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

Journal of Non-Crystalline Solids 513 (2019) 15–23

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


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

Effect of SiC content on viscosity and thermal properties of foam ceramic T


prepared from molybdenum tailings

Zhiheng Tanga, Mingxing Zhangb, Xuefeng Zhanga,b, , Jianli Hea, Xiaolin Jiab, Leibo Dengb
a
School of Physics, Inner Mongolia University of Science & Technology, Baotou 014010, China
b
Key Laboratory of Integrated Exploitation of Bayan Obo Multi-Metal Resources, Inner Mongolia University of Science and Technology, Baotou 014010, China

A R T I C LE I N FO A B S T R A C T

Keywords: A foam ceramic was prepared using molybdenum tailings (97–99.5 wt%) as the raw material and SiC as the
Foam ceramic foaming agent. The results showed that with increasing SiC content, the liquid-phase viscosity decreased, which
Molybdenum tailing in turn enhanced the overall expansion of the system, leading to a decrease in the thermal conductivity. The
Liquid-phase viscosity effects of the SiC content on the bulk density, porosity, pore size, and thermal conductivity were investigated in
Thermal conductivity
detail. Moreover, the experimental results were compared with several analytical models, and we found that the
parallel model closely agreed with the predictions. The sample prepared via sintering at 1150 °C using 2.5 wt%
SiC and molybdenum tailings exhibited a low bulk density of 0.37 g/cm3, high porosity of 85.14%, and low
thermal conductivity of 0.22W/m·K. Thus, the application of molybdenum tailings in the production of foamed
materials is considered to have potential economic and environmental advantages.

1. Introduction materials.
The control of liquid-phase viscosity (η) is considerably important in
A worldwide focus on energy efficiency has placed heat-insulating the foaming process because it severely affects the thermal and me-
properties high on the agenda; however, the widely used organic chanical properties, as well as the microstructure of the resulting
thermal insulation materials are flammable and highly hazardous, and samples [21–26]. For instance, Fernandes et al. [21] have suggested
these materials cause side effects during application [1–5]. A foam that the liquid-phase viscosity should be controlled in the range
ceramic (FC) exhibits several advantages over conventional insulating 107–108 Pa·s, which enables the foaming agent to release gas, leading to
materials (e.g., polyethylene foams): freeze-thaw tolerance, superior homogeneous porosity. However, Mugoni [22] and Souza et al. [23]
chemical and thermal stability, higher surface area and permeability, have recommended that a low viscosity, in the range 103–107 Pa·s, is
superior mechanical properties, and a longer lifespan [6–8]. sufficient to allow the expansion of the produced gas. Nevertheless,
Several researchers have synthesized FCs from various industrial some other researchers including Scarinci [24] have reported that the
wastes such as recycled glass [1–4,7–13], fly ash [14–16], and polishing viscosity should be in the range 103–105 Pa·s for optimal foaming
porcelain waste [17–19], via the formation of a glassy phase during conditions. Most researchers have focused on the feasibility and me-
heating. However, FC fabrication from solid waste has some limitations: chanical properties [4–7,13–16,20–24] of raw materials [15,16], and
(1) the amount of solid waste utilized is extremely low; (2) solid waste, only some of them have systematically studied the thermal conductivity
such as recycled glass, is mostly made up of components with different [27–30] of FCs, which is considered a relatively important parameter
colors; therefore, color sorting and impurity removal steps are required, that determines the heat-insulating properties of the materials.
leading to higher costs; and (3) recycled glass may contain con- In this paper, we demonstrate the effect of the SiC content on the
taminants such as metallic or nonmetallic fragments [6,20]. Never- liquid-phase viscosity and thermal conductivity of the fabricated FC.
theless, the main contents of molybdenum tailings (Mo-tailings) are
similar to the components of FC, and relatively, they have uniform
particle sizes and fixed compositions. These factors significantly reduce
the cost of FC production. Therefore, FCs prepared via recycling of Mo-
tailings is a promising alternative to organic thermal insulation


Corresponding author at: School of Physics, Inner Mongolia University of Science & Technology, Baotou 014010, China.
E-mail address: wjbltc@163.com (X. Zhang).

https://doi.org/10.1016/j.jnoncrysol.2019.02.024
Received 15 November 2018; Received in revised form 22 February 2019; Accepted 22 February 2019
Available online 15 March 2019
0022-3093/ © 2019 Elsevier B.V. All rights reserved.
Z. Tang, et al. Journal of Non-Crystalline Solids 513 (2019) 15–23

Table 1 S3 contained Fe2+ and Fe3+; this was confirmed from the presence of
The chemical composition of Mo-tailings. the 2P3/2 photoelectron lines at ca. 711.1 eV (Fe3+) and 713.6 eV
Component SiO2 Al2O3 Fe2O3 K2O Na2O CaO MgO TiO2 LOI (Fe2+) in the corresponding spectra. The collected XPS profiles were
analyzed using the CasaXPS software. All spectra were calibrated using
Mo-tailings 70.56 13.22 4.51 3.42 3.15 2.00 1.39 0.50 1.25 the adventitious C1s peak at 284.8 eV. Before analysis of the XPS
Est.Error 0.10 0.15 0.04 0.10 0.19 0.03 0.06 0.01
spectra, the background was subtracted using the multi-region Shirley
LOI (loss on ignition).
method.

