian tensile cracks were mainly observed in low speed at both
scratching speed than at a lower scratching speed, Fig. 5(a)–(f). Fur- lower and higher loads (Figs. 10(a), (c), 3(a), (c), 4(a), (c), 5(a)
ther, for a given scratching speed, higher applied normal load leads and (c)). The interaction among the Hertzian tensile cracks indicated
to higher magnitude of the maximum shear stress (Fig. 8). This esti- with a dotted circle leads to chipping and formation of micro‐wear
mation clarifies why for a given scratching speed the sub-surface de- debris (Figs. 10(b), 4(a), (c), (e), (f)) depicted as the gray dots in
formation was more acute at a higher applied normal load than at a Fig. 10. For higher applied normal loads (Fig. 10(c), (d)) the areas
lower applied normal load, Figs. 6 and 7(a), (c), (e) and (b), (d), (f). enclosed in between the Hertzian tensile ring cracks were usually
Similarly, the tensile stress (σ) active at the wake of the indenter found to be heavily plastically deformed showing signatures of exten-
was estimated as ~0.75 to 3 GPa (Fig. 9) following [16,37]. sive shear induced deformations and micro-fractures and the severity
of shear band formation were usually more for lower scratching
ð1−2νs ÞP speed (Figs. 10(c), 5(c), (e)). Presence of two mutually orthogonal
σ ¼ ð1 þ 15:5μ Þ  2=3 ð20Þ
2π 43kPr=Es
family of shear bands was indicated with red and blue colors
(Fig. 10(c), (d)). However, the shear deformation zones were arrested
where, at the boundary of the Hertzian tensile ring cracks (Figs. 10(d), 5(d), (f)).
At higher applied normal load (Figs. 10(c), (d), 5(e), (f)) the relative
  
9  2
 
2 Es amount of fracture events was usually more than the amount of defor-
k¼ 1−νs þ 1−νi ð21Þ
16 Ei mation and damage induced by shear deformation. The intra Hertzian
tensile cracks were formed at higher applied normal load at an
It is evident from the data presented in Fig. 9 that for a given angle greater than zero with the scratching direction (Figs. 10(c),
scratching speed the estimated tensile stress active at the wake of (d), 4(e), (f)) and the number of such cracks generally decreased
the indenter increased with applied normal load and for a given with the scratching speed (Figs. 10(d), 4(f)).

Fig. 10. Typical schematic model of damage evolution of the scratch grooves for increasing scratching speed at both low and high applied normal loads.
1906 P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907

Fig. 11. Typical schematic model of damage evolution in the cross-section of the scratch grooves for increasing scratching speed at both low and high applied normal loads.

In the sub-surface damage zone (Fig. 11), there were two distinct immediately that, for a given applied normal load the reduction in
areas of attention e.g. (a) the region encompassing the main scratch scratch speed will produce a smaller magnitude of tensile stress ac-
grooves and (b) the region encompassing the damage area beyond tive at the wake of the indenter. So, the extent of surface damage
the scratch groove. The main scratch grooves were full of crushed ma- could be minimized by such a combination, if desired. However,
terial. However, in the region beyond the scratch grooves a large such apparent contradictions may be easily resolved if we bear in
number of shear induced deformation bands had formed in the pre- mind that every grinding polishing operation is for a given applica-
sent SLS glass. For lower applied normal load the crushing in the tion which will demand an intelligent optimization of the process pa-
groove as well as micro‐chip formations (indicated with pink line) rameters that is needed to achieve the desired surface finish. For
was lower (Figs. 11(a), 6(a)) especially for higher scratching speed instance, given the present results the most obvious choice for the
(Figs. 11(b), 6(b)). The fracture as well as micro‐chip formations present SLS glass could be a properly optimized low load, low speed
was higher for higher applied normal loads (Figs. 11(c), (d), 6(c)–(f)). regime where a low load of magnitude just nominally beyond the
The extent of sub-surface damage zone beyond the main scratch critical load (Pc) value will create smaller amount of surface damages
grooves as well as the shear induced deformation increased with because the tensile stress active at the wake of the indenter is low and
scratching speed both for lower and higher applied normal loads this scenario will overcompensate for the damages that could be in-
(Figs. 11, 6). The micro-cracks through the different planes (indicated duced due to the higher contact time in low speed scratch process.
with red line) were generated at higher applied normal loads at higher Since a grinding polishing operation can be imagined as the sum
scratching speed (Figs. 11(d), 7(d)). In addition micro‐wear debris (in- total effect of many single pass scratch processes occurring simulta-
dicated with gray dots) and the micro-cracks (indicated with green neously, at least in principle the analogy could be extended to such
lines) were also formed in the sub-surface damage regions, Figs. 11, 7. operations. On the other hand the low speed will ensure that the ex-
tent of tensile cracking and the sub-surface damages in the glass are
5. Conclusions kept at a magnitude as minimum as possible. It will be extremely im-
portant to keep the sub-surface damages at a level as minimum as
For applied normal loads of 5, 10 and 15 N, the depth, width, wear possible because such micro-damages can lead to slow growth of
volume and corresponding wear rate of the scratch grooves in the SLS cracks in supportive environment and may degrade the actual service
glass showed inverse power law dependencies on scratching speed as life of the glass component. In fact the future aim of the present re-
predicted by a model developed in the present work and also because search is oriented towards such a humble beginning in search of an
the time of contact had an inverse power law dependence on effective optimization of the process parameters for a glass of given
scratching speed (100–1000 μm/s). The results show further that for chemical composition and the surface micromechanical as well
a given applied load more surface but lesser sub-surface damage is in- nanomechanical properties. It is also appreciated that the effect of
duced to the glass material at lower scratch speed. The lesser sub- local temperature increase and if the temperature is high enough,
surface damage happens because the lower scratch speeds contribute the role of temperature induced visco-elastic deformations during
to relatively smaller magnitudes of the shear stress active just under- the scratching process are vitally important issues in understanding
neath the indenter. This information appears to be in contrast with the material removal process. Further dedicated experimental design
the fact that the estimated tensile stress active at the wake of the and actual experimental verification will be needed to resolve such
indenter increased with applied normal load and for a given applied issues which were beyond the scope of the present work. Although
normal load with the scratching speed (Fig. 9). It then follows not separately looked into in the present investigation it is well
P. Bandyopadhyay et al. / Journal of Non-Crystalline Solids 358 (2012) 1897–1907 1907

