Foam Glass Processing Using A Polishing Glass Powder Residue

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Available online at www.sciencedirect.

com

CERAMICS
INTERNATIONAL
Ceramics International 39 (2013) 5869–5877
www.elsevier.com/locate/ceramint

Foam glass processing using a polishing glass powder residue


Yiğit Attilaa, Mustafa Güdenb,n, Alper Tas-demircib
a
Department of Materials Science and Engineering, Izmir Institute of Technology, Gülbahc- e Köyü, Urla, Izmir, Turkey
b
Department of Mechanical Engineering, Izmir Institute of Technology, Gülbahc- e Köyü, Urla, Izmir, Turkey
Received 8 November 2012; received in revised form 20 December 2012; accepted 20 December 2012
Available online 11 January 2013

Abstract

The foaming behavior of a powder residue/waste of a soda-lime window glass polishing facility was investigated at the temperatures
between 700 and 950 1C. The results showed that the foaming of the glass powder started at a characteristic temperature between 670
and 680 1C. The maximum volume expansions of the glass powder and the density of the foams varied between 600% and 750% and
0.206 and 0.378 g cm  3, respectively. The expansion of the studied glass powder residue resulted from the decomposition of the organic
compounds on the surfaces of the glass powder particles, derived from an oil-based coolant used in the polishing. The collapse stress of
the foams ranged between  1 and 4 MPa and the thermal conductivity between 0.048 and 0.079 W K  1 m  1. Both the collapse stress
and thermal conductivity increased with increasing the foam density. The foams showed the characteristics of the compression
deformation of the open cell brittle foams, which was attributed to the relatively thick cell edges.
& 2013 Elsevier Ltd and Techna Group S.r.l. All rights reserved.

Keywords: A. Sintering; C. Mechanical properties; C. Thermal properties; D. Glass

1. Introduction CO2, while redox agents react with glass, forming the
gaseous products of CO2, CO or N2, depending on the
Light-weight foam glass parts have been known since composition of the used blowing agent. Various types of
1930s [1]. Nevertheless, the interests in these materials glass powders and blowing agents were previously inves-
resume nowadays [2]. Foam glass parts are inert, incom- tigated for the processing of foam glass. The glass particles
bustible and relatively strong, and typically used as the polishing wastes containing SiC particles were used as
thermal insulation material. However, the application blowing agent for foaming soda-lime glass powder [5] and
areas have the potentials for widening, including light the powder mixture of soda-lime glass and glass fiber [6].
weight filler for the restoration of failed slopes, subgrade The glass cullet mixed with 20 wt% fly ash was foamed
improvement material, light-weight aggregate material in using marble polishing plant sludge (mainly composing of
concrete and water folding material for greening [3]. In the calcite and dolomite) and using SiC as blowing agent [7].
current foam glass manufacturing processes, a mixture of The dismantled waste cathode ray tube glasses were
glass powder and blowing agent, either in the form of loose foamed using CaCO3, SiC and TiN as blowing agents
powder or briquettes, is heated above the softening/melt- and hydrothermal hot-pressing [8–14]. Borosilicate glasses
ing point of glass. Then, the blowing agent decomposes or were reported to be effectively foamed using SiC and Si3N4
reacts with glass and releases gaseous products, which over a wide range of temperatures [15]. Water and liquid
drive the expansion of the glass. Blowing agents are hydrocarbons were further investigated as blowing agents
classified as neutralizers (e.g. CaCO3 and CaMg(CO3)2) in foam glass processing; the latter were recommended to
and redox agents (e.g. C, SiC and Si3N4) [4]. Neutralizers retard the effect of combustion reaction rate [16,17].
decompose upon heating, releasing the gaseous product of About 0.9 million ton of waste glass is generated in EU
annually [7]. The glass bottles in Japan were reported 1.98
n
Corresponding author. Tel.: þ 90 232 7506779; fax: þ 90 232 7506701. million tons in 1998 and 74% of the container glass
E-mail address: [email protected] (M. Güden). production was used as waste glass [3]. The cullet is

