Você está na página 1de 8

Journal of Non-Crystalline Solids 447 (2016) 190–197

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids

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

Influence of the glass particle size on the foaming process and physical
characteristics of foam glasses
Jakob König a,b, Rasmus R. Petersen a, Yuanzheng Yue a,⁎
a
Department of Chemistry and Bioscience, Aalborg University, DK-9220 Aalborg East, Denmark
b
Advanced Materials Department, Jožef Stefan Institute, SI-1000 Ljubljana, Slovenia

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

Article history: We have prepared low-density foam glasses from cathode-ray-tube panel glass using carbon and MnO2 as the
Received 23 February 2016 foaming agents. The effect of the glass particle size on the foaming process, the apparent density and the pore
Received in revised form 3 May 2016 morphology is revealed. The results show that the foaming is mainly caused by the reduction of manganese.
Accepted 4 May 2016
Foam glasses with a density of b150 kg m−3 are obtained when the particle size is ≤33 μm (D50). The foams
Available online 14 June 2016
have a homogeneous pore distribution and a major fraction of the pores are smaller than 0.5 mm. Only when
Keywords:
using the smallest particles (13 μm) does the pore size increase to 1–3 mm due to a faster coalescence process.
Foam glass However, by quenching the sample from the foaming to the annealing temperature the pore size is reduced by
Manganese oxide a factor of 5–10. The foams with an apparent density of b200 kg m−3 are predominantly open-porous. The
Thermal conductivity foams exhibit a thermal conductivity as low as 38.1 mW m−1 K−1 at a density of 116 kg m−3. For the investigated
Porosity foam glasses, there exists a great potential to further decrease their thermal conductivity by increasing the closed
Particle size porosity and by changing the trapped gas to CO2.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction (CRTs) [14], fly ashes from coal production [15] and vitrified solid
wastes from municipal waste combustion [16] can also be used. The
Foam glass is a highly porous (N 85 vol.% pores), lightweight material foaming agent, added as a minor part to the powdered glass, releases
used for thermal and acoustic insulation [1,2]. Depending on the gases based on a decomposition reaction or a redox reaction with the
application, e.g., heat insulation or soundproofing, the foam glass glass. The solid residues of the foaming agent or its reaction with the
has predominantly closed or open porosity, respectively. Foam glass have an impact on the glass network and foam morphology.
glass exhibits several advantages over conventional insulating materials Therefore, foaming agents must be selected according to the composition
such as organic foams and mineral wool, e.g., water and steam resis- of the glass.
tance, freeze-thaw tolerance, excellent chemical and thermal stability, Different foaming agents and/or their reactions with the glass lead to
high surface area and permeability, and superior mechanical properties. a wide variety of gases being found in the pores. In general, CO2 is the
Due to the chemical stability and the high degree of closed porosity, the best gas to be trapped in the pores, since it is easily obtainable, has a
lifespan of foam glass (N100 years [3]) is much longer than conventional low thermal conductivity and a low toxicity. Gases like H2O, O2 and N2
thermal insulation materials. are also acceptable, but they have a higher thermal conductivity than
Foam glass is produced from powdered glass admixed with a foaming CO2. More problematic is the presence of CO, H2 or SO3 due to their
agent that releases gases at elevated temperatures. In the softened glass, high toxicity, flammability and/or reactivity. CO2 gas is readily available
the expanding gas bubbles increase the volume of the sample, thus from the decomposition of carbonates [8,14,17] and carbon-based
forming a porous lightweight product. In industrial foam-glass production foaming agents, like pure carbon materials, carbohydrates and SiC [1,
the majority of the glass originates from waste glass [1]. Some production 18–20]. The latter are used by the foam-glass industry [1]. When carbon
facilities only make use of waste glass, thus eliminating the energy- is used as a foaming agent, the gas released at elevated temperatures is a
consuming melting step. Foam-glass production makes it possible to CO/CO2 mixture, where the ratio depends on the availability of oxygen.
recycle a great diversity of glassware or other amorphous materials During cooling the mixture then transforms to CO2, according to
into a high-value-added product. The waste glasses most commonly Boudouard's equilibrium. When applying carbohydrates or SiC as a
employed are mixed-color bottle glass [4–7] and window glass [8–12], foaming agent, the gas-releasing mechanism becomes more complex
but other glasses, like lamp glass [13], glasses from cathode ray tubes [18,21]. For this reason, pure carbon is preferred over carbohydrates
and SiC. The foaming starts after the glass has sintered to a state with
⁎ Corresponding author. closed pores, where the volume of the entrapped gases is too small for
E-mail address: yy@bio.aau.dk (Y.Z. Yue). a successful foaming. Therefore, a sufficient amount of free oxygen has

http://dx.doi.org/10.1016/j.jnoncrysol.2016.05.021
0022-3093/© 2016 Elsevier B.V. All rights reserved.
J. König et al. / Journal of Non-Crystalline Solids 447 (2016) 190–197 191