2. Materials and methods 2.3. Structure and performance characterization

2.1. Characterization of raw materials The bulk density of the large FC samples was calculated from the
sample masses and volumes after cutting the samples to
The chemical composition of the Mo-tailings was analyzed by X-ray 10 × 10 × 10 mm parallelepipedic pieces. The powder density (ρpow)
fluorescence (XRF, Rigaku 6486-0024, Japan) and the results are shown was measured using a gas pycnometer (Micromeritics Accupyc 1330,
in Table 1. The presence of modifiers (Fe2O3, CaO, MgO, K2O, and Norcross, GA, USA). The total porosity (φ) was obtained from the bulk
Na2O) in the samples ensured reduction in the liquid-phase viscosity at density and powder density using the following equation:
high temperatures, which was propitious to the entrapment of gases.
⎛ ρ ⎞
From Table 1, MoO3 was not detected in Mo-tailings by XRF. However, φ = ⎜1 − ∗ 100%
considering the potentially harmful effects of Mo on the environment, ρpow ⎟ (1)
⎝ ⎠
human and ecosystem health, the content of Mo in the Mo-tailings was
Optical microscopy and scanning electron microscopy (Zeiss Sigma
further determined by atomic emission spectrometery (AES, iCAP6300,
500) were employed to observe the foam structures. The number of
Thermo, USA). The content of metal Mo was determined to be
pores were marked with twenty-times-magnified optical microscope
0.17 ± 0.0126 mg/L. Because the content of Mo is extremely low, the
images and counted with the Nano software.
raw material is relatively safe for humans and environment.
The SiC-related bloating effect was observed using a heating mi-
The crystalline phases of the fired specimens were identified by X-
croscope (TOM-AC, Fraunhofer, Germany) during calcination.
ray diffraction (XRD, Philips PW-1710, Netherlands) using the Cu Kα
Cylindrical samples with diameter 13 mm and height 5 mm were pre-
radiation in the scanning range 10–80° at room temperature; the XRD
pared by uniaxial pressing at 5 MPa and subjected to a heat treatment
patterns of the Mo-tailings are shown in Fig. 1.
from 25 to 1500 °C at a heating rate of 10 °C/min. Changes in the shape
of the samples with increasing temperature were photographed at 1-
min intervals using the image analysis system of the instrument syn-
2.2. Preparation of FC
chronized with the programming schedule of the HSM.
The effective thermal conductivity (λeff) was measured via laser
For the preparation of FC, Mo-tailings and SiC (0–3 wt%) were used
flash analysis (LFA 447, Netzsch). Discs with thickness of 1 mm and
as the raw materials; the prepared samples were named as S0, S0.5, S1,
diameter of approximately 12.7 mm were prepared with the sintered
S1.5, S2, S2.5, and S3, depending on the SiC content.
samples, and then they were coated with a thin graphite layer. After Xe
FC samples were prepared by mixing the components in a planetary
laser irradiation with a long pulse width of 450 μs, the increase in
mill (Tencan Powder, China) for 40 min. The particle sizes of the
temperature with time was recorded using an infrared detector. The
powders determined by laser particle size analyzer (BT-9300S,
temperature-time curve was then fitted with a radiation numerical
Dandong Bettersize Instruments Ltd., China) varied from 0.1 to 55 μm,
model by including the pulse correction, from which the thermal dif-
and the median diameter was 9.5 μm.
fusivity (α) was calculated. Specific heat (Cp) was measured using the
Cylindrical samples with diameter 55 mm and height 20 mm were
temperature-modulated differential scanning calorimetry (MDSC)
obtained by uniaxial pressing of the FC powders at 5 MPa.
method (TA Q2000, USA). All measurements were carried out in air at
Subsequently, the samples were dried at 110 °C for 12 h for the residual
room temperature. Finally, with the values of α, the thermal con-
water removal and then heated in an electric furnace at 1130–1160 °C
ductivity was calculated using Eq. (2).
for 30 min at a heating rate of 5 °C/min.
X-ray photoelectron spectroscopy (XPS) was used to confirm the λ eff = α∗ρ∗cP (2)
presence of metals in the samples and the different oxidation states of
the metals. The XPS profiles were recorded on a VG Escalab 200R
spectrophotometer equipped with a hemispherical electron analyzer 3. Results and discussion
using monochromatic Mg Kα radiation as the X-ray source. Samples S1-
3.1. Liquid-phase viscosity analysis of the mixture