appreciated that equally important will be the exclusive investigation [15] J.B.D. Veldkamp, N. Hattu, V.A.C. Snijders, in: R.C. Bradt, D.P.H. Hasselman, F.F.
Lange (Eds.), Fracture Mechanics of Ceramics, Vol. 3, Plenum Press, USA, 1978,
about the role of the changes in relative humidity in governing the pp. 273–300.
scratch induced damage evolution and the consequent material re- [16] B.R. Lawn, Proc. R. Soc. Lond. A 299 (1967) 307–316.
moval mechanisms in SLS glass. It is planned to conduct further re- [17] K. Li, Y. Shapiro, J.C.M. Li, Acta Mater. 46 (1998) 5569–5578.
[18] F. Petit, C. Ott, F. Cambier, J. Eur. Ceram. Soc. 29 (2009) 1299–1307.
search in aforesaid areas so that the scratch damage process in [19] W. Gu, Z. Yao, X. Liang, Wear 270 (2011) 241–246.
commercially available SLS glass can be better understood. [20] H.B. Abdelounis, K. Elleuchb, R. Vargiolu, H. Zahouani, A.L. Bot, Wear 266 (2009)
621–626.
[21] V.L. Houerou, J.C. Sangleboeuf, S. Deriano, T. Rouxel, G. Duisit, J. Non-Cryst. Solids
Acknowledgements 316 (2003) 54–63.
[22] V.R. Howes, A. Szameitat, J. Mater. Sci. Lett. 3 (1984) 872–874.
Permission of Director, CSIR- CGCRI, Kolkata to publish this work [23] Y. Li, H. Huang, R. Xie, H. Li, Y. Deng, X. Chen, J. Wang, Q. Xu, W. Yang, Y. Guo, Opt.
Express 18 (2010) 17180–17186.
and financial support of CSIR (Project No. NWP 0027) are gratefully [24] J. Neauport, C. Ambard, P. Cormont, N. Darbois, J. Destribats, C. Luitot, O. Rondeau,
acknowledged. The experimental assistance of Mrs. S. Roy of CSIR- Opt. Express 17 (2009) 20448–20456.
CGCRI is gratefully acknowledged. [25] F. Yang, J. Non-Cryst. Solids 351 (2005) 3861–3865.
[26] P. Bandyopadhyay, A. Dey, A.K. Mukhopadhyay, Int. J. Appl. Glas. Sci. (2012),
http://dx.doi.org/10.1111/j.2041-1294.2011.00073.x.
References [27] P. Bandyopadhyay, A. Dey, S. Roy, A.K. Mukhopadhyay, J. Non-Cryst. Solids 358
(2012) 1091–1103.
[1] A.A. Tseng, C.-F.J. Kuo, S. Jou, S. Nishimura, J. Shirakashi, Appl. Surf. Sci. 257 (2011) [28] P. Bandyopadhyay, A. Dey, S. Roy, N. Dey, A.K. Mukhopadhyay, Appl. Phys. A
9243–9250. (2012), http://dx.doi.org/10.1007/s00339-012-6828-3.
[2] B. Ramezanzadeh, S. Moradian, H. Yari, A. Kashani, M. Niknahad, Tribol. Int. 47 [29] P. Bandyopadhyay, A. Dey, A.K. Mandal, N. Dey, S. Roy, A.K. Mukhopadhyay, Appl.
(2012) 77–89. Phys. A (2012), http://dx.doi.org/10.1007/s00339-012-6844-3.