0272-8842/$ - see front matter & 2013 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
http://dx.doi.org/10.1016/j.ceramint.2012.12.104
5870 Y. Attila et al. / Ceramics International 39 (2013) 5869–5877

estimated to reach about 0.5 million ton per year in UK The scanning electron microscopy (SEM) analyses were
[18]. Only 0.21 million ton of 0.67–0.77 million ton of performed in a FEI Quanta 205 FEG SEM. The density of
waste flat glass generated every year in UK are recycled, the foams was determined by measuring the weight and
while the rest is used as landfill [19]. Another study reports dimensions of the test samples.
that only 20% of the collected waste glass are returned to The foaming experiments were performed using the glass
glass factories for re-melting [20]. The current waste glass- powder briquettes, 3 cm in diameter and 7 mm in thickness
based foam glass manufacturing seems to be one of the (Fig. 1(b)). The briquettes were prepared inside a cylindrical
most economical ways of recycling large quantities of mold under 15 MPa. The foaming behavior of waste glass
waste glasses generated. In this study, the foaming beha- powder (70 mm in average particle size) obtained by grinding
vior of a polishing glass powder residue was investigated soda-lime window glass was also investigated for comparison.
experimentally. The powder residue is currently dumped in An oil-based coolant (Potenza 40) was mixed with window
a land near the facility. The total amount of powder glass powder with an amount of 10 wt% in order to
generated is estimated to be 30,000 ton per year. The investigate the effect of coolant on the foaming and TGA
expansion of the studied glass powder residue was shown behavior of the glass powder. The schematic and the picture
to be resulted from the decomposition of the organic of the front view of foam expansion set-up are shown in
compounds on the surfaces of the glass powder particles Fig. 2(a) and (b), respectively. The details of the experimental
derived from an oil-based coolant used in the polishing set-up are given elsewhere [21]. Briefly, the foam expansion
operations. The expansions of the glass powder were set-up consisted of a vertical tube furnace, linear expansion
measured in-situ using an experimental foaming set-up. measurement system and foaming mold. The bottom of
The compression mechanical behavior and thermal con- foaming mold (3 cm in diameter and 8 cm in height) was
ductivities of the prepared foams were determined and enclosed tightly and the glass powder briquette was placed at
compared with those reported in the literature. the bottom of the mold. A linear variable displacement
transducer (LVDT) was connected to the steel expansion
2. Materials and testing methods rod through a wire and two pulleys. A thermocouple directly
contacted to the bottom of the briquette was used to measure
As-received glass powder was a residue of a soda-lime the temperature of the briquette during foaming. Transducer
window/flat glass polishing facility, Camex (Bursa, Turkey). and thermocouple data were collected using a data logger
The polishing was accomplished with an oil-based coolant (Data Taker DT 80). In a typical expansion experiment, the
and the resultant polishing residue was damped into a field. weight of the expansion rod is balanced using a counter
The particle size of as-received glass powder (Fig. 1(a)) was weight in order to minimize the load exerted to the expanding
determined using a Malvern Mastersizer 2000 particle size glass powder briquette (Fig. 2(a)). The foaming mold accom-
analyzer. The crystallographic structures of the powders panying the glass powder briquette is inserted into the furnace
and foams were determined using a Philips X’Pert Pro X- by means of an elevator. After the briquette insertion into the
Ray Diffraction (XRD) device (Cu-Ka radiation, l ¼ 1.54 furnace, the top expansion rod connected to LVDT is lowered
A1 and 40 kV). The Fourier transformation infrared (FTIR) until the rod bottom head touches the briquette surface. The
analyses were performed using a Shimadzu 8400S FTIR linear expansion (mm) is converted into volume percent
spectrometer. The thermo-gravimetric analyses (TGA) were expansion (VE (%)) using the following equation:
conducted in a Perkin Elmer Diamond TG/DTA. The
elemental compositions of the powders and foams were hf ho
V E ð %Þ ¼  100 ð1Þ
determined using a Spectro IQ II X-ray fluorescence (XRF). ho

Fig. 1. The pictures of (a) as-received glass powder and (b) glass powder briquette used in the expansion experiments.
Y. Attila et al. / Ceramics International 39 (2013) 5869–5877 5871

Fig. 2. (a) The schematic of foam expansion measurement set-up and (b) the picture showing foaming mold and sliding top and bottom rods.