to be available from the glass melt. In order to promote the gas-releasing PM100). The powders were then sieved to obtain five powder fractions
reaction, i.e., the oxidation of carbon, an oxidizing agent can be added, with different particle sizes using laboratory sieves (38–250 μm). The
which acts as an oxidation/reduction couple with the carbonaceous chemical composition of the CRT panel glass is reported elsewhere
foaming agent [10,11,22]. The most suitable oxidizing agents are Fe2O3 [25]. Carbon (activated charcoal; Bie & Berntsen, Søborg, Denmark)
[10,16], MnO2 [10,11,18] and sulfates [16,18,19]. However, the solid and MnO2 (98%, Bie & Berntsen) were ball milled to pulverize the
residues of the foaming agent can act as a nucleation and crystallization agglomerates and larger particles. The foaming mixtures were prepared
agent. This means that the impact of the foaming agent on the glass sta- from panel glass, 0.93 wt.% carbon and 6.76 wt.% MnO2, according to
bility has to be considered. previously determined optimum concentrations [37]. The carbon/
Despite the fact that foam-glass technology has been known since MnO2 ratio was equal to one-half of the stoichiometric amount needed
the 1930s, the number of academic publications has only started to for the theoretical carbon oxidation reaction:
increase in the past 15 years. The main motivation for the research
seems to be environmental consciousness and the increasing price of C þ 2MnO2 →CO2 þ 2MnO ðReaction 1Þ
landfilling waste. The research on foam glass mainly concentrates on
its mechanical properties, only few papers report on the thermal The powder mixtures were homogenized on a carousel in a glass
conductivity [9,14,15,21,23–26]. However, a worldwide focus on energy bottle with four light plastic balls for 17 h at 30 rpm. The particle size
efficiency has put insulating properties higher on the agenda. Therefore, distribution of the powder mixtures was measured with a LS 13320
attention has been drawn to modified conventional and contemporary laser granulometer (Beckman Coulter, Brea, US-CA). The D50 and D90
materials with improved insulating properties [27–29]. In order to values of the powder mixtures as well as the foaming agents are given
promote the use of foam glass in thermal insulation, its insulating prop- in Table 1. The particle size distribution is shown in Fig. S1. Compared
erties need to be enhanced. to the previous study [37], the powder processing in the present study
The thermal insulation properties of a foam glass can be improved by was different, and thus the influence of the glass particle size on the
(i) decreasing the density, (ii) closing the porosity, (iii) entrapping a low foaming process could be revealed.
conducting gas, and (iv) employing glasses with a low thermal conduc- The powder mixtures were uniaxially pressed into disk-shaped
tivity. To prepare a low-density foam glass (b 150 kg m−3), optimum pellets with 40 MPa. Small pellets (1 g, Ø = 13 mm) were used to
conditions have to be found, i.e., a foaming agent compatible with the study the influence of the particle size and the foaming conditions on
composition of the glass, an optimum concentration and the right the density and homogeneity of the foams. Larger pellets (Ø =
heat-treatment conditions. Finely milled powder and homogenous mix- 35 mm) of selected powder mixtures were used to prepare large sam-
tures result in small pores [6,30]. If the pores are closed, then convective ples (23–45 g) in a stainless-steel cylinder (6 cm in diameter and
heat transfer is prevented. Moreover, the viscosity as well as the surface 5 cm in height) to study the thermal conductivity and compressive
tension should be in a defined range at the foaming temperature to strength. The pellets were placed in an electrical laboratory-chamber
support low density and small, closed pores [1,19]. When the pores furnace and heated at different heating rates (5 or 10 °C/min) to
are small and closed, the overall thermal conductivity of the foam 835 °C and heat-treated for different times (5 to 30 min). Subsequently,
glass can be estimated by summing the contributions from the conduc- the samples were cooled at 4 °C/min to the glass-transition temperature
tion through the solid and the gaseous phases [31,32]. Based on an (Tg) of 530 °C [35] and then slowly cooled at 1 °C/min to room temper-
assumption of linear contributions [31,32] the difference in the thermal ature. The temperature in the furnace was calibrated using an external
conductivity between a foam glass filled with O2 (24 mW m− 1 K−1) thermocouple.
and a foam glass filled with CO2 (15 mW m− 1 K− 1) [33] would be Thermogravimetic (TG) analyses were performed using a Jupiter
8.6 mW m− 1 K− 1 for a foam glass with 95% porosity. 449 simultaneous thermal analysis (STA) instrument coupled with a
The effect of the particle size on the foam density and homogeneity 403C Aëoloss mass spectrometer (MS) (Netzch, Selb, Germany). The
is, despite being well known, scarcely reported [1,6,25,30,34]. It is com- measurements were performed with a heating rate of 10 °C/min in an
mon to report the particle size of the glass and the foaming agent before air atmosphere. The powder mixture was compressed with 40 MPa
they are mixed. However, since the mixing is frequently carried out and a small piece (30–40 mg) was placed into an alumina crucible.
using milling equipment, the particle size changes, obscuring any clear The CO2 peak was integrated to calculate the amount of carbon burnt
relation between the particle size and the foam density. It was shown out at the defined temperature.
that when foaming with carbon the density of the foam glass decreases The apparent density (ρapp ) of the small and the large foam
rapidly with a glass particle size below 100 μm [1,30]. On the other samples was determined using Archimedes' principle in a 10 wt.%
hand, the relationship between the particle size and the foam density polyethylenglycol (PEG 3000 Da, Bie & Berntsen) solution or was
is not so straightforward when using carbonates as the foaming agent calculated from the samples' mass and dimensions, respectively. A
[25]. PEG solution was used to prevent absorption of the liquid medium
The aim of the present work is to study the influence of the particle during the measurement. The pycnometer density (ρ pyc ) of the
size on the foaming process and to prepare foam glasses with a low foam samples was determined with a helium pycnometer (Ultrapyc
thermal conductivity. We foamed waste CRT panel glass with carbon 1200e, Quantachrome Instruments, Boynton Beach, US-FL). The powder
and MnO2 as the foaming agents. CRT panel glass has a high glass stabil- density (ρpowder) of the crushed foamed-glass sample was measured
ity [25,35,36]. To reveal the influence of the particle size on the foaming with the pycnometer. The percentage of the closed porosity, out of the
process, pre-milled powder sieved to different fractions was mixed with
the foaming agents and the mixture was homogenized without changing
the particle size. In addition, the effect of the heat-treatment conditions Table 1
on the characteristics of the foam and its stability are revealed and Particle size (D50 and D90) of the powder mixtures, carbon and MnO2.
discussed. We characterized the apparent density, the closed porosity,
Sample D50 [μm] D90 [μm]
the compressive strength and the thermal conductivity of the foam
#1 13 32
glasses.
#2 33 80
#3 53 92
2. Experimental #4 103 163
#5 196 283
CTR panel glass was crushed in a jaw crusher (Retch BB51, Haan, carbon 31 54
MnO2 5.4 29
Germany) and subsequently ball milled in a planetary ball mill (Retsch
192 J. König et al. / Journal of Non-Crystalline Solids 447 (2016) 190–197