A quantitative analysis of η of the samples during sintering is highly


important [21,26]. Parameter η, which is in direct correlation with the
selected foaming temperature, controls the maximum foam stability,
and in turn, the shape, size, and homogeneity of the pores. When the
foaming temperature is relatively high, η becomes extremely low
(i.e., < 103 Pa·s), which makes the maintenance of a stable foam
structure difficult, leading to the collapse of samples due to the coa-
lescence of pores and gravity effect [24]. On the contrary, a low tem-
perature corresponds to a high η, which makes it difficult to achieve
well-foaming samples that require adequate pressure within the bubble
to resist high surface tension of the silicate glass melt. Hence, during
expansion, the glass melt needs a suitable η, allowing the samples to
expand under the internal pressure caused by the release of the gas
Fig. 1. The XRD result of Mo-tailings. generated during the oxidation-reduction reaction of the foaming

16
Z. Tang, et al. Journal of Non-Crystalline Solids 513 (2019) 15–23

Table 2
HSM and η fixed point analysis of sample S2.5 (inset figure: characteristic stages at different temperatures).
Feature point Figure Temperature Log η Description
(°C) (pa·s)

TFrist Shrinkage(TF1) 1010 9.1 ± 0.1 The temperature at which the sample begins to shrink.

TMaximum Shrinkage(TM) 1108 7.8 ± 0.1 The temperature at which the sample shrinks to the minimum volume.

T Softening Point(TS1) 1163 6.3 ± 0.1 The temperature at which the sample corners start toround.

T Sphere Point(TS2) 1285 5.4 ± 0.1 The temperature at which the sample turns into spherical.

T Half Ball Point(TH) 1376 4.1 ± 0.1 The temperature at which the sample turns into hemispherical.

T Flow Point(TF2) 1468 3.4 ± 0.1 The temperature at which the sample turns into flow state.

agent. an increase in the SiC content from 0 to 3.0 wt%.


In this study, η was analyzed using an inverse proportionality em- Hence, the aforementioned results indicate that a higher SiC content
pirical relation, known as the Vogel-Fulcher-Tammann (VFT) equation is favorable for the reduction of the characteristic sintering temperature
[22,31–33]: and melt viscosity, which causes the foaming effect of the SiC particles.
There are two main mechanisms of the redox reaction of SiC. Firstly,
log η = A + B/(T − T0) (3) the SiO2 coating on the surface of SiC particles prevents SiC from
coming into contact with O2, therefore, only when O2/air passes
where A, B, and T0 are three empirical constants. They were calculated
through the SiO2 coating, the oxidation-reduction reaction between SiC
using the six characteristic viscosity points (Table 2 and Fig. 2).
and O2 occurs. This implies that the oxidation reaction rate of SiC
It is well known that the proper range of porosity is strongly in-
largely depends on the diffusion rate of oxygen. At lower sintering
fluenced by the composition of mixtures, which comprise various kinds
temperatures, the rate of oxygen diffusion through the SiO2 coating is
of primary raw materials and foaming agents. As mentioned above, the
only 10−14–10−15 cm2/s, which inhibits the oxidation reaction of SiC
suitable η range for the diverse-component samples varies between
particles. However, when the sintering temperature rises (above
103 Pa·s and 108 Pa·s [23–25]. Hence, these two values were taken as
1100 °C), the SiO2 coating is easily destroyed by the formation of the
the reference in this study. Fig. 2 shows the viscosity-temperature
liquid phase silicate, which causes an increase in the diffusion rate of
curves of the samples containing 0–3.0 wt% SiC (with an interval of
O2, thus promoting the reaction of SiC particles with oxygen and gen-
0.5 wt%). As observed, the two values simultaneously decrease with
erating a large amount of CO2 [17]. Because the embryo generates a
increasing SiC content. The decrease in the lower and upper limit
large amount of the liquid phase above 1100 °C, it is difficult for the gas
temperatures is about 97 °C and 110 °C, respectively, corresponding to

17
Z. Tang, et al. Journal of Non-Crystalline Solids 513 (2019) 15–23

Fig. 2. The viscosity-temperature curves of samples doped with different additional amounts of SiC (a) 0 wt% (b) 0.5 wt% (c) 1.0 wt% (d) 2.0 wt% (e) 2.5 wt% (f)
3.0 wt%.