Fig. 3. (a) The pictures of glass powder used to prepare the foam samples for compression and thermal conductivity tests and (b) the pictures of the
foamed glass powder briquette after foaming and cutting in the compression test sample size.

where hf and ho refer to the final and initial height of the glass the dimensions of 50  50  50 mm (Fig. 3(b)). The quasi-
powder briquette, respectively. static compression tests were conducted using a displacement-
The foam samples for compression and thermal conductiv- controlled SHIMADZU AG-I universal tension–compression
ity measurements were prepared separately by foaming test machine at the strain rate of 2  10  3 s  1. The thermal
70  70  10 mm size rectangular glass powder briquettes conductivity measurements were performed on 100  50  10
(Fig. 3(a)). These briquettes were formed inside a steel die mm size foam samples using a KEM QTM-500 thermal
under 15 MPa. The powder briquette was then placed inside a conductivity detector.
steel mold. The inner surfaces of the mold were coated with a
thin layer of kaolin. The kaolin layer prevented the reaction 3. Results and discussion
between the molten glass and the mold [12]. The expansions of
the glass powder briquettes were limited to the vertical 3.1. Characterization and expansion experiments
direction. The foaming was performed in a furnace between
700 and 950 1C (25 1C intervals) with two heating rates: slow The particle size of as-received glass powder ranged
heating rate, 5 1C min  1, and relatively fast heating rate, between 5.7 mm (d10) and 266 mm (d90), with an average of
15 1C min  1. After foaming, the foam samples were cut into 22.7 mm (d50). The composition of as-received glass powder
compression and thermal conductivity test sample dimensions determined by X-ray fluorescence analysis is as following:
as seen in Fig. 3(b). The compression test samples were 72.76% SiO2, 11.18% Na2O, 11.31% CaO, 1.74% MgO
prepared in accord with ASTM C240-97 standard [22] with and 1.61% Al2O3. The composition of the glass cullet
5872 Y. Attila et al. / Ceramics International 39 (2013) 5869–5877

generated in various countries was reported as 67–72%


SiO2, 14–15% Na2O, 7–11% CaO, 1–7% MgO and 1–6%
Al2O3 [23]. Therefore, the chemical composition of the as-
received glass powder matches well with that of soda-lime
window glass and the cullet (except Na2O) collected
around the world [2,3].
Fig. 4 shows the TGA curves of as-received glass
powder, window glass powder (control) and window glass
powder mixed with 10 wt% oil-based coolant. The weight
loss of as-received glass powder is  11 wt% at 662 1C and
continues until about 860 1C. A weight gain,  1 wt%,
seen above 860 1C is likely due to the oxidation reactions
of molten glass. Somewhat, a similar behavior is seen in
the TGA curve of window glass powder mixed with
10 wt% oil-based coolant. The weight loss continues until
about 860 1C, while the weight loss at 662 1C is slightly
lower, 9 wt%. But, the weight loss of window glass powder
is only about 0.5 wt% until 860 1C as shown in Fig. 4. The
weight loss of as-received glass powder residue is, there-
fore, attributed to the burning of organic compounds on
the glass powder particle surfaces. The weight losses of the
glass powder briquettes after the foaming between 750 and
950 1C were also measured for comparison and found to
be very similar to that of TGA, 11 wt% on the average.
The XRD spectrum of the as-received powder at room
temperature showed typical soda-lime window glass amor-
phous structure [1], while quartz, wollastonite and diopsite
crystal phases of soda lime glass were detected in the XRD
spectra of foam glass samples foamed above 800 1C. The
wollastonite and diopside are the crystal phases of the soda
lime glass formed upon heating and were also previously
detected in the foaming of soda lime glass powder at
950 1C [5,24] and 1000 1C [20].
Fig. 5(a) shows the volume expansion–temperature and
volume expansion–time graphs of the glass powder bri-
quettes heated to 900 1C within 3000 s (50 min). The
numbers in the same figure represent the different stages

Fig. 5. The volume expansion–time and temperature–time graphs of glass


powder briquettes (a) heated to the same final temperature with the same
heating rate (the numbers corresponding to the different stages of the
foam expansion) and heated to different final temperatures (b) with the
same and (c) different heating rates.