total porosity, was calculated from the apparent and pycnometer The analysis revealed a significant difference in the weight-loss
densities according to: (WL) curves between the finest (D50 = 13 μm) and the coarsest
(D50 = 196 μm) powders (Fig. 1). Three steps in the WL can be distin-
!
ρapp guished in both samples, which were separated into the following
TP ¼ 1 temperature ranges: (I) 200–564 °C, (II) 564–730 °C and (III) above
ρpowder:
730 °C. In step I, the WL of the sample from the fine powder is larger
than that of the sample from the coarse powder (1.05 vs. 0.85 wt.%)
!
ρapp ρapp and the CO2 peak appears at a higher temperature (527 vs. 470 °C).
CP ¼  No O2 peak was detected, which suggests that the oxygen released by
ρpyc ρpowder
the decomposition of MnO2 reacts with the carbon. During step I partial
burning of the carbon and partial reduction of the MnO2 occurs. At the
CP end of step I, 14 and 5.8 wt.% of unburned carbon remains in the fine
CP ½% ¼  100
TP and coarse powder samples, respectively. The larger WL and the more
unburned carbon in the fine powder sample indicate that more carbon
The compressive strength was measured on large cylinder-shaped is burnt out by the oxygen originating from MnO2 reduction than in
samples using a table-top universal tester (Autograph AGS-J, Shimadzu, the sample of coarse powders. In the coarse powder sample more
Kyoto, Japan) with a crosshead speed of 0.5 mm min− 1. Prior to carbon can be burnt by the atmospheric oxygen because the green den-
measurements the top and bottom surfaces of the foam samples were sity is lower, and hence more atmospheric oxygen is present between
coated with a 1-mm layer of a commercial mineral building glue. The the particles, enabling a higher gas flow in and out of the sample. Thus
thermal conductivity was measured from the top and bottom sides oxidation of the carbon by the atmosphere occurs at lower tempera-
and the average value is reported in relation to the density of the sam- tures (CO2 peak at 470 °C). On the other hand, less carbon is burnt by
ple. The thermal conductivity was measured with a surface probe the atmospheric oxygen in the fine powder sample, but it eventually
(25 × 25 mm2) at 10 °C using a heat-transfer analyzer (ISOMET Model burns when the oxygen is released by the decomposition of MnO2 as
2104 Heat Transfer Analyzer, Applied Precision, Bratislava, Slovakia). the temperature increases, with the CO2 peak at 527 °C. The carbon
The accuracy of the equipment in the relevant range was tested by mea- that is not burnt by the atmosphere shifts the reduction of MnO2 to
suring commercial expanded-polystyrene and foam-glass samples with lower temperatures, and thus the overall WL in the fine powder sample
known values. The obtained results were in perfect agreement with the in step I is higher, despite more unburnt carbon staying in the sample at
declared values. 564 °C.
In step II (564–730 °C), the rest of the carbon is burnt out and the
3. Results and discussion MnO2 decomposes to Mn2O3 [26,38]. At around 690 °C the sample
densifies and the pores close [25,26]. This is seen as a halt in the WL.
The combination of a carbon-based foaming agent and an oxidizing The closing of the pores leads to an increase in the oxygen partial
agent is an effective way to produce low-density foam glass [10,11,18, pressure in the sample, which inhibits further reduction of the MnO2.
22]. In the present investigation we applied pure carbon and MnO2, Therefore, not all of the MnO2 is decomposed to Mn2O3 at the end of
which would provide CO2 gas according to Reaction 1. However, since step II. A complete reduction of MnO2 to Mn2O3 and complete burning
the thermal treatment was performed in an air atmosphere, it is expect- of the carbon would result in a WL of 1.55 wt.%; however, only
ed that part of the carbon will be oxidized before the pores between the 1.33–1.37 wt.% is observed. The same mechanism appears in both
softened glass particles are closed as a result of the sintering process. samples; however, the WL in step II for the coarse powder sample is
This was observed for the sample foamed with carbon in a previous larger as less MnO2 was decomposed in step I.
investigation [37], according to which foaming occurred only when In step III (N 730 °C) the WL is caused by a further reduction of
the particle size (D90) is lower than 27 μm. In the present investigation, manganese. The reduction in the fine powder sample continues when
coarser particles are used, with the lowest D90 value equal to 32 μm. In the temperature reaches 790 °C and the WL is accelerated above
this respect, it is expected that all carbon will be burnt by the oxygen 820 °C. The total WL at 900 °C (1.74 wt.%) indicates that almost all the
from surrounding atmosphere. The release of gases and weight loss manganese is in the form of Mn3O4 (theoretical WL is 1.76 wt.%).
were determined by performing a TG-MS analysis. When pure MnO2 is heated in air the decomposition from Mn2O3 to
Mn3O4 begins at 900 °C [26,38]. The presence of a glass melt favors a
higher ratio of Mn2+/Mn3+ and shifts the reduction to lower tempera-
tures. This is accordance with our previous study [26]. We suggested
that Mn2O3 first dissolves in the glass and then Mn3+ reduces to
Mn2+. In the coarse powder sample there is a small, gradual WL from
730 to 860 °C, which then accelerates above 910 °C. Such behavior can
be explained if the agglomeration of small MnO2 particles (5.4 μm,
D50) between large glass particles (196 μm, D50) in the mixture is
considered. Thereafter, a much smaller contact surface between the
agglomerated manganese oxide and the glass is available and the disso-
lution is slower. Consequently, the decomposition of Mn2O3 to Mn3O4 is
observed at the same temperature as in pure manganese oxide [26,38].
Some of the agglomerates located on, or close to, the surface of the pellet
gradually decompose (gradual WL in the range 730–860 °C), while the
reaction in the interior is decelerated due to the high oxygen partial
pressure.
The TG-MS results show that the carbon is burnt out before the pores
close (b 690 °C). Thus the foaming is caused by the oxygen gas released
with the reduction of manganese. A great part of the oxygen released by
Fig. 1. Sample weight (W) and CO2 evolution (MS) from powder mixtures with different the decomposition of MnO2 to Mn2O3 is lost to the atmosphere as the
particle sizes (D50) during heating in an air atmosphere. reaction occurs before the sample is sintered to closed pores [26]. To
J. König et al. / Journal of Non-Crystalline Solids 447 (2016) 190–197 193