to be discharged outward. Therefore, with increase in the amount of gas SiC content, the peak area and half-peak width of the Fe3+ peak in the
remaining in the embryo, the sample expands [34–36]. sample gradually decrease. However, the Fe2+ peak displays the op-
The aforementioned phenomenon was confirmed by XRD analysis. posite trend, indicating that the Fe3+ content in the sample decreases,
As shown in Fig. 3, all samples doped with different amounts of SiC and while the Fe2+ content increases. Because Fe2+ is a network-modifying
sintered at 1150 °C for 30 min consist of the quartz, anorthite, and iron ion, it destroys the [SieO] net structure and causes a decrease in η
oxide phases, in agreement with the XRD result of the Mo-tailings [33,37,38].
(Fig. 1). The high content of SiO2 in the Mo-tailings is attributed to the To further investigate the effect of SiC on η and shrinkage (also
significant precipitation of quartz, while the weak iron oxide traces called the densification process), we studied the swelling behavior of
(featuring only Fe3+, Fe2O3) are attributed to the Fe2O3 (about 4.51 wt samples with different SiC concentrations by high-temperature thermal
%) in the raw material (Table 1). These results reveal that high contents synthetic analysis with the HSM. As shown in Fig. 5, no obvious ex-
of SiC do not cause any crystalline phase transformation of the main pansion or shrinkage is observed from room temperature to approxi-
raw material. However, the intensities of the different diffraction peaks mately 1020 °C. However, at temperatures higher than 1020 °C (up to
gradually decrease with increasing SiC content, implying that SiC im- 1130 °C), shrinkage occurs at a low rate (i.e., densification rate (in
pedes crystallization, causing a decrease in η. On the other hand, as seen terms of dimensional variation) tending to zero), after which, the ex-
in Fig. 3, the oxidation-reduction reaction between SiC and Fe2O3 in the pansion attributed to the release of the gas generated during the oxi-
glass melt results in the transformation of Fe3+ to Fe2+, which causes a dation-reduction of the foaming agent begins. Nonetheless, the densi-
decrease in the precipitation of iron oxide. The XPS results (Fig. 4 and fication rate gradually increases as the SiC content increases from 1.0 to
Table 3) agree with the XRD results. Notably, with an increase in the 3.0 wt%, and the maximum shrinkage temperature decreases from

Fig. 3. The XRD patterns of FC with variable SiC content at 1150 °C for 30 min.

18
Z. Tang, et al. Journal of Non-Crystalline Solids 513 (2019) 15–23

Fig. 5. Shrinkage and swelling behavior of powder compacts with different SiC
contents; (A/A0 corresponds to the ratio of final area/initial area of the samples
powder compacts. The uncertainty associated with A/A0 is about 0.15%.)

the high η and low gas production, as aforementioned, lead to in-


sufficient foaming, and eventually, a pore structure with thick cell walls
and inhomogeneous pores is formed. With increase in the SiC content, η
decreases; consequently, the atoms gain sufficient mobility to diffuse
towards any vacancy concentration gradient [20]. Subsequently, the
oxidation reaction of SiC particles intensifies, leading to enhanced CO2
gas generation and pore growth acceleration; thereby, the bulk density
drastically decreases. However, at SiC contents > 2.5 wt%, a higher
amount of gas is generated, increasing the gas pressure and commu-
nicating pore number; consequently, the foam collapses and columns
loosen. Therefore, the density of S3 did not significantly decrease.
On the other hand, S0.5 exhibits the maximum thermal conductivity
because of the relatively large bulk density (Fig. 6(a), which is attrib-
uted to a lower foaming capacity and more heat-transfer channels.
However, with increasing SiC component, the thermal conductivity
Fig. 4. The XPS patterns of FC with variable SiC content at 1150 °C for 30 min.
sharply decreases to about 0.22 W/m·K. Nevertheless, S3 exhibits a
(a) 1.0 wt% (b) 2.0 wt% (c) 3.0 wt%.
higher thermal conductivity than S2.5. The reason is that although S3
has a lower bulk density, it possesses enhanced open porosity, which
1130 °C to 1099 °C. The decrease in the maximum shrinkage tempera- has a negative effect on the heat insulation property. Therefore, the
ture is attributed to the decrease in η with increasing SiC content. final heat-insulating properties of the FC are determined by not only the
bulk density but also the porosity and pore size of the samples.
3.2. Effect of bulk density on the thermal conductivity of FC The samples prepared in this study at 1150 °C, S0.5-S3, were not
much superior to those prepared by other researchers with regard to
The thermal conductivity of FC depends on the solid conduction, gas thermal conductivity and bulk density (Fig. 6(b)). However, the raw
conduction, thermal radiation, and convection [1]. Fig. 6(a) shows the materials we used were almost pure wastes and released no harmful
evolution of the sample bulk density and thermal conductivity with substances during the preparation, which was economical and en-
increment in the SiC content. In the sample sintered at 1130 °C, absence vironment-friendly. Note that the materials with thermal con-
of SiC, the gas production becomes limited, and consequently, the bulk ductivity < 0.25 W/m·K can be regarded as thermal insulators [39,40].
density becomes 2.42 g/cm3 [7,37], which is the general density of the Therefore, the preparation of FCs with potential applications from Mo-
identified porcelain tile (not shown in the diagram). However, the bulk tailings as a raw material, solves the problem of massive accumulation
density significantly decreases from 0.97 to 0.39 g/cm3 with increasing of Mo-tailings.
SiC content (Fig. 6(a)). Therefore, in order to obtain an FC with low
bulk density, the addition of SiC is necessary [17]. At 0.5 wt% of SiC,