of glass powder briquette expansion. In Region 1, the glass


Fig. 4. The TGA curves of as-received glass powder, window glass powder sinters as the temperature increases. A slight
powder and 10 wt% oil-based coolant mixed window glass powder. increase in the volume of the briquette until about 700 1C
Y. Attila et al. / Ceramics International 39 (2013) 5869–5877 5873

in this region is due to the thermal expansion of the glass powder. The cells are initially small and spherical in
briquette. The foaming starts and reaches the maximum shape. As the foaming proceeds, the cells not only become
expansion in Region 2 at the temperatures between 670 and larger but also change the shape from sphere to poly-
680 1C and 866 and 872 1C, respectively. While, the expan- hedral. The decaying processes become operative in the
sion is nearly constant in Region 3 and decreases slightly in later stages as cell coarsening, cell wall rupture and
Region 4, as the temperature of the foamed glass powder drainage. The drainage decreases the thickness of cell
briquette reaches the furnace temperature. The transition walls, leading to thicker cell edges and cell wall rupture.
from Region 1 to Region 2 indicates the temperature at The cellular structure of a foam glass is shown in Fig. 6(a).
which the glass powder transforms partially or completely The foam sample was prepared by inserting the foaming
into a viscous liquid. However, the decomposition of the mold into a preheated furnace (800 1C) and holding the
blowing agent starts before the softening of glass particles. mold for 3000 s (50 min) in the furnace. The foam cell
The sintering of the particles through neck regions is walls seen in Fig. 6(a) accommodate interconnections with
believed to prevent the escape of the gases before the neighboring cells as marked by arrows. The bursting action
melting of the glass powder, causing the rapid expansion of the gaseous products and the partial melting/sintering of
of the briquette upon the sintering and/or partial melting of the glass particles on the cell walls lead to the formation of
the glass particles. The continued decomposition of the holes between the neighboring cells, resulting in an open
blowing agent at increasing temperatures further drives the cell foam structure. This naturally increases the water
expansion of the glass powder briquette. The maximum absorption of the foamed glass. Fig. 6(b) and (c) are the
expansion is also well agreed with the TGA curve shown in SEM micrographs showing the presence of small size pores
Fig. 4, in which the blowing agent decomposition vanishes on the cell edges and cell walls. The presence of small pores
after about 860 1C. Therefore, no foam expansion or on the cell edges and cell walls were also observed
slightly reduced foam expansions are detected in Region 3 previously in a similar foamed glass [24]. The needle-like
and Region 4. Fig. 5(b) shows the volume expansion– features seen in Fig. 6(c) on the cell walls (marked with
temperature and volume expansion–time curves of glass arrows) are noted to be wollastonite and diopsite crystals.
powder briquettes heated to the final temperatures between The bottom and cross-section pictures of a foamed widow
700 and 900 1C with the same heating rates. The final glass powder/10 wt% oil-based coolant briquette are shown
expansions (at 3000 s) are noted to increase rapidly, when in Fig. 6(d). As with foams processed using polishing glass
the glass powder briquettes are heated up to 850 1C. powder residue, these foams possess open cell foam structure.
However, the final expansions show saturation tendency Fig. 7 shows the FTIR spectra of as-received
between 850 and 900 1C. The final expansions are 600– glass powder, 10 wt% oil-based coolant added window
750% between 850 and 900 1C; the briquette thickness glass powder and foamed glass (powdered). The FTIR
increases about 6 and 7 times. The selected volume spectra of glass powder and oil-based coolant added
expansion–time and temperature–time graphs of the powder window glass powder are very similar: the peaks between
briquettes foamed with various heating schedules (shown by 3600 and 3200 cm  1 correspond to the broad stretching of
letters) and different heating rates are shown in Fig. 5(c). O–H and between 2960 and 2800 cm  1 to the stretching
The briquette heated quickly to 900 1C in 2500 s (50 min) vibration of C–H aldehyde group. The peak at 1660 cm  1
shows the maximum expansion (a),  700%, while the corresponds to the stretching vibration of C–C, 1460 cm  1
briquette heated quickly to 800 1C shows a lower final to the asymmetrical stretching of CH group and
expansion (b), 540%. The briquette heated to 800 1C with a 1500 cm  1 to the aromatic structure of CQC [25]. The
slow heating rate (c) results in similar final expansion with above groups of bonds are absent in the FTIR spectrum of
the quick heating to 800 1C (b). The briquettes heated to foamed glass powder. The main strong peak at 1050 cm  1
700 1C (d) in 8400 s (140 min) and to 750 1C in 13,800 s and the shoulder peak at 1200 cm  1 are the asymmetrical
(230 min) also show comparably lower expansions. It is also stretching of Si–O–Si bonds; the low frequency band
noted that the expansions decrease after about 870 1C, between 400 and 500 cm  1 is due to the rocking motion
regardless of the heating schedule and rate, as shown by of Si–O–Si bridges, whereas the corresponding bending
arrows in Fig. 5(c). The reductions in the expansions above mode is responsible for adsorption at 700–850 cm  1 [26].
870 1C are partly due to the vanishing of the decomposition The peaks seen between 3600 and 1500 cm  1 are due to
reactions of the blowing agent and partly due to the reduced the presence of oil-based coolant. The oil-based coolant is
melt viscosity leading to the collapse of the cells walls. The a neutralizing blowing agent, dissociating to give gaseous
reduced expansions of powder briquettes at slow heating products similar to CaMg(CO3)2 and CaCO3, as
rates were also reported previously [10]. The slow heating m  m
rates cause the early release of the gases from the blowing Cn Hm þ þ n O2 - H2 O þ nCO2 ð2Þ
4 2
agent, before the viscosity of the glass becomes low enough
to allow the glass to expand rapidly [11]. The organic compounds wetting the surfaces of the glass
The foam glass cellular structure evolution may be particles dissociate to give water vapor and carbon dioxide
considered composing of several consecutive stages. The during heating in the furnace, leading to the expansion of
foaming starts in the solid state just before the melting of the glass powder residue.
5874 Y. Attila et al. / Ceramics International 39 (2013) 5869–5877