the exposure of the samples to high temperatures is longer, which is


reflected in a lower density. This is to be expected since the TG results
(Fig. 1) show the gas-forming reaction starts at 790 °C and then acceler-
ates at 820 °C (curve for the fine powder sample). The main reaction in
this temperature range is a reduction of the Mn2O3 dissolved in the
glass [26]. A larger contact area when using finer powders leads to a
faster dissolution and thus the reaction occurring at a lower tempera-
ture. On the other hand, in the coarse powder sample the dissolution
of Mn2O3 is slower and the temperature has to be increased to 900 °C
(see Fig. 1) before the thermal decomposition of Mn2O3 occurs [38].
All the foam glasses are homogeneous and contain only a few larger
pores (Figs. 3 and 4). The photographs and SEM images indicate a
narrow pore size distribution. In most foam glasses the pore size is
below 0.5 mm, except in the sample prepared from the finest particles
foamed for 30 min. In this sample the pore size is 1–3 mm. The foam
glass made from the coarsest powders has a pore size in the range of
50–300 μm, while the concentration of pores is much lower compared
Fig. 2. Influence of the particle size (D50) on the apparent density (ρapp) of the foam to the other samples (Fig. 4). In this sample only a small amount of
glasses prepared at 835 °C with different heating rates and foaming times (see legend). gas is released in the softened glass, as seen from the TG results
The inset magnifies the density data for smaller particle sizes. (Fig. 1). Pore-free regions of 200–300 μm that are visible in the sample
most probably originate from large glass particles, as the D90 value of
this powder is 283 μm.
The amount of closed pores is above 95 vol.% in the samples heated
increase the CO2 content in the foam glass prepared with carbon and at 10 °C/min (Fig. 5). These samples were exposed to high temperatures
manganese oxide in an air atmosphere, we suggest that the powders for a shorter time. The sample with the lowest density and a high degree
are finely milled [37] and Mn2O3 is used instead of MnO2. However, of closed pores, i.e., the sample that was prepared from 53-μm powders
even when Mn2O3 is used, partial burning of the carbon by the atmo- (D50), heated at 5 °C/min and foamed for 5 min, exhibits a pore size of
spheric oxygen is expected. To prevent burning of the carbon an 50–250 μm (Fig. 4b) and a density of 200 kg m−3. In contrast, open
oxygen-deficient atmosphere should be used. porosity dominates in the samples with a density smaller than
The influence of the particle size on the apparent density is shown in 200 kg m−3 (inset in Fig. 4). However, the pore size remains small
Fig. 2. The apparent density decreases with a decreasing particle size (e.g., 0.1–0.6 mm in Fig. 3c), except for the sample from the finest
and a prolonged foaming time. Very limited foaming is observed powders foamed for 30 min.
when 196-μm powder (D50) is used. This is in accordance with the TG In general, an open porosity and an inhomogeneous microstructure
results that show little WL in the temperature range 730–890 °C. In are characteristic of decomposing foaming agents like carbonates [14,
order to accelerate the gas release the temperature should be increased 19,25]. The reason for such an occurrence is the sudden increase in the
to above 890 °C. However, similarly low densities as observed for the pressure inside the pores at the temperatures where the main part of
finer powders cannot be obtained as the foam stability decreases with the decomposition occurs. The pressure increase is fast enough to tear
the drop in viscosity [36]. apart the thin walls, so the connections between the pores are formed.
The upper limit of the particle size at which an important foaming The mechanism in MnO2, which is also considered a decomposing
occurs, (i.e. a density below 500 kg/m3) is found at 163 μm (D90). foaming agent [6], is different (this study and ref. [26]). The gas release
This is significantly lower than the particle size limit reported in the when foaming with MnO2 is not as fast and detrimental as is the case for
literature, i.e., 400 μm (D90) [1,6]. On the other hand, the particle size carbonates [25,35,37]. The TG analysis for a carbonate sample showed a
required for the preparation of low-density foams (b150 kg m−3) is 0.7% WL over a range of 50 °C (see Fig. 1b in ref. [25]), while in this study
reported to be well below 10 μm (D50) [13,30,39,40]. Here, we obtained the WL for the fine powder sample is one-third of that (0.23%) in the
foam glass with a density of 158 kg m−3 using 53 μm powders (D50). temperature range 835–885 °C. However, with prolonged exposure to
Hence, less energy is needed for the processing of the raw materials. high temperatures the pore walls become very thin and connections
When heating the powders with a slower rate (5 °C/min vs. are formed (Fig. 4c). The sharp edges of the pore walls indicate that
10 °C/min) or when prolonging the dwell time (30 min vs. 5 min), they broke during the sample preparation. The thickness of these