Table 3
Comparative data of Fe3+ and Fe2+ atomic ratio in foam ceramic used fitting methods.
Element S1 S2 S3

Fe3+ Fe2+ Fe3+ Fe2+ Fe3+ Fe2+

Spectral line 2P3/2 2P3/2 2P3/2 2P3/2 2P3/2 2P3/2


Binding energy (eV) 711.10 713.60 711.10 713.60 711.10 713.60
FWHM 4.37 ± 0.31 0.82 ± 0.09 3.97 ± 0.65 3.22 ± 0.40 2.47 ± 0.17 3.73 ± 0.33
(eV)
Fe2+/Fe3+ 0.04 ± 0.005 0.59 ± 0.16 1.24 ± 0.03
(%)

Fe2+/Fe3+ = (Fe2+(Area)/Fe3+(Area)).

19
Z. Tang, et al. Journal of Non-Crystalline Solids 513 (2019) 15–23

Fig. 6. (a) Evolutions of bulk density and thermal conductivity with SiC contents. (Solid lines denote density, dashed lines denote thermal conductivity). (b) The
thermal conductivity variations of FC with bulk density from other reseaches.

Fig. 7. (a) The thermal conductivity variations of FC with porosity from other researches. (b) Evolution curves of thermal conductivity with different SiC contents. (c)
Evolution curves of porosity and thermal conductivity with different SiC contents. (d) Optical images of different SiC contents at 1150 °C.

3.3. Effect of porosity and pore size on the thermal conductivity of FC porosity of S2.5 at 1160 °C is lower than that at 1150 °C, which is the
same as that of S3.
As shown in Fig. 7(a), the relationship between porosity and Although, the porosities of S3 and S2 are similar, the thermal con-
thermal conductivity is inversely proportional. Fig. 7(b) can be divided ductivity of S3 is lower than that of S2, which is attributed to the larger
into two parts: the low-temperature part (1130 °C) and high-tempera- pore size of S3, resulting from the higher gas production and lower η.
ture part (1150 °C to 1160 °C). In the low-temperature part, the porosity Therefore, the effect of pore size on the thermal conductivity must be
increases with increasing SiC content because the addition of SiC con- taken into consideration, as well.
tributes to the gradual decrease in η and increase in the gas production, As shown in Fig. 8, S0.5 exhibits a narrow pore size distribution, and
which finally, cause an increase in the porosity, and in turn, a decrease the sizes of about 90% of pores are in the range 0–0.5 mm. Notably,
in the thermal conductivity. Nevertheless, for the samples sintered at with increase in the SiC content, the average pore size gradually in-
higher temperatures, with increasing SiC content, the gas production creases from 0.5 to 5.0 mm for S0.5-S3. In addition, with increase in the
increases and η decreases. In addition, at these temperatures, larger SiC content, η decreases, which provides more active sites for nuclea-
pores form due to the multiple porosities. As the number of larger pores tion, and the oxidation of SiC intensifies, leading to the generation of
increases, the pores rupture, and finally, the gas escapes. Therefore, the more gas; consequently, pores easily connect together to form larger

20
Z. Tang, et al. Journal of Non-Crystalline Solids 513 (2019) 15–23

Fig. 8. Pores size and distribution of samples with different SiC amounts and sintered at 1150 °C for 30 min. (a) 0.5 wt% (b) 1.0 wt% (c) 1.5 wt% (d) 2.0 wt% (e)
2.5 wt% (f) 3.0 wt%.