Fig. 6. The SEM micrographs of foamed glass powder (800 1C) showing (a) the cells, (b) small pores on the cell walls and edges, (c) wollastonite and
diopsite crystals and (d) the pictures of foamed window glass powder using oil-based coolant as blowing agent.

3.2. Compression mechanical behavior and thermal


conductivity

The compressive stress–strain curves of the foams prepared


with relatively slow (5 1C min  1), and fast heating (15 1C
min  1) rates are shown in Fig. 8(a) and (b), respectively. The
tested foam samples show the typical stress–strain behavior
of the brittle foams: following the collapse stress (maximum
stress) the samples fracture catastrophically. The variation of
the foam collapse stresses with density is further shown
Fig. 9(a). The collapse stress varies between 1 and 4 MPa and
increases with increasing foam density. No significant effect
of heating rate on the collapse stresses of the foams is also
seen in the same figure; therefore, the effect of heating rate is
excluded in the following strength analysis. The foam
collapse stress is given as [27],
Fig. 7. The FTIR spectra of glass powder, oil-based coolant mixed
window glass powder and foamed glass powder. sf ¼ sbend ½Cfrrel 3=2 þ ð1fÞrrel  ð3Þ
Y. Attila et al. / Ceramics International 39 (2013) 5869–5877 5875

Fig. 8. The compressive stress–strain curves of foamed powder briquettes; Fig. 9. (a) The variations of collapse stress with foam density and (b) the
heating rates (a) 5 and (b) 15 1C min  1. fitting results of the collapse stresses of the tested foam glass and the
previously investigated foam glass with Eq. (3).

where, sbend is the bending strength of the soda lime glass; C


is a constant; rrel is the relative density of the foam and f is 300 kg cm  3 and with the f values between 0.7 and 0.9 for
the volume fraction of the solids located on the cell edges. the foam densities above 300 kg cm  3. The determined fitting
The value of C and the bending strength of soda lime glass values of f, in addition to the microscopic observations,
are sequentially given as 0.2 [27] and 70 MPa [5]. The first confirm the open cell foam structure of the tested and
term in Eq. (3) is due to bending and the second term is due previously investigated foam glasses. This is mainly caused
to membrane stretching of the cell walls. Eq. (3) predicts the by a high volume fraction of the solids located on the
collapse stresses of open-cell foams, when f equals to 1, and cell edges.
the collapse stresses of closed cell foams, when f equals to 0. The thermal conductivities of the foam glass samples as
The collapse stresses of the tested foam glass samples and function of density are shown in Fig. 10, together with the
previously investigated foam glasses are fitted with Eq. (3) thermal conductivities of the similar foam glasses reported
using different values of f. The fitting results, collapse stress in the literature (references in Fig. 10: 1—[11], 2—[29],
vs. density, are shown in Fig. 9(b) (references in Fig. 9(b): 3—[19], 4—[19], 5—[23]). Similar to the collapse stress, the
1—[19], 2—[19], 3—[18], 4—[28], 5—[29], 6—[30], 7—[7], thermal conductivity increases as the foam density
8—[6], 9—[31],10—[32], 11—[33]). The collapse stresses of increases. The thermal conductivities of the tested foam
the investigated and reported foam glasses are well fitted with glass range between 0.048 and 0.079 W m  1 K  1. The
the f values between 0.7 and 1 for the foam densities up to tested foam glass samples, as noted in Fig. 10, show similar
5876 Y. Attila et al. / Ceramics International 39 (2013) 5869–5877