Fig. 3. Cross-section images of the foam glasses prepared from powders with different particle sizes (D50). Samples were prepared at 835 °C with 5 °C/min for a) 5 min and b) 30 min.
194 J. König et al. / Journal of Non-Crystalline Solids 447 (2016) 190–197

Fig. 4. SEM micrographs of foam glasses prepared at 835 °C with a 5 °C/min heating rate using different particle sizes (D50) and treatment times: a) 196 μm, 30 min; b) 53 μm, 5 min;
c) 33 μm, 30 min; d) 13 μm, 30 min, and subsequently quenched. Open and full arrows in c) show the connections between the neighboring pores and the thin pore wall, respectively.
The inset in c) shows a broken pore wall with a thickness estimated from the rolled edges (white arrows) to b100 nm.

walls is estimated to be below 100 nm (inset in Fig. 4c). This is in accor- This sample was first heated to 835 °C at 5 °C/min, held for 30 min and
dance with the thickness reported in the literature [41]. then quenched from the foaming temperature to the glass-transition
The formation of connections is unwanted because when a pore temperature (~530 °C) by moving it directly into an annealing furnace.
opens to the surrounding atmosphere the driving force for further Thus the time the sample spent above the softening temperature was
expansion is lost. Closed porosity also decreases the heat convection shortened by 30 min. The density of the quenched sample is slightly
through the sample and provides a better thermal insulation when a lower than the density of the slowly cooled sample, i.e., 116 kg m−3 vs.
gas with a low thermal conductivity is captured in the pores. When 120 kg m−3, respectively (Table 2). In contrast, the pore size is 5–10-
the majority of the pores are interconnected, the density of the foam times smaller (0.1–0.6 mm) in comparison to the sample cooled at
can only decrease further with the nucleation and growth of new 5 °C/min (1–3 mm large), as shown in Fig. 6a and b. Such a large differ-
pores [26]. These usually nucleate in the thicker walls or struts, which ence suggests that the coalescence proceeds very quickly, even at
results in a bimodal pore size distribution. Initially, the size of an temperatures below the foaming temperature, i.e., during cooling.
individual pore increases due to the release of gases and the tempera-
ture increase. When the gas release ceases, the density of the foam starts
to increase, while the pores continue to grow due to coalesce. This is
observed for the sample from the finest particle size (13 μm, D50)
foamed for 30 min that has a slightly higher density compared to the
sample from 33-μm powders (D50), but has much larger pores (Fig. 3).
The small pore size and homogeneous microstructure observed in all
the samples indicate that the coalescence proceeds evenly throughout
the sample. To further investigate the coalescence and growth of the
pores, as well as the stability of the foam glass, two experiments with
a shorter and longer foaming time were performed. In the first experi-
ment, the treatment time was prolonged by decreasing the cooling
rate from 4 °C/min to 1 °C/min for the sample with 33-μm powders
(D50) foamed for 30 min. Accordingly, the time during which the
sample was heated above the softening temperature (i.e. 710 °C [36])
increased from 30 min to 2 h. The resulting foam glass had a high
density (320 kg m−3), a very irregular shape, an inhomogeneous pore
size distribution and thick walls (Fig. S2). Such an appearance indicates
extensive pore coalescence, which resulted in the partial collapse of the
foam.
Fig. 5. Closed porosity (CP) in relation to the particle size (D50) and density (ρapp; inset) of
The second experiment with a shortened treatment time was the foam glasses prepared at 835 °C with different heating rates and durations (see
performed on a large sample with the finest powder (13 μm, D50). legend).
J. König et al. / Journal of Non-Crystalline Solids 447 (2016) 190–197 195

Table 2
Comparison of the apparent density (ρapp) and closed porosity (CP) between small and
large samples prepared from powders with different particle sizes (D50).

Small samples Large samples

D50 ρapp CP ρapp CP


−3 −3 a
[μm] [kg m ] [%] [kg m ] [%]

13 143 10.9 121 ± 2 5.4


33 137 10.5 151 ± 7 3.4
103 282 88.1 355 ± 25 67
13 (quenched) – – 116 9.6
a
The average foam density and the variation in the density between three foam samples.