Fig. 9. SEM images of samples with different SiC amounts and sintered at 1150 °C for 30 min: (a) 0 wt% (b) 0.5 wt% (c) 1.0 wt% (d) 2.0 wt% (e) 2.5 wt% (f) 3.0 wt%.

pores. Moreover, the pore wall stretches when the gas pressure exceeds 3.4. Comparison between predicted and experimental values of thermal
the forces of surface tension and η, resulting in pore expansion. The gas conductivity
pressure in large pores is lower than that in small ones because the
pressure in a spherical pore is inversely proportional to the pore radius. The effective thermal conductivity of an FC can be predicted using
Consequently, larger pores grow faster than the smaller ones. However, an analytical model, in which the value is assimilated to a two-phase
a large pore size causes an uneven distribution of pores and high open system constituted by a dense solid skeleton and gas [27]. The pores are
porosity in the samples, as shown in Fig. 9(d)–(f); this is the reason for usually filled with CO2 and air. Notably, in the prepared FC samples,
the high thermal conductivity of S3. CO2 is the main entrapped gas. Therefore, in this study, we have con-
sidered the thermal conductivity of CO2 as a major parameter.
We tried to fit the experimental data to the existing models, namely,

21
Z. Tang, et al. Journal of Non-Crystalline Solids 513 (2019) 15–23

However, at porosities higher than 85%, the prediction starts to fail as


the porosity increases, possibly because the open porosity of the sample
gradually increases, resulting in an increase in the air content of the
sample.

4. Conclusions

In summary, the variation in η of the sintered samples with in-


creasing SiC content was studied using an HSM. The following con-
clusions were drawn according to the results presented and discussed in
this paper: with decreasing η, both the porosity and pore size increased,
and the thermal conductivity first decreased and then increased.
Experimental results agreed with the parallel-model-based predictions.
Thus, the preparation of FCs with a large amount of Mo-tailings could
Fig. 10. Effective thermal conductivity as a function of porosity for analytical facilitates potential advantages from both the economic and ecological
predictions and experimental measurements. Calculation Parameters: aspects.
λs = 1.62 W/m K and λp = 0.025 W/m K.
Acknowledgments
the parallel model [28], Hashin-Shtrikman model [28], Russel model
[27,41], and Landauer model [27,41], as shown in Eqs. (4)-(7). These This research was financially supported by the Inner Mongolia
models can be divided into two groups according to the assumptions University of Science and Technology Innovation Fund (No. 2016QDL-
proposed to describe the properties of the microstructure. The first 13).
group is constituted by the asymmetrical models: the parallel model,
Hashin-Shtrikman model, and Russel model, which consider the iso- Ethical statement
lated, cylindrical, and cubic inclusions dispersed in the matrix, re-
spectively. The foam ceramics prepared from a large amount of molybdenum
The parallel model: tailings (97–99.5 wt%) is successfully synthesized. The innovation of
the article is to characterize the change of viscosity in the process of
λ eff = (1 − φp ) ∗ λs + φp∗λp (4) tailings sintering and the effect of SiC on the thermal properties of
molybdenum tailings foam ceramics. The results showed that with in-
The Hashin-Shtrikman model:
creasing SiC content, the liquid-phase viscosity decreased, which in
∗ turn enhanced the overall expansion of the system. The influence of SiC
⎡ λp + 2λs + 2φp (λp − λs ) ⎤
λ eff = λs∗ ⎢
λ + 2λs − φp (λp − λs ) ⎥