[5] E. Bernardo, R. Cedro, M. Florean, S. Hreglich, Reutilization and


stabilization of wastes by the production of glass foams, Ceramics
International 33 (6) (2007) 963–968.
[6] E. Bernardo, G. Scarinci, P. Bertuzzi, P. Ercole, L. Ramon, Recy-
cling of waste glasses into partially crystallized glass foams, Journal
of Porous Materials 17 (3) (2010) 359–365.
[7] H.R. Fernandes, D.U. Tulyaganov, J.M.F. Ferreira, Production and
characterisation of glass ceramic foams from recycled raw materials,
Advances in Applied Ceramics 108 (1) (2009) 9–13.
[8] E. Bernardo, G. Scarinci, S. Hreglich, Foam glass as a way of
recycling glasses from cathode ray tubes, Glass Science and Technol-
ogy 78 (1) (2005) 7–11.
[9] F. Mear, P. Yot, M. Cambon, A.M. Ribes, Elaboration and
characterisation of foam glass from cathode ray tubes, Advances in
Applied Ceramics 104 (3) (2005) 123–130.
[10] F. Mear, P. Yot, M. Ribes, Effects of temperature, reaction time
and reducing agent content on the synthesis of macroporous foam
glasses from waste funnel glasses, Materials Letters 60 (7) (2006)
929–934.
[11] F. Mear, P. Yot, R. Viennois, M. Ribes, Mechanical behaviour and
thermal and electrical properties of foam glass, Ceramics Interna-
Fig. 10. The variations of thermal conductivities of the tested foam
tional 33 (4) (2007) 543–550.
glasses and the previously investigated foam glasses with density.
[12] H.W. Guo, Y.X. Gong, S.Y. Gao, Preparation of high strength foam
glass–ceramics from waste cathode ray tube, Materials Letters 64 (8)
thermal conductivities with the foam glasses reported in (2010) 997–999.
[13] Z. Matamoros-Veloza, J.C. Rendon-Angeles, K. Yanagisawa,
the literature: the thermal conductivities range between M.A. Cisneros-Guerrero, M.M. Cisneros-Guerrero, L. Aguirre, Pre-
0.04 and 0.05 W m  1 K  1 for the foam densities between paration of foamed glasses from CRT TV glass by means of
100 and 200 kg m  3. hydrothermal hot-pressing technique, Journal of the European
Ceramic Society 28 (4) (2008) 739–745.
[14] M.J. Chen, F.S. Zhang, J.X. Zhu, Lead recovery and the feasibility
4. Conclusions of foam glass production from funnel glass of dismantled cathode ray
tube through pyrovacuum process, Journal of Hazardous Materials
The foaming behavior of a powder residue/waste of a 161 (2–3) (2009) 1109–1113.
[15] B. Gerhard, Foaming of borosilicate glasses by chemical reactions in
soda-lime window glass polishing facility was investigated
the temperature range 950–1150 Å 1C, Journal of Non-Crystalline
using an in-situ foam expansion measurement set-up at the Solids 38–39 (Part 2 (0)) (1980) 855–860.
temperatures between 700 and 950 1C. The foaming was [16] S.K. Goyal, I.B. Cutler, Absorption of water in waste glass as a
shown to start at a characteristic temperature between 670 precursor for foam formation, Journal of Non-Crystalline Solids 19
and 680 1C. The maximum volume expansions of the glass (0) (1975) 311–320.
[17] V.E. Manevich, K.Y. Subbotin, Mechanism of foam-glass formation,
powder and the density of the foams varied between 600%
Glass and Ceramics 65 (5–6) (2008) 154–156.
and 750% and 0.206 and 0.378 g cm  3, respectively. The [18] J.P. Wu, A.R. Boccaccini, P.D. Lee, M.J. Kershaw, R.D. Rawlings,
expansion of the studied glass powder residue was shown Glass ceramic foams from coal ash and waste glass: production
to be resulted from the decomposition of the organic and characterisation, Advances in Applied Ceramics 105 (1) (2006)
compounds on the surfaces of the glass powder particles, 32–39.
[19] Wrap, /http://www.wrap.org.ukS, Environment Agency, Waste &
derived from oil-based coolant used in the polishing. The
Resources Action Programme, 2011.
heating rate of the foaming process had no significant [20] K.K. Éidukyavichus, V.R. Matseikene, V.V. Balkyavichus,
effect on the collapse stress and thermal conductivity. Both A.A. Shpokauskas, A.A. Laukaitis, L.Y. Kunskaite, Use of cullet
the collapse stress and thermal conductivity increased with of different chemical compositions in foam glass production, Glass
increasing the foam density and showed well agreements and Ceramics 61 (3) (2004) 77–80.
[21] M. Guden, S. Yuksel, SiC-particulate aluminum composite foams
with those reported in the literature. The foams showed the
produced from powder compacts: foaming and compression beha-
characteristics of the deformation of the open cell brittle vior, Journal of Materials Science 41 (13) (2006) 4075–4084.
foams, which was attributed to the relatively thick [22] ASTM, C240-97, in: Standard Test Methods of Testing Cellular
cell edges. Glass Insulation Block, ASTM International, West Conshohocken,
PA, 2003.
[23] N.M. Bobkova, S.E. Barantseva, E.E. Trusova, Production of foam
References glass with granite siftings from the Mikashevichi deposit, Glass and
Ceramics 64 (1–2) (2007) 47–50.
[1] A.H. Baker, Pittsburg Corning Corp, (1938). [24] D.U. Tulyaganov, H.R. Fernandes, S. Agathopoulos,
[2] V.A. Lotov, E.V. Krivenkova, Kinetics of formation of the porous J.M.F. Ferreira, Preparation and characterization of high compres-
structure in foam glass, Glass and Ceramics 59 (3–4) (2002) 89–93. sive strength foams from sheet glass, Journal of Porous Materials
[3] J. Lu, K. Onitsuka, Construction utilization of foamed waste glass, 13 (2) (2006) 133–139.
Journal of Environmental Sciences—China 16 (2) (2004) 302–307. [25] R.M. Silverstein, F.X. Webster, D.J. Kiemle, Spectrometric Identi-
[4] Y.A. Spiridonov, L.A. Orlova, Problems of foam glass production, fication of Organic Compounds, Hoboken, NJ, John Wiley & Sons,
Glass and Ceramics 60 (9–10) (2003) 313–314. 2005.
Y. Attila et al. / Ceramics International 39 (2013) 5869–5877 5877

