100% found this document useful (1 vote)
43 views12 pages

Wang 2022

This research investigates the impact of varying amounts of palygorskite on the thermal behaviors and properties of fired clay bricks. The study finds that adding palygorskite enhances compressive strength and reduces porosity, with an optimal addition of 30 wt% yielding the best results. Additionally, the thermal analysis reveals significant changes in mineral composition and physical properties during the firing process.

Uploaded by

souka bed
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
43 views12 pages

Wang 2022

This research investigates the impact of varying amounts of palygorskite on the thermal behaviors and properties of fired clay bricks. The study finds that adding palygorskite enhances compressive strength and reduces porosity, with an optimal addition of 30 wt% yielding the best results. Additionally, the thermal analysis reveals significant changes in mineral composition and physical properties during the firing process.

Uploaded by

souka bed
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 12

Applied Clay Science 216 (2022) 106384

Contents lists available at ScienceDirect

Applied Clay Science


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

Research paper

Influence of palygorskite on in-situ thermal behaviours of clay mixtures and


properties of fired bricks
Sen Wang a, b, e, Lloyd Gainey c, Xiaodong Wang a, b, Ian D.R. Mackinnon b, d, Yunfei Xi a, b, e, *
a
Central Analytical Research Facility (CARF), Queensland University of Technology (QUT), Brisbane, Queensland 4001, Australia
b
Centre for Clean Energy Technologies and Practices & Centre for Materials Science, Queensland University of Technology (QUT), Brisbane, Queensland 4001, Australia
c
Austral Bricks, Rochedale, Queensland 4123, Australia
d
School of Earth and Atmospheric Sciences, Faculty of Science, Queensland University of Technology (QUT), Brisbane, Queensland 4001, Australia
e
School of Chemistry and Physics, Faculty of Science, Queensland University of Technology (QUT), Brisbane, Queensland 4001, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Clay mixtures have been added with a palygorskite containing clay at 0 wt% to 50 wt% for brick button firing.
Palygorskite Through an effective integration of in-situ temperature controllable XRD, thermogravimetric and dilatometric
Fired clay brick analyses, the real-time thermal behaviours of clay mixtures were studied. Also, the physical, mechanical and
Cordierite
thermal performance of obtained fired brick buttons were investigated. The addition of palygorskite containing
Phase transition
Compressive strength
clay has increased the content of amorphous phases in fired brick buttons, while quartz, cristobalite, mullite and
Thermal behaviour hematite were suppressed. Cordierite, spinel and magnetite are new minerals developed in obtained brick but­
tons. From 25 to 1150 ◦ C, the brick green bodies underwent five mass loss steps and six dilatometric/contrac­
tions. The fibrous morphology of palygorskite no longer existed after 1150 ◦ C firing. Instead, larger-sized acicular
cordierite particles are developed. Adding palygorskite has resulted in lower porosity, higher drying shrinkage,
firing shrinkage and compressive strength in fired buttons due to reinforced liquid-phase sintering. The effect of
palygorskite on vitrification tended to be non-linear and non-uniform. In addition, the colours of brick buttons
changed from “light camel” to “black olive” with palygorskite content. Since cracks were developed on the brick
buttons added with 40 wt% and 50 wt% palygorskite containing clay, the optimum addition dosage was set at 30
wt%. The obtained compressive strength and thermal diffusivity (25 ◦ C) at this addition amount are 153.8 MPa
(24.5% higher than the control sample) and 0.788 mm2/s, respectively.

1. Introduction investigations on initial clayey soils (e.g., mineralogical and chemical


compositions, particle size, and thermal properties) the most urgent and
Fired clay brick is a type of classic brick form, considered to be a valuable considerations before actual production.
long-lasting and robust ceramic building material. It accounts for As the most crucial component in clayey soils, clay minerals can
75–80% of the global brick market and has been widely used for walls, impart the required plasticity so the raw materials can be shaped as
paving, flooring, roads etc. (Maithel et al., 2016). Since fired clay brick is bricks after pressing or extrusion. In addition, vitrification of clay mix­
produced from clayey soils heated at 700–1200 ◦ C, the quality of ob­ tures at high temperatures can bond mineral particles, reduce connec­
tained products mainly depends on the nature of starting material and tive porosity and produce a dense structure (Cultrone et al., 2004). The
firing process (Cultrone and Sebastian, 2009; Elavarasan et al., 2021; interpretation of the roles of clay minerals during brick firing is essential
Johari et al., 2010; Vasic et al., 2021b). In modern brick production, because thermal behaviours of clay minerals are closely related to the
although different clayey soils with variable recipes are adopted, the properties of obtained brick products (Chalouati et al., 2020). In addi­
procedures for manufacturing bricks in one factory are generally fixed to tion, different types of clay minerals behave differently when heated
simplify the production and save costs. This one-fits-all approach inev­ (Vasic et al., 2017; Wang et al., 2021).
itably leads to defects (e.g., cracking, bloating, efflorescence, and Palygorskite (also known as attapulgite) is a hydrated magnesium-
deformation etc.) when inappropriate clayey soils are utilised, making aluminium silicate clay mineral with an ideal chemical formula (Mg,

* Corresponding author at: Queensland University of Technology, Room 605C, Level 6, P block, Gardens Point Campus, Brisbane, QLD 4001, Australia.
E-mail address: y.xi@qut.edu.au (Y. Xi).

https://doi.org/10.1016/j.clay.2021.106384
Received 19 October 2021; Received in revised form 14 December 2021; Accepted 16 December 2021
Available online 24 December 2021
0169-1317/© 2021 Elsevier B.V. All rights reserved.
S. Wang et al. Applied Clay Science 216 (2022) 106384