(5)
content on the thermal conductivity of foamed ceramic was measured
⎣ p ⎦ by 4 theoretical models and the actual measurement method.
The Russel model:
2 ∗
References
λs∗ ⎡λs + φ 3 (λ
p − λs ) ⎤
λ eff = ⎣ ⎦
[1] M.B. Østergaard, R.R. Petersen, J. König, H. Johra, Y. Yue, Influence of foaming
(
λs + (λp − λs ) ∗ φ
2
3 −φ ) (6) agents on solid thermal conductivity of foam glasses prepared from CRT panel glass,
J. Non-Cryst. Solids 465 (2017) 59–64.
On the other hand, the second group comprises the symmetrical [2] A.C. Bento, E.T. Kubaski, T. Sequinel, S.A. Pianaro, J.A. Varela, S.M. Tebcherani,
Glass foam of macroporosity using glass waste and sodium hydroxide as the
models, such as the Landauer model, which consider a more continuous
foaming agent, Ceram. Int. 39 (2013) 2423–2430.
property of the second phase. Each phase exhibits a more or less con- [3] J. König, R.R. Petersen, Y. Yue, D. Suvorov, Gas-releasing reactions in foam-glass
tinuous character depending on its volume fraction. formation using carbon and MnxOy as the foaming agents, Ceram. Int. 43 (2017)
4638–4646.
The Landauer model:
[4] H. Wang, Y. Sun, L. Liu, R. Ji, X. Wang, Integrated utilization of fly ash and waste
1∗ glass for synthesis of foam/dense bi-layered insulation ceramic tile, Energy Build.
λ eff = ⎡ (λp∗ (3φ − 1)) + λs∗ (3φs − 1) 168 (2018) 67–75.
4 ⎣ [5] H. Wang, Z. Chen, R. Ji, L. Liu, X. Wang, Integrated utilization of high alumina fly
1
+ ([λp∗ (3φ − 1) + λs∗ (3φs − 1)]2 + 8λs∗λp ) 2⎤ ash for synthesis of foam glass ceramic, Ceram. Int. 44 (2018) 13681–13688.
⎦ (7) [6] G. Kyaw Oo D'Amore, M. Caniato, A. Travan, G. Turco, L. Marsich, A. Ferluga,
C. Schmid, Innovative thermal and acoustic insulation foam from recycled waste
where λs and λp are the thermal conductivities of the solid matrix and glass powder, J. Clean. Prod. 165 (2017) 1306–1315.
pores, respectively, and φs and φp are the corresponding volume frac- [7] M. Zhu, R. Ji, Z. Li, H. Wang, L. Liu, Z. Zhang, Constr. Preparation of glass ceramic
tions of these phases. foams for thermal insulation applications from coal fly ash and waste glass, Build
Mater. 112 (2016) 398–405.
It can be seen from Fig. 10 that several theoretical calculations and [8] J. König, R.R. Petersen, N. Iversen, Y. Yue, Suppressing the effect of cullet com-
actual measurements show the same trend: the thermal conductivity position on the formation and properties of foamed glass, Ceram. Int. 44 (2018)
decreases with increasing porosity. Interestingly, most models seem to 11143–11150.
[9] R.R. Petersen, J.K. Nig, M.M. Smedskjaer, Y. Yue, Effect of Na2CO3 as foaming agent
underestimate the actual thermal conductivity of the samples, com- on dynamics and structure of foam glass melts, J. Non-Cryst. Solids 400 (2014) 1–5.
pared with the experimental data. The analytical-theory model based [10] J. König, R.R. Petersen, Y. Yue, Fabrication of highly insulating foam glass made
on isolated inclusions, such as the Hashin-Shtrikman model, does not from CRT panel glass, Ceram. Int. 41 (2015) 9793–9800.
[11] J. König, R.R. Petersen, Y. Yue, Influence of the glass particle size on the foaming
yield a good prediction for the effective thermal conductivity of the FC process and physical characteristics of foam glasses, J. Non-Cryst. Solids 447 (2016)
prepared in this study. Moreover, the Landauer model, which considers 190–197.
the uneven distribution of the gas phase in the solid phase at a rela- [12] R. Lebullenger, S. Chenu, J. Rocherullé, O. Merdrignac-Conanec, F. Cheviré,
F. Tessier, A. Bouzaza, S. Brosillon, Glass foams for environmental applications, J.
tively lower porosity, fails to predict the thermal conductivity of the Non-Cryst. Solids 356 (2010) 2562–2568.
solid phase as well as the gas phase [28]. Hence, most models sum- [13] E. Bernardo, G. Scarinci, P. Bertuzzi, P. Ercole, L. Ramon, Recycling of waste glasses
marized above cannot be effectively applied to the FC, except the into partially crystallized glass foams, J. Porous. Mat. 17 (2010) 359–365.
[14] H.R. Fernandes, D.U. Tulyaganov, J.M.F. Ferreira, Preparation and characterization
parallel model. As depicted in Fig. 9, the experimental results of this
of foams from sheet glass and fly ash using carbonates as foaming agents, Ceram.
study closely agree with the predictions proposed by the parallel model.