This shows the importance of a fast cooling rate after the foaming
process.
The apparent densities and closed porosities of the large and small
samples prepared under the same conditions are compared in Table 2.
The density of the large sample prepared from the finest powder is
smaller compared to the small foam sample, while the opposite is Fig. 7. Thermal conductivity (λ) of the foam glasses prepared with 0.93 wt.% carbon and
observed for the other two powder sizes. The difference occurs because 6.76 wt.% MnO2, heated at 5 °C/min to 835 °C and held for 30 min. The measurements
there are different optimum foaming times for the small and large were performed at 10 °C and the precision was ±0.5 mW m−1 K−1 (equal to the
samples. For the small samples prepared from 13-μm powders (D50), marker size). For comparison, data for samples prepared with CRT panel glass, 0.95 wt.%
carbon and 5.12 wt.% MnO2 [37] and commercial foam glasses [12] are shown. The lines
the pore size (much greater in comparison to the coarser powders)
represent linear fits.
and the density (greater than in the large sample) indicate that a
30-min foaming time is too long. The 30-min foaming time is closer to
the optimum for the large sample size from the 13-μm powders indicat- small samples (Fig. 2). Despite the large difference in the pore size
ed by the lower density (120 kg m−3). In the case of the 33-μm powders between the large samples prepared from 13-μm powders (Fig. 6a and
(D50), however, the 30-min foaming time is optimum for the small b), both samples exhibit comparable densities and a low percentage of
sample, while it is not long enough for the large sample. The extension closed porosity (Table 2). Such behavior indicates that the stability of
of the optimum foaming time with the increase in the sample size can the foam is good and not affected by the coalescence when the
be related to the difference in the heat transfer during the heat sample cools to the softening temperature within a reasonable
treatment. Since foaming takes place inside a metal cylinder and the time (i.e., 30 min). The variation in the density presented in Table 2
heat transfer is relatively low, the heating and cooling rates of the is in major part caused by the temperature gradients due to different
large samples are slower than those of the small samples. The fractions positions of the samples in the furnace. In comparison with a previous
of closed pores in the large samples are lower, which is most probably investigation [37] where the same raw materials were used, but the
related to slower cooling, during which time coalescence occurs. mixture was homogenized by ball milling, the foam samples obtained
The cross-sections of the large foam samples (Fig. 6) show that the in this study are more uniform throughout the sample height and the
samples are homogeneous and the pore sizes are comparable to the pores are smaller. Another difference is that the mixed powders were

Fig. 6. Cross-sections of the large foam glasses prepared at 835 °C for 30 min with a 5 °C/min heating rate. The foams were prepared from powders with different particle sizes (D50):
a) 13 μm (quenched), b) 13 μm, c) 33 μm, d) 103 μm.
196 J. König et al. / Journal of Non-Crystalline Solids 447 (2016) 190–197