Al)5(Si, Al)8O20(OH)2⋅8H2O (Wang et al., 2020). It is generally neo­ wt%, 10 wt%, 20 wt%, 30 wt%, 40 wt% and 50 wt%, respectively, which
formed in soils and paleosols under dry or semi-dry climates (Liu et al., are denoted as CM, CMP10, CMP20, CMP30, CMP40 and CMP50,
2013). The occurrence of palygorskite in nature is associated with three accordingly. 4 mL of tap water was added accurately to each mixture
conditions (Chen et al., 2004; Galan, 2006; Zhang et al., 2018b): 1) in using an automatic pipettor. After a manual stirring for ~5 min, samples
modern soils affected by groundwater of pH 7–8 and abundant salinity, were pressed into a steel mould using a hydraulic press (5 tons for 2
2) in soils and paleosols with clear and significant textural transitions, min). The shaped buttons (denoted as green buttons) were obtained, and
and 3) in calcretes/caliches. Due to the nanofibrous micromorphology, each had a diameter of 4 cm and a thickness of ~2 cm.
palygorskite has been used as supporting materials for nanocatalysts The green buttons were dried at 110 ◦ C for 24 h. Then dried buttons
(Ezzatahmadi et al., 2017; Xi et al., 2014); it is also applied to reinforce were transferred to a Carbolite Gero AAF 11/3 furnace for firing. The
numerous products, including plastics (Garcia et al., 2009; Zhang et al., firing temperature was 1150 ◦ C with a heating rate of 5 ◦ C/min and
2016), rubbers (Tang et al., 2016), membranes (Ding et al., 2021; Ji soaking at 1150 ◦ C for 2 h. After the firing was completed, the furnace
et al., 2015; Zhang et al., 2019a) and coatings (Zhang et al., 2018a). was turned off and samples were cooled down to room temperature in
Specifically, palygorskite can appear in the raw material for brick the furnace. 1150 ◦ C (rather than 950 ◦ C) was chosen because this
manufacturing, and its mass percentage can reach as high as 50% temperature was customized for Queensland clays rich in well-
(Awadh and Abdullah, 2011). Loutou et al. (2019) have prepared fired crystallised clay minerals but with fewer fluxes. Local brick manufac­
bricks using clayey sediment which contains 22.49 wt% of palygorskite turers routinely use this higher temperature which was also discussed by
and obtained acceptable water absorption capacity, firing shrinkage and Wang et al. (2021).
flexural strength. Zhang et al. (2019b) reported that palygorskite could
improve the flexural strength and fracture toughness of ceramic prod­ 2.2. Characterisation methods
ucts by 10.9% and 4.7%, respectively. However, to the best of the au­
thors' knowledge, there is no interpretation of the function of A Bruker D8 Advance diffractometer (Co Kα source at 1.7902 Å, 35
palygorskite during the firing process of brick production so far, as well kV, 40 mA) was used to obtain X-ray diffraction (XRD) patterns from 2◦
as the effects of this clay mineral on obtained brick properties. A sig­ to 90◦ 2θ with a step size of 0.02◦ and a scanning speed of 1.5◦ 2θ/min. A
nificant knowledge gap exists particularly from a combined perspective 10 wt% of internal standard (corundum) was used for quantitative phase
of mineralogy and material science, which makes it challenging to analysis. Clay slides were prepared through the gravity sedimentation
provide a practical reference for industrial production. experiment to determine clay minerals in CM. Detailed procedures are
In addition, few researchers have focused on the morphologic evo­ shown in the supplementary document. For the in-situ XRD measure­
lution of palygorskite nanofibers at high temperatures - particularly ment, a Rigaku Smartlab X-ray diffractometer (Cu Kα source at 1.54184
whether the fibrous shape can be retained to strengthen the mechanical Å, 40 kV, 40 mA) equipped with an Anton Paar HTK-1200 N chamber
performance. Some of the current conclusions are controversial. For and a Rigaku Reactor-X chamber was used. XRD patterns were collected
example, Chen et al. (2006) demonstrated that palygorskite fibres con­ from 3◦ to 45◦ 2θ with a step size of 0.02◦ and a scanning speed of 9◦ 2θ/
tract and bend at 1000 ◦ C to form an “earthworm” appearance. Zhang min under a 100 mL/min dry airflow. The temperature was increased
et al. (2019b) observed the remnants of palygorskite fibres at 1240 ◦ C, from 30 ◦ C to 1150 ◦ C using a heating rate of 5 ◦ C/min, and the powder
while Boudriche et al. (2012) and Wang et al. (2020) reported the melt sample heights were aligned at each temperature to count for the sample
and disappearance of initial shape at 1000 ◦ C and 1100 ◦ C, respectively. and sample holder expansions. A PANalytical AXIOS 1 kW wavelength
Thus, a systematic investigation of the thermal response of palygorskite dispersive X-ray fluorescence (XRF) spectrometer was utilised to mea­
becomes essential. Special concerns should be put especially on a tem­ sure the chemical compositions. The loss on ignition (LOI) was calcu­
perature range higher than 1000 ◦ C. lated before and after calcination at 1050 ◦ C for 2 h.
In this study, a palygorskite containing clay at different dosages A Netzsch STA 4493F was used to perform thermogravimetric
(0–50 wt%) has been added into a clay mixture for brick button making analysis (TGA) at 5 ◦ C/min from 25 ◦ C to 1150 ◦ C in a 100 mL/min
at 1150 ◦ C. Through effective integration of in-situ XRD, thermogravi­ airflow. The temperature was kept at 1150 ◦ C for 2 h. The porosities of
metric and dilatometric analyses, the effects of palygorskite on real-time fired brick buttons were investigated through a combination of water
thermal behaviours have been interpreted, and the microstructure and absorption experiments (a routine practice used by brick manufacturers)
physical-mechanical properties of obtained brick specimens have been and N2 isothermal adsorption and desorption measurement. For brick
investigated. This paper aims to reveal changes and corresponding manufacturers, the cold water adsorption (CWA) test is usually used to
mechanisms in the presence of palygorskite in the brickmaking process estimate the number of pores in the brick filled by capillary action
and provide suggestions for enhanced industrial production of higher- (approximately 10 μm or less in size), while the boiling water adsorption
performing brick products. (BWA) test can evaluate smaller pores that are not accessible through
capillary force. N2 isothermal absorption and desorption measurements
2. Materials and methods were performed using the Tristar II 3020 and a 99-point Brunauer-
Emmett-Teller (BET) method. Before the test, about 1.0 g of sample
2.1. Preparation of fired brick buttons was degassed at 150 ◦ C for 24 h. The average pore size is estimated on
the desorption branch by the Barrett-Joyner-Halenda (BJH) method.
A clay mixture (denoted as CM) was collected from Austral Bricks A field emission scanning electron microscope (FESEM, JEOL 7001F)
Pty. Ltd. in Queensland, Australia. CM is composed of five clays/shales was used to image the surface morphology of the brick buttons using
mined from local Late Permian and Triassic coal basins. These clays/ backscattered electron (BSE) imaging. The elemental composition was
shales are rich in clay minerals altered from volcanic rocks (i.e., granite collected through an X-Max 80 (Oxford, UK) energy-dispersive X-ray
and tuff). Less than 5 wt% of perlite (an amorphous volcanic glass) and spectroscopy (EDS) system. The SEM sample preparation procedures
brick grogs (crushed fired bricks) were added to CM as a routine in­ have been included in the supplementary document.
dustrial practice. In addition, a natural palygorskite containing clay Small rectangular bars (2 × 0.5 × 0.5 cm) were cut from the brick
(denoted as P) was purchased from Hudson Marketing Pty. Ltd., buttons and measured in a Netzsch 402C dilatometer for the thermal
Australia. P comes from Lake Nerramyne in Western Australia. expansion/shrinkage. The same heating program and atmosphere are
Before use, both CM and P were dried at 40 ◦ C for 24 h. Then samples used as in-situ XRD and TGA measurements. A Netzsch LFA 467 flash
were crushed until they could pass a 160 μm sieve. A digital balance was device was used to measure the thermal diffusivity on brick pellets (Φ1 x
used to weigh different amounts of CM and P for mixing. The total 0.5 cm). Drying and firing shrinkages were determined according to the
amount of the mixture was 40 g, and the P proportions were set up to 0 change in the diameter of the brick button with ten positions chosen and