22
Z. Tang, et al. Journal of Non-Crystalline Solids 513 (2019) 15–23

Int. 35 (2009) 229–235. conductivity of glasses foamed with MnO2, J. Non-Cryst. Solids 425 (2015) 74–82.
[15] X. Chen, A. Lu, G. Qu, Preparation and characterization of foam ceramics from red [28] S. Ren, J. Liu, A. Guo, W. Zang, H. Geng, X. Tao, H. Du, Mechanical properties and
mud and fly ash using sodium silicate as foaming agent, Ceram. Int. 39 (2013) thermal conductivity of a temperature resistance hollow glass microspheres/bor-
1923–1929. osilicate glass buoyance material, Mater. Sci. Eng. A 674 (2016) 604–614.
[16] Y. Guo, Y. Zhang, H. Huang, K. Meng, K. Hu, P. Hu, X. Wang, Z. Zhang, X. Meng, [29] D. Zorić, D. Lazar, O. Rudić, M. Radeka, J. Ranogajec, H. Hiršenberger, Thermal
Novel glass ceramic foams materials based on red mud, Ceram. Int. 40 (2014) conductivity of lightweight aggregate based on coal fly ash, J. Therm. Anal.
6677–6683. Calorim. 110 (2012) 489–495.
[17] H. Wang, Z. Chen, L. Liu, R. Ji, X. Wang, Synthesis of a foam ceramic based on [30] S. Arcaro, B.G.D.O. Maia, M.T. Souza, F.R. Cesconeto, L. Granados,
ceramic tile polishing waste using SiC as foaming agent, Ceram. Int. 44 (2018) A.P.N.D. Oliveira, Thermal insulating foams produced from glass waste and banana
10078–10086. leaves, Mater. Res. 19 (2016) 1064–1069.
[18] X. Xi, L. Xu, A. Shui, Y. Wang, M. Naito, Effect of silicon carbide particle size and [31] N. Sasmal, M. Garai, B. Karmakar, Preparation and characterization of novel
CaO content on foaming properties during firing and microstructure of porcelain foamed porous glass-ceramics, Mater. Charact. 103 (2015) 90–100.
ceramics, Ceram. Int. 40 (2014) 12931–12938. [32] D. Hesky, C.G. Aneziris, U. Gro, A. Horn, Water and waterglass mixtures for foam
[19] X. Xi, A. Shui, Y. Li, Y. Wang, H. Abe, M. Naito, Effects of magnesium oxychloride glass production, Ceram. Int. 41 (2015) 12604–12613.
and silicon carbide additives on the foaming property during firing for porcelain [33] Y. Liao, C. Huang, Glass foam from the mixture of reservoir sediment and Na2CO3,
ceramics and their microstructure, J. Eur. Ceram. Soc. 32 (2012) 3035–3041. Ceram. Int. 38 (2012) 4415–4420.
[20] D.U. Tulyaganov, H.R. Fernandes, S. Agathopoulos, J.M.F. Ferreira, Preparation and [34] X. Xi, H. Xiong, A. Shui, M. Huang, S. Xiao, H. Lin, Foaming inhibition of SiC-
characterization of high compressive strength foams from sheet glass, J. Porous. containing porcelain ceramics by using Si powders during sintering, J. Eur. Ceram.
Mat. 13 (2006) 133–139. Soc. 37 (2017) 5044–5050.
[21] H.R. Fernandes, F. Andreola, L. Barbieri, I. Lancellotti, M.J. Pascual, J.M.F. Ferreira, [35] R. Ji, Z. Zhang, Y. He, L. Liu, X. Wang, Synthesis, characterization and modeling of
The use of egg shells to produce Cathode Ray Tube (CRT) glass foams, Ceram. Int. new building insulation material using ceramic polishing waste residue, Build.
39 (2013) 9071–9078. Mater. 85 (2015) 119–126.
[22] C. Mugoni, M. Montorsi, C. Siligardi, F. Andreola, I. Lancellotti, E. Bernardo, [36] A. Shui, X. Xi, Y. Wang, X. Cheng, Effect of silicon carbide additive on micro-
L. Barbieri, Design of glass foams with low environmental impact, Ceram. Int. 41 structure and properties of porcelain ceramics, Ceram. Int. 37 (2011) 1557–1562.
(2015) 3400–3408. [37] J. García-Ten, A. Saburit, E. Bernardo, P. Colombo, Development of lightweight
[23] M.T. Souza, B.G.O. Maia, L.B. Teixeira, K.G. de Oliveira, A.H.B. Teixeira, porcelain stoneware tiles using foaming agents, J. Eur. Ceram. Soc. 32 (2012)
A.P. Novaes De Oliveira, Glass foams produced from glass bottles and eggshell 745–752.
wastes, Process. Saf. Environ. 111 (2017) 60–64. [38] I. Ponsot, E. Bernardo, Self glazed glass ceramic foams from metallurgical slag and
[24] G. Scarinci, G. Brusatin, E. Bernardo, Glass foams, in: M. Scheffler, P. Colombo recycled glass, J. Clean. Prod. 59 (2013) 245–250.
(Eds.), Cellular Ceramics: Structure, Manufacturing, Properties and Applications, [39] R. Zhang, J. Feng, X. Cheng, L. Gong, Y. Li, H. Zhang, Porous thermal insulation
Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, FRG, 2005, p. 166. materials derived from fly ash using a foaming and slip casting method, Energy.
[25] H.R. Fernandes, D.D. Ferreira, F. Andreola, I. Lancellotti, L. Barbieri, Buildings. 81 (2014) 262–267.
J.M.F. Ferreira, Environmental friendly management of CRT glass by foaming with [40] S.H. Wang, Construction Materials Science, China Construction Industry Publisher,
waste egg shells, calcite or dolomite, Ceram. Int. 40 (2014) 13371–13379. Beijing, 1988.
[26] R.R. Petersen, J. König, Y. Yue, The viscosity window of the silicate glass foam [41] B. Nait-Ali, K. Haberko, H. Vesteghem, J. Absi, D.S. Smith, Thermal conductivity of
production, J. Non-Cryst. Solids 456 (2017) 49–54. highly porous zirconia, J. Eur. Ceram. Soc. 26 (2006) 3567–3574.
[27] R.R. Petersen, J.K. Nig, Y. Yue, The mechanism of foaming and thermal

23

You might also like