foamed at a higher temperature than the ball-milled powders. This is in from the finest powder (13 μm, D50) foamed at 835 °C for 30 min, for
accordance with the observations of the present study (Figs. 1 and 2) which the pore size is in the range of 1–3 mm. However, the pore size
that the gas-forming reaction is observed at lower temperatures when remains in the range 0.1–0.6 mm if the sample is quenched from the
finer powders are used. foaming to the annealing temperature. The foams with an apparent
The compressive strength was measured for the large samples density of b200 kg m−3 are predominantly open-porous. The coales-
prepared from 33-μm powders (D50). At a density of 144 kg m−3 the cence and growth of the pores progresses at similar rates throughout
compressive strength is 0.48 MPa (Fig. S3). The compressive strength the sample and the connections between the pores are formed after
of a commercial foam glass (ρapp = 115 kg m−3) [12] was found to be the pore-wall thickness is reduced to below 100 nm. The compressive
0.4 MPa using the same method. If the difference in the density between strength of the foam glass is 0.48 MPa at a density of 144 kg m−3,
the samples is taken into account, the compressive strength of the which can potentially be improved by increasing the closed porosity.
investigated foams is lower than that of the commercial product. The The foams exhibit a thermal conductivity as low as 38.1 mW m−1 K−1
smaller compressive strength is attributed to a high content of open at a density of 116 kg m−3. Due to the predominantly open-porous struc-
porosity, which weakens the foam structure [14,31]. ture the foams could be used as indoor thermal and acoustic insulation.
The thermal conductivity was measured in samples with a density However, a great potential exists to further decrease the thermal conduc-
ranging from 116 to 158 kg m− 3 (Fig. 7). The thermal conductivity tivity of the investigated foam glasses.
decreases with increasing density. The lowest value obtained is Supplementary data to this article can be found online at http://dx.
38.1 mW m − 1 K − 1 for the quenched sample with a density of doi.org/10.1016/j.jnoncrysol.2016.05.021.
116 kg m−3. The slope of the thermal conductivity decrease with the
density is smaller in comparison to the foams prepared by ball-milling Acknowledgement
the powder mixture [37] and the commercial foam glass [12]. The reasons
for such a difference are not known at this point. In the literature only We thank L.S. Trankjær, R.E. Gissel and N.J. Klitmøller for preparing
scarce data on the thermal conductivity of foam glasses are available the powders, foam glasses, and measuring He-pycnometer and appar-
[15,21,37,42]. Our results indicate (Fig. 7) that foam glasses with smaller ent density. We also thank C. Prinds (Danish Technological Institute)
pores have a slightly lower thermal conductivity in comparison to for measuring the thermal conductivity and compressive strength of
samples from a previous report [37]. At a density of 150–160 kg m−3 the foam glasses. We thank the Danish National Advanced Technology
the difference is around 2–3 mW m−1 K−1, while at lower densities the Foundation for financial support under grant number 012-2011-3.
difference diminishes. However, another study on foam glass prepared
under similar conditions, but with a density above 200 kg m−3, showed
that there is no influence of the pore size on the thermal conductivity References
[26]. A detailed analysis of the thermal conductivity of foam glasses [1] G. Scarinci, G. Brusatin, E. Bernardo, Glass foams, in: M. Scheffler, P. Colombo (Eds.),
with similar densities and different pore sizes will have to be performed Cellular Ceramics, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 2005, pp. 158–176.
to address this discrepancy. [2] Foamglas Industrial Insulation Handbook, Pittsburgh Corning Europe, Waterloo,
Belgium, 1992.
In comparison to commercial products [12], the thermal conductivity [3] O.V. Kaz'mina, V.I. Vereshchagin, B.S. Semukhin, Structure and strength of foam-glass-
of the samples prepared in this study is lower when comparing values for crystalline materials produced from a glass granulate, Glas. Phys. Chem. 37 (2011)
the same densities (Fig. 7). The best commercial foam-glass product has a 371–377.
[4] http://www.glapor.com (accessed 20.4.16).
thermal conductivity of 38 mW m−1 K−1 at a density of 100 kg m−3, [5] O.V. Kaz'mina, V.I. Vereshchagin, A.N. Abiyaka, Prospects for use of finely disperse
while in this study the same thermal conductivity is obtained at a density quartz sands in production of foam-glass crystalline materials, Glas. Ceram. 65
of 116 kg m−3. Moreover, the prepared foam glasses have a high degree (2008) 319–321 (English translation of Steklo i Keramika).
[6] V. Ducman, M. Kovačević, The foaming of waste glass, Key Eng. Mater. 132–136
of open porosity. This shows that the potential for a further decrease in (1997) 2264–2267.
the thermal conductivity lies in optimizing the closed porosity and the [7] J.P. Wu, A.R. Boccaccini, P.D. Lee, M.J. Kershaw, R.D. Rawlings, Glass ceramic foams
pore-gas composition. The thermal conductivity is often expressed as a from coal ash and waste glass: production and characterisation, Adv. Appl. Ceram.
105 (2006) 32–39.
linear contribution from the solid and gas conductivities [31,32]. Thereaf-
[8] H.R. Fernandes, D.U. Tulyaganov, J.M.F. Ferreira, Preparation and characterization of
ter, a decrease in the gas conductivity could be significant if the air was to foams from sheet glass and fly ash using carbonates as foaming agents, Ceram. Int.
be exchanged with CO2. However, a more detailed analysis of the heat 35 (2009) 229–235.
[9] C. Vancea, I. Lazău, Glass foam from window panes and bottle glass wastes, Cent.
transfer in foam glasses needs to be performed to evaluate the
Eur. J. Chem. 12 (2014) 804–811.
contributions of the solid and gaseous conductivity to the overall [10] J. García-Ten, A. Saburit, M.J. Orts, E. Bernardo, P. Colombo, Glass foams from oxidation/
thermal conductivity. reduction reactions using SiC, Si3N4 and AlN powders, Glass Technol. Eur. J. Glass Sci.
The most important parameter in order to achieve CO2-filled pores is Technol. Part A 52 (2011) 103–110.
[11] A.S. Llaudis, M.J.O. Tari, F.J.G. Ten, E. Bernardo, P. Colombo, Foaming of flat glass cullet
the use of an oxygen-free atmosphere during the heat treatment in using Si3N4 and MnO2 powders, Ceram. Int. 35 (2009) 1953–1959.
order to prevent the premature oxidation of the carbon. Despite a low [12] http://www.foamglas.com (accessed 20.4.16).
thermal conductivity of the prepared foam glass, its use in thermal insu- [13] V.A. Lotov, E.V. Krivenkova, Kinetics of formation of the porous structure in foam
glass, Glas. Ceram. 59 (2002) 89–93 (English translation of Steklo i Keramika).
lation is limited to indoor applications as the open-pore structure is [14] E. Bernardo, F. Albertini, Glass foams from dismantled cathode ray tubes, Ceram. Int.
vulnerable to water penetration and freeze-thaw cycling. However, 32 (2006) 603–608.
due to the open pore structure it can be used effectively as an acoustic [15] H. Hojaji, Development of foam glass structural insulation derived from fly ash, Proc.
MRS Fall Meeting, Boston, US-MA, 136 1988, pp. 185–206.
insulator. [16] A. Steiner, Foam Glass Production From Vitrifies Municipal Waste Fly AshesPhD Thesis
Eindhoven University of Technology, Eindhoven, The Netherlands, 2006.
4. Conclusions [17] H.R. Fernandes, F. Andreola, L. Barbieri, I. Lancellotti, M.J. Pascual, J.M.F. Ferreira, The
use of egg shells to produce Cathode Ray Tube (CRT) glass foams, Ceram. Int. 39
(2013) 9071–9078.
Foaming of mixtures of CRT panel glass, carbon and MnO2 is possible [18] G. Bayer, Foaming of borosilicate glasses by chemical reactions in the temperature
when a particle size of ≤ 103 μm is used, while the foaming is almost range 950–1150 °C, J. Non-Cryst. Solids 39 (1980) 855–860.
[19] Y.A. Spiridonov, L.A. Orlova, Problems of foam glass production, Glas. Ceram. 60
completely halted in 196-μm powders (D50). The major part of the
(2003) 313–314 (English translation of Steklo i Keramika).
foaming results from the reduction of manganese, while the contribu- [20] V.E. Manevich, K.Y. Subbotin, Foam glass and problems of energy conservation, Glas.
tion of the carbon is negligible. The apparent density of the obtained Ceram. 65 (2008) 105–108 (English translation of Steklo i Keramika).
foam glasses is below 150 kg m− 3 when a particle size of ≤ 33 μm is [21] S. Köse, G. Bayer, Schaumbildung im System Altglas-SiC und die Eigenschaften
derartiger Schaumgläser, Glastech. Ber. 7 (1982) 151–160.
used (D50). All the foams are homogeneous and possess very small [22] E. Bernardo, R. Cedro, M. Florean, S. Hreglich, Reutilization and stabilization of
pores (the majority smaller than 0.5 mm), except for the foam prepared wastes by the production of glass foams, Ceram. Int. 33 (2007) 963–968.
J. König et al. / Journal of Non-Crystalline Solids 447 (2016) 190–197 197