2
S. Wang et al. Applied Clay Science 216 (2022) 106384

average values were calculated. The 24-h cold water absorption, 5-h 10.58 Å], followed by minor amounts of two other clay minerals -
boiling water absorption and compressive strength tests were conduct­ kaolinite [PDF 00-058-2005, d001 = 7.17 Å] (13.22%) and illite/mica
ed according to ASTM Standards (ASTMC67/C67M, 2020). The instru­ [PDF 04–014-1813, d002 = 9.97 Å] (5.15%). Quartz [PDF 01-079-1906,
ment used for the compressive test was 300-LX Universal Testing d101 = 3.34 Å] only constitutes 6.71%, but P contains 9.50% of dolomite
System, Instron (USA). [PDF 01-081-8226, d104 = 2.89 Å] (Fig. 1a and Table 1). In addition, the
amorphous content in P (18.34%) is significantly higher than that in CM
3. Results and discussion (9.95%) because clay minerals usually exhibit poor crystallinity. Ac­
cording to Fig. 1b, kaolinite flakes are oriented parallel in CM accom­
3.1. Raw material mineralogy and chemistry panied by angular quartz grains. Palygorskite shows fibrous morphology
in P with single fibre widths of ~100 nm and length ≤ 1 μm (Fig. 1c).
According to the XRD patterns in Fig. 1a, kaolinite [PDF 00-058- In terms of chemical composition, the major non-oxygen elements in
2004, d001 = 7.14 Å], quartz [PDF 04-012-0490, d101 = 3.34 Å], and both CM and P are Si and Al. However, CM contains higher Si and Al
illite/mica [PDF 04-014-1813, d002 = 9.90 Å] are predominant minerals than P (Table 2). The lower Si content in P is attributed to a lower
in CM, with calculated mass percentages of 34.65%, 31.22% and amount of quartz. Al2O3 is generally related to clay minerals, but since
11.64%, respectively (Table 1). A small amount of mixed layer illite- the major clay mineral in P is palygorskite which contains both Mg and
smectite (3.28%) has also been identified through the clay separation Al, P shows a lower amount of Al2O3. In addition, compared with CM, P
experiment (Fig. 1d). K-feldspar [PDF 01-071-0955, d040 = 3.24 Å] ac­ exhibits higher concentrations of Fe, alkalis (Na and K) and alkaline-
counts for 2.45 wt% in CM, while plagioclase [PDF 00-009-0466, d002 = earth (Ca and Mg) oxides. These elements can act as fluxes to promote
3.19 Å] accounts for 0.52 wt%. A Ti-containing mineral is anatase [PDF partial or complete vitrification of the green body during firing (Rehman
04-001-7641, d101 = 3.52 Å], and trace amounts of mullite [PDF 04-016- et al., 2020). According to the mineralogy, Na in P mainly comes from
1588, d110 = 5.37 Å] and cristobalite [PDF 04-008-7742, d101 = 4.04 Å] halite [PDF 01-070-2509, d200 = 2.82 Å]; Mg is related to palygorskite
were added from the brick grogs. and dolomite, while Ca derives from dolomite.
Compared to CM, P exhibits distinct mineralogy with a high con­
centration (40.92 wt%) of palygorskite [PDF 00–031-0783, d110 =

Fig. 1. XRD patterns of CM and P (a); SEM images of CM (b) and P (c); XRD patterns of oriented clay slides: CM (d) and P (e); Legend: Qz = quartz, Ilt = illite/mica,
Kln = kaolinite, Plg = palygorskite, Mul = mullite, Ant = anatase, Crn = corundum (internal standard), K-fsp = K-feldspar, Pl = plagioclase, Cal = calcite, Dol =
dolomite, and Hl = halite. Mineral abbreviations after (Whitney and Evans, 2010).

3
S. Wang et al. Applied Clay Science 216 (2022) 106384

Table 1
Mineralogical compositions of raw materials (wt%).
Qz Kln Ilt Plg Ilt-Sme K-fsp Pl Cal Dol Ant Mul Crs Hl Amp

CM 31.22 34.65 11.64 n.d. 3.28 2.45 0.52 1.03 n.d. 1.54 2.51 1.21 n.d. 9.95
P 6.71 13.22 5.15 40.92 n.d. 2.62 1.01 n.d. 9.50 1.01 n.d. n.d. 1.52 18.34

Note: Qz = quartz, Kln = kaolinite, Ilt = illite/mica, Plg = palygorskite, Ilt-Sme = mixed layer illite-smectite, K-fsp = K-feldspar, Pl = plagioclase, Cal = calcite, Dol =
dolomite, Ant = anatase, Mul = mullite, Crs = cristobalite, Hl = halite, Amp = amorphous, and n.d. = not detected. Mineral abbreviations after (Whitney and Evans,
2010).

Table 2
Chemical compositions of raw materials (wt%).
SiO2 Al2O3 Fe2O3 TiO2 CaO MnO K2 O MgO Na2O BaO P2O5 SO3 LOI.

CM 60.09 22.12 2.93 0.86 0.76 0.02 1.30 0.52 0.18 0.04 0.04 0.06 10.76
P 46.78 13.96 4.70 0.71 3.47 0.04 1.93 6.36 1.82 0.07 0.03 0.28 20.16

3.2. Real-time phase transition and quantitative phase analysis from palygorskite by heating alone but at higher temperatures.
In this study, enstatite/clinoenstatite can only be observed when
On the in-situ XRD patterns of CM and CMP20 (Fig. 2), three humps palygorskite was fired alone without mixing with CM, and its content
centred at 4.4◦ , 6.7◦ and 13.5◦ 2θ are from the Kapton and graphite increases with temperature (XRD patterns after firing at 850 ◦ C, 950 ◦ C
window material of the HTK-1200 N chamber, not associated with the and 1050 ◦ C in Fig. S2 in the supplementary document and 1150 ◦ C in
samples. These humps can be clearly seen on the XRD pattern of a Fig. 3). This is consistent with that observed by Kulbicki (1959) which
standard corundum sample measured within the HTK-1200 N chamber fired palygorskite at above 850 ◦ C. Interestingly, when palygorksite was
(Fig. S1 in the supplementary document). Since the 6.7◦ 2θ hump fired with CM, no enstatite/clinoenstatite is identified probably because
overlaps with palygorskite basal 110 reflections in the low-temperature neighbouring minerals/elements can affect the evolution of clay mineral
region, another batch of in-situ XRD patterns was collected using a (palygorskite in this case) at high temperatures and inhibit the forma­
Rigaku Reactor-X chamber equipped with a beryllium window up to tion of enstatite/clinoenstatite, which is consistent with Wang et al.
750 ◦ C to reveal the presence of palygorskite (Fig. 3b). According to (2021).
Fig. 2a and b, most minerals in CM and CMP20 experienced similar According to the quantitative phase analysis results in Table 3,
phase evolutions, thus only typical phase changes in CMP20 are inter­ quartz and amorphous fraction are two abundant phases in fired CM,
preted below. consistent with previous literature (Viani et al., 2018). 30.56% of
The 110 reflection of palygorskite disappears at 460 ◦ C (Fig. 3b) due mullite corresponds to the dominant kaolinite in the raw material
to the removal of structural H2O and formation of a folded structure. (Table 3). Other identified phases are cristobalite (7.22%), hematite
This change has also been reported by (Post and Heaney, 2008). In (2.65%), anorthite (1.25%) and anatase (0.26%). The addition of P
Fig. 2b, kaolinite dehydroxylates between 460 and 610 ◦ C indicated by decreases the percentage of quartz content in the starting material; thus,
the disappearance of the 001 reflection (7.17 Å, 12.3◦ 2θ), while illite/ less quartz and cristobalite are obtained. Also, mullite is reduced to only
mica persists, showing the 002 reflection (9.97 Å, 8.7◦ 2θ) until 1020 ◦ C. 8.19% in fired CMP50 due to: 1) presence of more Ca because of the
The dehydroxylation of illite/mica is identified by a slight increase of existence of dolomite in the raw material, which prevents the formation
d002 value at 625–820 ◦ C. A detailed explanation of the dehydroxylation of mullite but yields anorthite instead (El Ouahabi et al., 2015; Zhou
of kaolinite and illite can be found in the work by Wang et al. (2021). At et al., 2010) and 2) decrease of Al concentration after P has been
460–660 ◦ C, quartz shifts its reflections to lower angles due to a tran­ introduced. An Al-deficient environment can facilitate Si-rich amor­
sition from α- to β-phase (Fig. 2b). This process results in a sharp phous content since Si cannot coordinate with Al to produce mullite. In
expansion of the brick body which has been observed on dilatometry addition, a higher amount of flux because of P addition will also enhance
curves (Fig. 4c). In addition, calcite and dolomite are two carbonates the liquefaction of aluminosilicate, increasing the amorphous content in
decomposing at 570–645 ◦ C and 460–570 ◦ C, respectively. A weak fired CMP50 to 59.13% (Table 3).
reflection coming from the 200 plane of lime is found at around 37.3◦ 2θ Interestingly, the content of hematite declines with the increase of P
in Fig. 2a and b. content, although P brings more Fe to the clay mixture (Table 2). This
In terms of high-temperature related phases, mullite becomes trend corresponds well to the obtained appearance of brick in which no
noticeable after 945 ◦ C through recrystallisation of clay minerals (e.g., red colour is observed (Fig. 6). However, less hematite is formed which
kaolinite and illite/mica). Cristobalite is formed from β-quartz and/or is in contrast to the common sense view that Fe usually becomes Fe2O3
amorphous SiO2 at 1100 ◦ C. Cordierite and spinel in P added brick in an oxidising environment. A possible explanation is that Fe is
buttons (Fig. 3) are originated from decomposed palygorskite. The substituting Mg in cordierite and spinel, evidenced by the SEM/EDS data
decomposed dolomite may also provide MgO to form cordierite and in point 10 in Table 4 and Fig. 5. In addition, a small portion of Fe forms
spinel as described by Zhou et al. (2010). Spinel is formed by the reac­ magnetite at high temperatures (Table 3). The formation of trace
tion: MgO + Al2O3 → MgAl2O4 (El Ouahabi et al., 2015). Cordierite is magnetite in fired CMP40 and CMP50 is probably due to two reasons: 1)
only visible in ex-situ XRD patterns (Fig. 3a) due to the lower limit of Gas diffusion during firing is a two-way process. The addition of P has
detection compared to in-situ XRD. Mumpton and Roy (1958), Hirsiger increased the gas diffusing out of the clay body, confirmed by a higher
et al. (1975) and Wu et al. (2015) also observed the cordierite phase mass loss of P-bearing samples during firing (Fig. 4a). More gas coming
after firing palygorskite containing clays at 1150–1200 ◦ C. This trans­ out results in deficient oxygen diffusing into the clay body; 2) Compared
formation may throw light on the genesis of cordierite in nature. Rein­ to CM, P contains a higher amount of Fe. The P-added clay mixtures need
hardt (1987) reported the formation of cordierite from evaporitic clay more oxygen to be oxidised, thus forming magnetite when O2 is rela­
rich in palygorskite, sepiolite, chlorite, talc and corrensite in Queens­ tively insufficient.
land, Australia. These clays have undergone metamorphism under
620–650 ◦ C and 4 kbar of pressure. However, no additional pressure was
applied in this study, indicating that cordierite can also be generated