[3] D. Barnes, S. Johnson, R. Snell, S. Best, J. Mech. Behav. Biomed. Mater. 6 (2012) [30] www.pilkington.com/resources/ats129swproperties20090804.doc.
128–138. [31] Jing Li, J.C.M. Li, Mater. Sci. Eng. A 409 (2005) 108–119.
[4] S. Hazra, P.P. Bandyopadhyay, Mater. Des. 35 (2012) 243–250. [32] M. Tomozawa, C.-Y. Li, T.M. Gross, J. Non-Cryst. Solids 356 (2010) 1194–1197.
[5] E. Rayón, V. Bonache, M.D. Salvador, E. Bannier, E. Sánchez, A. Denoirjean, H. [33] T.M. Gross, M. Tomozawa, J. Non-Cryst. Solids 354 (2008) 4056–4062.
Ageorges, Surf. Coat. Technol. 206 (2012) 2655–2660. [34] B.R. Lawn, N.P. Padture, H.D. Cai, F. Guiberteau, Science 263 (1994) 1114–1116.
[6] Q. Zheng, M. Potuzak, J.C. Mauro, M.M. Smedskjaer, R.E. Youngman, Y. Yue, J. Non- [35] I.M. Petersonl, A. Pajares, B.R. Lawn, V.P. Thompson, E.D. Rekow, J. Dent. Res. 77
Cryst. Solids (2012), http://dx.doi.org/10.1016/j.jnoncrysol.2012.01.030. (1998) 589–602.
[7] Asma Perveen, M.P. Jahan, M. Rahman, Y.S. Wong, J. Mater. Process. Technol. 212 [36] P.K. Giri, J. Banerjee, Introduction to statistics, second ed., Academic Publisher,
(2012) 580–593. India, 2002, pp. 298–299.
[8] M.V. Swain, Proc. R. Soc. Lond. A 366 (1979) 575–597. [37] B.R. Lawn, F.C. Frank, Proc. R. Soc. Lond. A 299 (1967) 291–306.
[9] A.K. Mukhopadhyay, D. Chakroborty, M.V. Swain, Y.-W. Mai, J. Eur. Ceram. Soc. 17 [38] C.E. Packard, C.A. Schuh, Acta Mater. 55 (2007) 5348–5358.
(1997) 91–100. [39] B.R. Lawn, S.M. Wiederhorn, D.E. Roberts, J. Mater. Sci. 19 (1984) 2561–2569.
[10] M. Klecka, G. Subhash, Wear 265 (2008) 612–619. [40] P.E. Miller, T.I. Suratwala, L.L. Wong, M.D. Feit, J.A. Menapace, P.J. Davis, R.A.
[11] H. Kasem, S. Bonnamy, Y. Berthier, P. Jacquemard, Tribol. Int. 43 (2010) Steele, in: G.J. Exarhos, A.H. Guenther, K.L. Lewis, D. Ristau, M.J. Soileau, C.J.
1951–1959. Stolz (Eds.), Proc. Of SPIE, 5991, 2005, pp. 599101-1 – 599101-25.
[12] O. Borrero-López, M. Hoffman, A. Bendavid, P.J. Martin, Thin Solid Films 519 [41] R.W.K. Honeycombe, Plastic Deformation of Metals, second ed. Edward Arnold
(2011) 7925–7931. Ltd., London, UK, 1984.
[13] A. Vencl, S. Arostegui, G. Favaro, F. Zivic, M. Mrdak, S. Mitrovic, V. Popovic, Tribol. [42] A. Puthucode, R. Banerjee, S. Vadlakonda, R. Mirshams, M.J. Kaufman, Metall.
Int. 44 (2011) 1281–1288. Mater. Trans. A 39A (2008) 1552–1559.
[14] R. Crombez, J. McMinis, V.S. Veerasamy, W. Shen, Tribol. Int. 44 (2011) 55–62.

You might also like