[23] Y. Attila, M. Güden, A. Taşdemirci, Foam glass processing using a polishing glass [33] W.G. Kannuluik, H.B. Donald, The pressure dependence of the thermal conductivity
powder residue, Ceram. Int. 39 (2013) 5869–5877. of polyatomic gases at 0 °C, Aust, J. Sci. Res. Ser. A Phys. Sci. 3 (1950) 417–427.
[24] F. Méar, P. Yot, R. Viennois, M. Ribes, Mechanical behaviour and thermal and electrical [34] A. Pokorny, J. Vicenzi, C.P. Bergmann, Influence of particle size of raw materials in
properties of foam glass, Ceram. Int. 33 (2007) 543–550. the microstructure and properties of vitreous foam, Proc: Congress SAM/CONAMET,
[25] J. König, R.R. Petersen, Y.Z. Yue, Influence of the glass–calcium carbonate mixture's San Nicolás, Argentina 2007, pp. 1027–1032.
characteristics on the foaming process and the properties of the foam glass, J. Eur. [35] R.R. Petersen, J. König, M.M. Smedskjaer, Y.Z. Yue, Foaming of CRT panel glass powder
Ceram. Soc. 34 (2014) 1591–1598. using Na2CO3, Glass Technol. Eur. J. Glass Sci. Technol. Part A 55 (2014) 1–6.
[26] R.R. Petersen, J. König, Y.Z. Yue, The mechanism of foaming and thermal conductivity of [36] R.R. Petersen, J. König, M.M. Smedskjaer, Y.Z. Yue, Effect of Na2CO3 as foaming agent
glasses foamed with MnO2, J. Non-Cryst. Solids 425 (2015) 74–82. on dynamics and structure of foam glass melts, J. Non-Cryst. Solids 400 (2014) 1–5.
[27] J. Fricke, U. Heinemann, H.P. Ebert, Vacuum insulation panels—from research to [37] J. König, R.R. Petersen, Y.Z. Yue, Fabrication of highly insulating foam glass made
market, Vacuum 82 (2008) 680–690. from CRT panel glass, Ceram. Int. 41 (2015) 9793–9800.
[28] K.S. Jorgensen, U. Bauer, G. Rosenberg, K. Christensen, Method for manufacturing an [38] K. Terayama, M. Ikeda, Study on thermal-decomposition of MnO2 and Mn2O3 by
aerogel-containing composite and composite produced by that method, World Patent thermal-analysis, Trans. Jpn. Inst. Metals 24 (1983) 754–758.
WO2011012710, 2011. [39] M. Tasserie, D. Bideau, J. Rocherulle, P. Verdier, Y. Laurent, Study of a new foam material
[29] B. Wicklein, A. Kocjan, G. Salazar-Alvarez, F. Carosio, G. Camino, M. Antonietti, L. based on industrial glass, Verre 6 (1992) 2–15.
Bergström, Thermally insulating and fire-retardant lightweight anisotropic foams [40] J.D. Kirkpatrick, Process for making a cellulated vitreous material, US Patent,
based on nanocellulose and graphene oxide, Nat. Nanotechnol. 9 (2014) 1–7. US4198224, 1980.
[30] A.A. Ketov, An experience of reuse of a glass cullet for production of foam structure [41] Y.-N. Qu, J. Xu, Z.-G. Su, N. Ma, X.-Y. Zhang, X.-Q. Xi, J.-L. Yang, Lightweight and
material, Proc. Int. Symp.: Recycling and Reuse of Glass Cullet, Dundee, Scotland high-strength glass foams prepared by a novel green spheres hollowing technique,
2001, pp. 1–7. Ceram. Int. 42 (2016) 2370–2377.
[31] L.J. Gibson, M.F. Ashby, Cellular Solids: Structure and Properties, second ed. Cambridge [42] J.P. Wu, A.R. Boccaccini, P.D. Lee, R.D. Rawlings, Thermal and mechanical properties
University Press, Cambridge, 1999. of a foamed glass-ceramic material produced from silicate wastes, Glass Technol.
[32] E. Placido, M.C. Arduini-Schuster, J. Kuhn, Thermal properties predictive model for Eur. J. Glass Sci. Technol. Part A 48 (2007) 133–141.
insulating foams, Infrared Phys. Technol. 46 (2005) 219–231.

Você também pode gostar