4
S. Wang et al. Applied Clay Science 216 (2022) 106384

Fig. 2. 2D in-situ XRD patterns of CM (a) and CMP20 (b). Legend: Ilt = illite/mica, Kln = kaolinite, Qz = quartz, Mul = mullite, Crs = cristobalite, Ant = anatase, Cal
= calcite, Dol = dolomite, Sa = sanidine, Mc = microcline, Ab = albite, Hem = hematite, and Spl = spinel. Mineral abbreviations after (Whitney and Evans, 2010).

3.3. Thermogravimetric and dilatometric analyses palygorskite channels. According to Gonzalez et al. (1990), palygorskite
contains water in four types: adsorbed water, zeolite water within the
Thermal mass loss is crucial for brick production because it is related channels, bound water coordinated to the edge octahedral cations, and
to the density of obtained products. Since mass loss is also accompanied structural water (e.g. Mg-OH, Al-OH and Fe-OH groups). In “Step 2”
by gas release, “bloating” may happen if the brick body loses too much (155–235 ◦ C), the first part of bound water is released from palygorskite
mass. Generally, a good quality brick exhibits less than 15% mass loss in agreement with (Chen et al., 2006). As the temperature increases, a
when fired (Tantawy and Mohamed, 2017). In this study, 8.5% of mass major mass loss appears at ~270–620 ◦ C in “Step 3”, which rises from
has been lost for CM. Addition of P promotes this value to 9.03% 6.90% (CM) to 7.51% (CMP10), 7.83% (CMP20), 8.41% (CMP30),
(CMP10), 9.52% (CMP20), 10.24% (CMP30), 11% (CMP40) and 8.88% (CMP40) and 9.34% (CMP50) with P content. In the control
11.67% (CMP50) (Fig. 4a). In addition, five steps are identified from 25 sample (CM), “Step 3” is attributed to kaolinite dehydroxylation and
to 1150 ◦ C on the first derivative (dL/dt) curves in Fig. 4b. calcite decomposition as indicated by in-situ XRD patterns and (Cheng
“Step 1” from room temperature to 125 ◦ C is assigned to the removal et al., 2015; Zuo et al., 2018), while in P added samples, dolomite
of absorbed water from the sample surface and zeolitic water from decomposition and evaporation of the second part of palygorskite bound

5
S. Wang et al. Applied Clay Science 216 (2022) 106384

Fig. 3. Ex-situ XRD patterns of fired brick buttons (a) and in-situ XRD patterns of palygorskite at 25–750 ◦ C (b). Legend: En = enstatite, Crd = cordierite, Mul =
mullite, Qz = quartz, Crs = cristobalite, Crn = corundum (internal standard), Hem = hematite, Plg = palygorskite, and Spl = spinel. Mineral abbreviations after
(Whitney and Evans, 2010).

Fig. 4. Mass and dimensional changes of clay mixture samples by TGA (a), DTG (b), thermal dilatometry and dilatometric curves highlighted in the 25 to 750 ◦ C
range (c and its inset, respectively), and the first derivative (dL/dt) of thermal dilatometry (d).

water are also involved. Cheng et al. (2011) also reported that the bound dehydroxylation is not noticeable. Finally, the mass loss in “Step 5” is
water of palygorskite was removed in two steps, similar to this research. ascribed to dehydroxylation of the -OH units (structural water) in
Furthermore, “Step 4” at 605–690 ◦ C is due to dehydroxylation of illite/ palygorskite (Frost and Ding, 2003). An obvious mass loss increase can
mica. This mass loss is only visible in CM and CMP10 because P addition be seen from CM to CMP50 at this period (Fig. 4a).
dilutes illite/mica concentration in the raw material so the In terms of thermodilatometry, it measures length changes

6
S. Wang et al. Applied Clay Science 216 (2022) 106384

Table 3
Quantitative phase analysis by the Rietveld refinement (wt%).
Brick button Qz Amp Mul Hem An Crs Ant Crd Spinel Mag

Fired CM 28.60 29.46 30.56 2.65 1.25 7.22 0.26 n.d. n.d. n.d.
Fired CMP10 26.45 35.10 27.21 2.53 1.75 6.25 0.35 0.36 n.d. n.d.
Fired CMP20 22.36 42.81 23.08 2.27 2.27 4.76 n.d. 0.78 1.67 n.d.
Fired CMP30 19.71 48.30 19.87 2.18 3.25 3.28 n.d. 1.25 2.16 n.d.
Fired CMP40 17.00 54.60 13.34 1.89 4.29 2.44 n.d. 2.39 3.97 0.08
Fired CMP50 14.85 59.13 8.19 1.80 6.01 2.21 n.d. 3.49 4.28 0.04

Note: Qz = quartz, Mul = mullite, Hem = hematite, An = anorthite, Crs = cristobalite, Ant = anatase, Crd = cordierite, Spl = spinel, Mag = magnetite, Amp =
amorphous, and n.d. = not detected. Mineral abbreviations after (Whitney and Evans, 2010).