[26] M. Dussauze, V. Rodriguez, A. Lipovskii, M. Petrov, C. Smith, materials from aluminosilicate initial materials, Glass and Ceramics
K. Richardson, T. Cardinal, E. Fargin, E.I. Kamitsos, How does 66 (3–4) (2009) 82–85.
thermal poling affect the structure of soda-lime glass?, Journal of [31] J. Garcia-Ten, A. Saburit, M.J. Orts, E. Bernardo, P. Colombo,
Physical Chemistry C 114 (29) (2010) 12754–12759. Glass foams from oxidation/reduction reactions using SiC, Si3N4 and
[27] L.J. Gibson, M.F. Ashby, Cellular Solids, Second ed., Cambridge AlN powders, Glass Technology: European Journal of Glass Science
University Press, Cambridge, UK, 1999. and Technology Part A 52 (4) (2011) 103–110.
[28] A.I. Shutov, L.I. Yashurkaeva, S.V. Alekseev, T.V. Yashurkaev, [32] A.S. Llaudis, M.J.O. Tari, F.J.G. Ten, E. Bernardo, P. Colombo,
Study of the structure of foam glass with different characteristics, Foaming of flat glass cullet using Si3N4 and MnO2 powders,
Glass and Ceramics 64 (9–10) (2007) 297–299. Ceramics International 35 (5) (2009) 1953–1959.
[29] O.V. Kaz’mina, V.I. Vereshchagin, A.N. Abiyaka, Prospects for use [33] M.S. Garkavi, O.K. Mel’chaeva, A.I. Nazarova, Effect of the process
of finely disperse quartz sands in production of foam-glass crystalline parameters of mix preparation on the properties of foam glass, Glass
materials, Glass and Ceramics 65 (9–10) (2008) 319–321. and Ceramics 68 (1–2) (2011) 44–46.
[30] O.V. Kaz’mina, V.I. Vereshchagin, A.N. Abiyaka, Assessment of the
compositions and components for obtaining foam-glass-crystalline

You might also like