Table 4
Chemical composition of measured points by EDS.
Element Element concentration of points measured by EDS (atom%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14

O 61.6 66.7 63.7 53.4 66.0 66.6 63.0 65.0 62.8 52.4 66.7 60.9 66.6 60.2
Si 16.0 33.3 21.1 3.9 3.3 33.0 18.4 26.1 16.9 40.8 33.3 15.6 0.3 18.3
Al 13.2 n.d. 13.5 5.7 2.7 0.2 16.3 7.9 17.3 4.2 n.d. 13.3 0.3 2.0
Fe 0.6 n.d. 0.3 37.0 0.3 0.2 0.4 0.6 1.6 2.6 n.d. 2.2 0.2 1.3
Ca 6.1 n.d. 0.2 n.d. n.d. n.d. 0.3 0.1 0.3 n.d. n.d. 0.4 n.d. 7.0
Mg 1.9 n.d. 0.2 n.d. n.d. n.d. 0.2 n.d. 0.4 n.d. n.d. 6.4 n.d. 10.4
K n.d. n.d. 0.8 n.d. 0.2 n.d. 0.8 0.1 0.5 n.d. n.d. 0.6 n.d. 0.2
Na n.d. n.d. n.d. n.d. n.d. n.d. 0.1 n.d. n.d. n.d. n.d. 0.4 n.d. 0.1
Ti 0.5 n.d. 0.1 n.d. 27.5 n.d. 0.3 0.1 0.2 n.d. n.d. 0.1 32.6 0.3
Cl n.d. n.d. n.d. n.d. n.d. n.d. 0.4 0.1 n.d. n.d. n.d. n.d. n.d. 0.2

Note: n.d. - not detected.

(expansion or shrinkage) of the green body on heating which are related β-quartz with the rising temperature (Antao, 2016).
to reactions/transformations such as dehydroxylation, quartz inversion, In “Stage 5” and “6”, all the clay mixtures show strong contractions
vitrification, etc. (Wang et al., 2021). For the CM sample, six stages of which increase with the P content. The contraction in “Stage 5” pre­
dimensional change are defined on the dilatometric curve of CM cor­ dominantly comes from crystallisation of mullite from metakaolin and
responding to the first derivative (dL/dt) data (Fig. 4d). collapse of palygorskite, illite/mica and mixed layer illite-smectite
In “Stage 1” (25–375 ◦ C, Fig. 4d), clay mixtures exhibit linear ex­ (Jimenez-Millan et al., 2018; McConville and Lee, 2005). The dL/dt
pansions with a dL/dt value of ~0.012 K− 1. No reaction happens in this peak decreases from 956 ◦ C (CM) to 889 ◦ C (CMP10), 881 ◦ C (CMP20),
temperature range, so a steady enlargement of the crystal lattice causes 869 ◦ C (CMP30), 865 ◦ C (CMP40) and 850 ◦ C (CMP50) with P addition,
this expansion. Although palygorskite and other clay minerals undergo which can be explained by the proportion of dehydroxylation of paly­
dehydration confirmed by the TG data (Fig. 4a), the corresponding gorskite that displays a peak at lower temperature (Fig. S6b). Besides,
contractions are relatively weak and thus, offset by the expansion the addition of P notably enhances the shrinkage in “Stage 5”. A 5.83%
process. length reduction is observed in sample CMP50, approximatively ten
The second length change locates in the range of ~375–540 ◦ C in times larger than that in CM (0.56%).
“Stage 2” (Fig. 4d). All clay mixtures display significant shrinkages with “Stage 6” starts at 1010 ◦ C and extends to 1150 ◦ C on the dL/dt curve
obtained values increasing from 0.02% (CM) to 0.03% (CMP10), 0.06% of the CM sample (Fig. 4d). Clay minerals begin to melt in this stage,
(CMP20), 0.08% (CMP30), 0.09% (CMP40) and 0.13% (CMP50) (see the accompanied by the crystallisation of some new phases. Mullite is
inset, Fig. 4c). According to the in-situ XRD data, this shrinkage is formed from dehydroxylated illite/mica and mixed layer illite-smectite,
contributed by three different transformations, including 1) dehydrox­ while cordierite precipitates from palygorskite. Wang et al. (2020) also
ylation of kaolinite, 2) decomposition of dolomite, and 3) dehydration of reported the melt of palygorskite nanofibers at 1100 ◦ C with the pro­
palygorskite. duction of a large amount of liquid phase. In addition, the corresponding
The third length change is a slight expansion between 540 and dL/dt peak gradually widens and becomes asymmetric from CM to
605 ◦ C (Stage 3) caused by the transition from α- to β-quartz (Vasic et al., CMP40. This change is due to an overlap of melt processes coming from
2021a). Kaolinite continues to dehydroxylate and causes shrinkage in different types of clay minerals. The dL/dt peak becomes symmetrical
this stage, as supported by in-situ XRD data. With P content increase, the again on CMP50 related to the domination of palygorskite liquefaction.
length increase of the clay mixture gradually slows down until it be­ Meanwhile, CMP50 exhibits an intensive shrinkage (9.75%) compared
comes unnoticeable in CMP50. This is because the addition of P dilutes to the other clay mixtures, concluding that the effect of P on the vitri­
the concentration of quartz in the starting material. fication process is a non-linear behaviour. Vitrification will be sharply
“Stage 4” represents a contraction occuring at 605–875 ◦ C. A few enhanced if the P concentration exceeds a specific value (i.e. 50 wt% in
reactions/changes are believed to contribute to this stage: 1) removal of this study).
the residual -OH units in dehydroxylated kaolinite (metakaolin), which
is accompanied by a slight shrinkage of clay mixture. Also, after dehy­
droxylation, Al atoms in metakaolin start to achieve the coordination of 3.4. Morphology and microstructure
4- or 6-fold and a short-range order material - the precursor of mullite is
formed (Gualtieri and Bellotto, 1998). The original crystal lattice of Fired CM exhibits an interconnected pore system as shown in Fig. 5a.
metakaolin collapses, then the clay shrinks; 2) dehydroxylation of illite/ These pores can be summarised in the following two formats: 1) pores
mica which is confirmed by in-situ XRD data. Slight structural expan­ alongside particle interfaces which indicate a low degree of sintering; 2)
sions occur due to the increased d002 values. 3) A gradual contraction of secondary pores in elongated or irregular shapes due to deformation
associated with clumps of clay minerals. In Fig. 5b, the partial melting of

7
S. Wang et al. Applied Clay Science 216 (2022) 106384

Fig. 5. BSE images of fired brick buttons: overall view of fired CM (a);partially molten clay minerals (b); anatase (Ant), hematite (Hem) and quartz/cristobalite (Qz/
Crs) in fired CM (c); overall view of fired CMP20 (d); overall view of fired CMP50 (e); residue of decomposed dolomite (f); Insets: higher magnification images of
molten clay minerals in (b), hematite (Hem) in (c) and cordierite (Crd) in (e).

clay minerals can be clearly seen. Some light-grey bands along the clay This phenomenon is even more evident in fired CMP50 when a contin­
minerals sheets (point 9, Fig. 5b) are richer in Al and Fe but poorer in Si uous and homogeneous matrix is formed (Fig. 5e). The particle bound­
than the grey matrix surroundings (Point 8, Fig. 5b). At higher magni­ aries and secondary pores related to clay material melting are almost
fication (bright coloured particles in Fig. 5c), hematite particles invisible due to the formation of a more glassy phase, which progres­
agglomerate closely in quartz/cristobalite grains, as indicated by EDS sively merges connected irregular pores to form closed oval ones. This
data (Table 4). These hematite particles are “short sausage-like” with observation is also supported by the quantitative XRD results showing
sizes less than 2 μm (inset, Fig. 5c). According to Nodari et al. (2007), that the amorphous content increases from 39.46% in fired CM to
they are inferred to be derived from on-site oxidation of Fe after Fe- 59.13% in CMP50, respectively. A higher degree of vitrification is
bearing clay minerals are melted (i.e., illite/mica or illite-smectite in obviously achieved in CMP50, attributed to increased flux contents
this case). In addition, the TiO2 phase accompanied with trace amounts including Fe, alkalis (Na and K) and alkaline-earth (Ca and Mg). The
of Si, Al, Fe, and K corresponds to anatase, supported by XRD results water released by clay minerals due to dehydroxylation may also facil­
(Table 4). itate melt formation (Rodriguez Navarro et al., 2003).
Adding 20 wt% of P has considerably changed the microstructure of Importantly, palygorskite fibres disappeared completely in the
brick buttons, resulting in obscurer interfaces between particles (white 1150 ◦ C fired brick buttons. This result is consistent with the findings
arrows, Fig. 5d) and reduced connected porosity than fired CM (Fig. 5a). from Boudriche et al. (2012) and Wang et al. (2020), which reported

8
S. Wang et al. Applied Clay Science 216 (2022) 106384

melting of palygorskite at 1000 ◦ C and 1100 ◦ C, respectively. Interest­ temperatures.


ingly, crystallised acicular cordierites are widely dispersed in the glassy
matrix of fired bricks (inset, Fig. 5e). These cordierites are randomly 3.5.2. Porosity
oriented with length (0–5 μm) and diameter (0–1 μm) larger than The porosities of fired brick buttons were investigated through a
original palygorskite fibres (length ≤ 1 μm and width of ~100 nm). combination of water absorption experiments (a routine practice used
Given the needle-like morphology, cordierite may assist in enhancing by brick manufacturers) and N2 isothermal adsorption and desorption
mechanical strength due to: 1) fibre reinforcement and 2) a higher measurement (Table 5). For brick manufacturers, the cold water
Young's modulus of 139 GPa in comparison with that of the glassy adsorption (CWA) test is usually used to estimate the number of pores in
matrix (68.8 GPa) (Ogiwara et al., 2010; Zheng et al., 2007). the brick that are filled by capillary action (approximately 10 μm or less
In addition, spot-like pores are densely packed together after in size), while the boiling water adsorption (BWA) test can evaluate
decomposition of dolomite and release of CO2 (Fig. 5f). This porous area smaller pores that are not accessible through capillary force (Nieminen
is rich in Mg and Ca as confirmed by the EDS mapping images in Fig. S3. and Romu, 1988; Wilson et al., 1999). For fired CM, CWA and BWA are
Some phases identified through EDS analysis include anorthite (point 1), 10.78% and 11.33%, respectively. These values are comparable to those
quartz (points 2 and 9), glassy matrix (point 3), hematite (points 4 and in the literature (Mezencevova et al., 2012; Ukwatta and Mohajerani,
8), anatase (points 5 and 11), molten metakaolin (point 7), cordierite 2017). With an increase of P content, both BWA and CWA decrease, and
(point 10), the residue of dolomite after decomposition (point 12), and this is consistent with the increased shrinkage on the dilatometric curves
quartz/cristobalite (point 6). when more P is introduced (Fig. 4c). In fired CMP50, the 24 h CWA falls
to only 0.94%, showing a very dense structure derived from intense
3.5. Physical, mechanical and thermal diffusivity properties vitrification and liquid-phase sintering. In addition, although the satu­
ration coefficient decreases from fired CM to fired CMP40 (Table 5),
3.5.1. Appearance which may result in more robust frost durability as saturation coefficient
Introduction of P significantly changes the colours of green bodies reflects the space available in the brick to accommodate the expansion
and fired buttons, while dried buttons have a barely noticeable differ­ pressure of freezing water (Netinger et al., 2014), reported that
ence in colour (Fig. 6). The green bodies exhibit lighter colours, and the
fired buttons become darker with increasing P content. A change from Table 5
“light camel” (fired CM) to “black olive” (fired CMP50) is presented, Cold and boiling water absorption and BET results for fired brick buttons.
which is presumably due to the formation of the cordierite phase. In
Fired 24 h cold 5 h boiling Saturation BET Average
addition, fired CMP40 and CMP50 display smooth surfaces with glazed- samples water water coefficient specific pore sizea
like texture, and the deformation of these two samples is more notice­ absorption absorption (CWA/ surface (nm)
able than others. This phenomenon agrees with Zhang and He (2014), (CWA, %) (BWA, %) BWA) area
which described that higher content of glass phases in ceramic brick (m2/g)
materials can cause visible distortion. In addition, some pinholes on the CM 10.78 11.33 0.95 2.44 4.96
surface of fired CMP50 are visible to the naked eyes, which may corre­ CMP10 9.33 10.62 0.88 2.72 4.89
CMP20 6.98 10.28 0.68 2.78 5.12
spond to the oval-pores observed in SEM images. Wang et al. (2022) also
CMP30 6.32 9.97 0.63 2.25 5.12
reported spherical and oval-shaped pores in the vermiculite-added fired CMP40 3.65 7.50 0.49 2.55 5.15
brick buttons. After the P concentration reaches more than 40%, a few CMP50 0.94 1.07 0.88 2.83 4.74
small cracks are observed on the surface of fired brick buttons, especially a
Based on BET specific surface area analysis, limited to pores between 1.7 and
for CMP50 (white arrows in Fig. 6). These cracks are speculated to be
300 nm.
due to excessive and uneven shrinkage of the brick body at high

Fig. 6. The appearance of brick buttons at different fabrication stages. In the left image, top to bottom pictures: green bodies, dried brick buttons, and fired brick
buttons. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

9
S. Wang et al. Applied Clay Science 216 (2022) 106384

saturation coefficient cannot predict the durability of brick reliably then increases with P content. The density can be calculated as ρ = m/V,
because it is based on an incomplete understanding of the physics of where m stands for mass and V for volume. From CM to CMP50, the
freeze-thaw damage, an oversimplification of field exposure condition, obtained brick mass decreases linearly (Fig. 4a). However, the volume
and testing that focus on unit (not material) response to freeze-thaw shrinkage from CM to CMP50 is not linear (Fig. 7b). Consequently,
cycling. Also, for all samples, BET specific surface area and the although both 10 wt% of P is added to clay mixtures, a larger volume
average pore size show very close values at 2.25–2.83 g/cm3 and shrinkage (− 8.58%) is observed from CMP40 to CMP50 than that
4.74–5.15 nm, respectively, indicating that micropores (< 2 nm) and (− 3.87%) from CM to CMP10. It seems that with the increase of P
mesopores (2 nm to 50 nm) in fired bricks are consistent. content in the clay mixture, further addition of P can cause considerable
volume shrinkage. This is why the density first decreases from CM to
3.5.3. Shrinkage and bulk density CMV30 (when the change of brick mass is faster than that of the volume)
The evolution of drying and firing shrinkage is quite complicated and then increases from CM30 to CMP50 (when the change of volume is
after P has been added to the clay mixture (Fig. 7a). Regarding the faster than that of the mass).
drying shrinkage, comparable results from 0.1312% (CM) to 0.1318%
(CMP10) are observed, followed by a rapid rise to 0.1780% for the 3.5.4. Compressive strength
CMP30 sample. Then, drying shrinkage remains nearly stable from The compressive strength of obtained brick buttons improves with P
CMP30 to CMP50, and a possible reason is that P has introduced more content (Fig. 7c). This is caused by reduced porosity of brick buttons and
clay minerals in the green body, thus a higher plastic limit value is ob­ fibre reinforcement by acicular cordierite particles in the glassy matrix.
tained. In CMP40 and CMP50, 4 mL of water may not be enough to wet Similar to firing shrinkage, compressive strength grows steadily initially,
and connect all particles, as more clay minerals generally need more then followed by a more substantial rise when the concentration of P
water. This plastic limit may explain why a further drying shrinkage, increases from 30% to 40%. However, although an enhanced firing
caused by shortening of the particle distance, is not achieved when P shrinkage occurs from CMP40 to CMP50 (Fig. 7a) with a significant
content rises from 30 wt% to 50 wt%. decrease in porosity (5 h boiling water absorption drops rapidly,
In terms of the firing shrinkage, it grows steadily from CM to CMP30 Table 5), the obtained compressive strength does not change too much
prior to a sharper increase afterwards. A similar trend is also seen for (a slight increase from 200.2 Mpa to 210.2 Mpa, Fig. 7c). There are two
volume shrinkage in Fig. 7b. The volume of the clay button shrinks by possible reasons: 1) Cracks on the CMP50 surface may serve as starting
14.61% from CMP30 to CMP50 while a volumetric reduction of only points of failure during compression measurement. These cracks cannot
6.21% from CM to CMP20 is presented. This demonstrates a non-linear be detected by water absorption experiment because water tends to flow
and non-uniform effect of P on firing shrinkage, as also supported by the down in them due to gravity; 2) The mineral/phase coordination in fired
dilatometric data (Fig. 4c). According to (Weng et al., 2003), a good CMP50 may not be ideal. Since quartz shows higher hardness and
quality brick exhibits a firing shrinkage below 8%. However, the firing modulus than the amorphous phase (Krakowiak et al., 2011), high
shrinkages of CMP40 and CMP50 climb to 8.71% and 11.48%, respec­ compressive strength can be achieved when quartz particles are welded
tively, which is attributed to the enhanced liquid-phase sintering in the relatively weak amorphous phase. However, the higher amor­
(Fig. 4c). phous to quartz ratio in the fired CMP50 make this structure unable to be
Interestingly, the density of fired brick buttons decreases first and achieved, consequently, the brick strength has been slightly improved

Fig. 7. Physical and mechanical properties of green bodies and brick buttons: drying and firing shrinkages (a), density and volume shrinkage (b), compressive
strength (c) and thermal diffusivity (d). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

10
S. Wang et al. Applied Clay Science 216 (2022) 106384

although with a significant drop of the brick porosity. colour change from “light camel” (fired CM) to “black olive”(fired
CMP50) is observed. When the P content is greater than 40 wt%,
3.5.5. Thermal diffusivity brick buttons exhibit a glazed texture, with tiny pores on the surface
P addition in CM increases the thermal diffusivity of obtained brick and more severe deformation caused by a higher degree of
buttons. This enhancement becomes more evident when the P content vitrification.
exceeds 20% (Fig. 7d). At 25 ◦ C, the thermal diffusivity of CM, CMP10, 5) CMP40 and CMP50 have significantly increased bulk density,
CMP20, CMP30, CMP40 and CMP50 are 0.742 mm2/s, 0.753 mm2/s, compressive strength and thermal diffusivity. They also have huge
0.756 mm2/s, 0.788 mm2/s, 0.833 mm2/s and 0.853 mm2/s, respec­ volume shrinkages at 27.0% and 35.0%, respectively. Higher
tively. The enhanced thermal diffusivity is believed to be caused by a compressive strength of fired buttons with P results from reduced
denser structure since thermal diffusivity likely reflects the overall porosity and fibre reinforcement by acicular cordierite particles in
porosity of the materials. the glassy matrix. Small cracks appear on the brick button surface,
In short, the higher firing temperature used in this study has resulted especially for fired CMP50, so some products with higher P are
in improved compressive strength than that of other bricks generally unqualified.
described in the literature. According to (ASTMC62, 2017), CMP10 and
CMP50 can be classified as the Grade MW brick (brick intended for use Therefore, addition of 30 wt% P in the clay mixture is considered the
in moderate weathering areas), requiring a compressive strength larger optimal ratio for brickmaking. Fired CMP30 can be classified as Grade
than 17.2 MPa with water absorption and saturation coefficient less than SW brick with compressive strength and thermal diffusivity (25 ◦ C) of
22% and 0.88, respectively, while CMP20, CMP30 and CMP40 belong to 153.8 MPa (up 24.5% over the control sample) and 0.788 mm2/s,
the Grade SW brick (brick intended for use in a severe weathering area), respectively. In addition, the costs of producing bricks at higher tem­
which requires a compressive strength larger than 20.7 MPa with water perature of 1150 ◦ C may not be the most energy efficient for some brick
absorption and saturation coefficient less than 17% and 0.78, manufacturers. But as already mentioned, this firing temperature is
respectively. essential for Queensland brick manufacturers.

4. Conclusions Credit authorship contribution statement

This research has quantitatively investigated the evolution of in-situ Sen Wang: Methodology, Investigation, Formal analysis, Writing –
thermal behaviours as well as physical, mechanical and thermal diffu­ original draft. Lloyd Gainey: Methodology, Resources, Supervision,
sivity properties of brick buttons when palygorskite containing clay (P) Writing – review & editing. Xiaodong Wang: Formal analysis, Valida­
is added to the starting material. Although palygorskite is relatively tion. Ian D.R. Mackinnon: Methodology, Supervision, Writing – review
rarely distributed in nature compared to other clay minerals like & editing. Yunfei Xi: Conceptualization, Resources, Visualization,
kaolinite, illite and montmorillonite, it can significantly affect the Writing – review & editing, Supervision, Funding acquisition.
manufacturing process and performance of fired clay bricks as follows:
Declaration of Competing Interest
1) Palygorskite can notably change the mineralogical compositions of
obtained brick buttons. The amorphous fraction increases from The authors declare that they have no known competing financial
29.46% to 59.13% after 50 wt% of P has been added. Mullite drops interests or personal relationships that could have appeared to influence
from 30.56% to 8.19%, while two Mg-rich phases - cordierite and the work reported in this paper.
spinel appear with total contents less than 10% (in fired CMP50).
Although the addition of P increases the concentration of Fe in the Acknowledgments
starting material, the presence of hematite after 1150 ◦ C is sup­
pressed through locking of Fe into cordierite, spinel and magnetite An Advance Queensland Research Fellowship with Brickworks Ltd.
phases. and the Centre for Clean Energy Technologies and Practices have funded
2) P promotes the total mass loss and shrinkage of brick bodies during this work (Grant No. AQRF01716-17RD2). The authors would like to
the firing process. The increase in mass loss is caused by the dehy­ acknowledge Mr. Mike Newitt, Ms. Brianna Deer, Mr. Julius Marinelli,
droxylation of palygorskite and decomposition of dolomite between Mr. Mick Bolton and Mr. Calum Henderson from Brickworks Ltd., Mr.
500 and 800 ◦ C, while the rise in dimensional shrinkage primarily Donald McAuley, Ms. Elizabeth Graham and Miss Karine Harumi
comes from collapse and melt of palygorskite nanofibers at Moromizato from the Central Analytical Research Facility (CARF) for
~800–1150 ◦ C. Based on the first derivative curves of TGA and the assistance with sample characterisations.
dilatometry, five mass loss steps and six thermal expansion/
contraction stages are defined between 25 and 1150 ◦ C. Appendix A. Supplementary data
3) The addition of P results in obscurer interfaces of particles and
reduced connective porosity in fired brick buttons. A continuous and Supplementary data to this article can be found online at https://doi.
homogeneous matrix is formed in fired CMP50 with big oval pores org/10.1016/j.clay.2021.106384.
inside. The original fibrous palygorskite disappears after 1150 ◦ C
firing with larger acicular cordierite particles (length ≤ 5 μm and References
diameter ≤ 1 μm) widely developed. These cordierite particles may
bring enhanced mechanical properties due to fibre reinforcement Antao, S.M., 2016. Quartz: structural and thermodynamic analyses across the alpha <− >
beta transition with origin of negative thermal expansion (NTE) in beta quartz and
and a high Young's modulus. calcite. Acta Crystallogr. Sect. B: Struct. Sci. Cryst. Eng. Mater. 72, 249–262.
4) The drying shrinkage of the clay mixture remains stable at ~0.178% ASTMC62, 2017. Standard Specification for Building Brick (Solid Masonry Units Made
when P content reaches more than 30 wt%, probably caused by the from Clay or Shale). ASTM International, West Conshohocken, PA, USA.
ASTMC67/C67M, 2020. Standard Test Methods for Sampling and Testing Brick and
increased plasticity limit of the clay mixture. P can promote firing
Structural Clay Tile. ASTM International, West Conshohocken, PA, USA.
shrinkage, but its effect is non-linear and non-uniform. When the P Awadh, S.M., Abdullah, H.H., 2011. Mineralogical, geochemical, and geotechnical
content exceeds 30 wt%, the firing shrinkage significantly increases evaluation of Al-Sowera soil for the building brick industry in Iraq. Arab. J. Geosci. 4
due to intense vitrification and liquid-phase sintering. In addition, (3–4), 413–419.
Boudriche, L., Calvet, R., Hamdi, B., Balard, H., 2012. Surface properties evolution of
the colour of the brick button becomes darker as more palygorskite is attapulgite by IGC analysis as a function of thermal treatment. Colloids. Surf. A
added. Due to the formation of cordierite in the fired product, a Physicochem. Eng. 399, 1–10.

11
S. Wang et al. Applied Clay Science 216 (2022) 106384

Chalouati, Y., Bennour, A., Mannai, F., Srasra, E., 2020. Characterization, thermal Ogiwara, T., Noda, Y., Shoji, K., Kimura, O., 2010. Solid state synthesis and its
behaviour and firing properties of clay materials from Cap Bon Basin, north-East characterization of high density cordierite ceramics using fine oxide powders.
Tunisia, for ceramic applications. Clay Miner. 55 (4), 351–365. J. Ceram. Soc. Jpn. 118 (1375), 246–249.
Chen, T.H., Xu, H.F., Lu, A.H., Xu, X.C., Peng, S.C., Yue, S.C., 2004. Direct evidence of Post, J.E., Heaney, P.J., 2008. Synchrotron powder X-ray diffraction study of the
transformation from smectite to palygorskite: TEM investigation. Sci. China Earth structure and dehydration behavior of palygorskite. Am. Mineral. 93 (4), 667–675.
Sci. 47 (11), 985–994. Rehman, M.U., Ahmad, M., Rashid, K., 2020. Influence of fluxing oxides from waste on
Chen, T.H., Wang, J., Qing, C.S., Peng, S.C., Song, Y.X., Guo, Y., 2006. Effect of heat the production and physico-mechanical properties of fired clay brick: a review.
treatment on structure, morphology and surface properties of palygorskite. J. Chin. J. Build. Eng. 27, 100965.
Ceram. Soc. 34 (11), 1406–1410. Reinhardt, J., 1987. Cordierite-anthophyllite rocks from north-west Queensland,
Cheng, H., Yang, J., Frost, R.L., 2011. Thermogravimetric analysis-mass spectrometry Australia: metamorphosed magnesian pelites. J. Metamorph. Geol. 5 (4), 451–472.
(TG-MS) of selected Chinese palygorskites—implications for structural water. Rodriguez-Navarro, C., Cultrone, G., Sanchez-Navas, A., Sebastian, E., 2003. TEM study
Thermochim. Acta 512 (1–2), 202–207. of mullite growth after muscovite breakdown. Am. Mineral. 88 (5-6), 713–724.
Cheng, H.F., Zhang, S., Liu, Q.F., Li, X.G., Frost, R.L., 2015. The molecular structure of Tang, Q.G., Wang, F., Liu, X.D., Tang, M.R., Zeng, Z.G., Liang, J.S., Guan, X.Y., Wang, J.,
kaolinite-potassium acetate intercalation complexes: a combined experimental and Mu, X.Z., 2016. Surface modified palygorskite nanofibers and their applications as
molecular dynamic simulation study. Appl. Clay Sci. 116, 273–280. reinforcement phase in cis-polybutadiene rubber nanocomposites. Appl. Clay Sci.
Cultrone, G., Sebastian, E., 2009. Fly ash addition in clayey materials to improve the 132, 175–181.
quality of solid bricks. Constr. Build. Mater. 23 (2), 1178–1184. Tantawy, M.A., Mohamed, R.S.A., 2017. Middle Eocene clay from Goset Abu Khashier:
Cultrone, G., Sebastian, E., Elert, K., de la Torre, M.J., Cazalla, O., Rodriguez-Navarro, C., Geological assessment and utilization with drinking water treatment sludge in brick
2004. Influence of mineralogy and firing temperature on the porosity of bricks. manufacture. Appl. Clay Sci. 138, 114–124.
J. Eur. Ceram. Soc. 24 (3), 547–564. Ukwatta, A., Mohajerani, A., 2017. Leachate analysis of green and fired-clay bricks
Ding, J., Hui, A., Wang, W., Yang, F., Kang, Y., Wang, A., 2021. Multifunctional incorporated with biosolids. Waste Manag. 66, 134–144.
palygorskite@ ZnO nanorods enhance simultaneously mechanical strength and Vasic, M.V., Pezo, L., Zdravkovic, J.D., Backalic, Z., Radojevic, Z., 2017. The study of
antibacterial properties of chitosan-based film. Int. J. Biol. Macromol. 189, 668–677. thermal behavior of montmorillonite and hydromica brick clays in predicting tunnel
El Ouahabi, M., Daoudi, L., Hatert, F., Fagel, N., 2015. Modified mineral phases during kiln firing curve. Constr. Build. Mater. 150, 872–879.
clay ceramic firing. Clay Clay Miner. 63 (5), 404–413. Vasic, M.V., Goel, G., Vasic, M., Radojevic, Z., 2021a. Recycling of waste coal dust for the
Elavarasan, S., Priya, A.K., Kumar, V.K., 2021. Manufacturing fired clay brick using fly energy-efficient fabrication of bricks: a laboratory to industrial-scale study. Environ.
ash and M-sand. Mater. Today Proc. 37, 872–876. Technol. Innov. 21, 101350.
Ezzatahmadi, N., Ayoko, G.A., Millar, G.J., Speight, R., Yan, C., Li, J., Li, S., Zhu, J., Vasic, M.V., Pezo, L., Gupta, V., Chaudhary, S., Radojevic, Z., 2021b. An artificial neural
Xi, Y., 2017. Clay-supported nanoscale zero-valent iron composite materials for the network-based prediction model for utilization of coal ash in production of fired clay
remediation of contaminated aqueous solutions: a review. Chem. Eng. J. 312, bricks: a review. Sci. Sinter. 53 (1), 37–53.
336–350. Viani, A., Cultrone, G., Sotiriadis, K., Sevcik, R., Sasek, P., 2018. The use of mineralogical
Frost, R.L., Ding, Z., 2003. Controlled rate thermal analysis and differential scanning indicators for the assessment of firing temperature in fired-clay bodies. Appl. Clay
calorimetry of sepiolites and palygorskites. Thermochim. Acta 397 (1–2), 119–128. Sci. 163, 108–118.
Galan, E., 2006. Genesis of clay minerals. In: Bergaya, F., Theng, B.K.G., Lagaly, G. Wang, K.L., Wang, L.J., Zhang, Y., Zhang, Y.D., Liang, J.S., 2020. Microstructural
(Eds.), Handbook of Clay Science. Elsevier, Newnes, pp. 1129–1162. evolution and sintering properties of palygorskite nanofibers. Int. J. Appl. Ceram.
Garcia, N., Hoyos, M., Guzman, J., Tiemblo, P., 2009. Comparing the effect of nanofillers Technol. 17 (4), 1833–1842.
as thermal stabilizers in low density polyethylene. Polym. Degrad. Stab. 94 (1), Wang, S., Gainey, L., Baxter, D., Wang, X.D., Mackinnon, D.R.I., Xi, Y.F., 2021. Thermal
39–48. behaviours of clay mixtures during brick firing: a combined study of in-situ XRD,
Gonzalez, F., Pesquera, C., Benito, I., Mendioroz, S., Pajares, J., 1990. Isomorphlc TGA and thermal dilatometry. Constr. Build. Mater. 299, 124319.
substitution in attapulgite and its influence on the dehydration process. Wang, S., Gainey, L., Marinelli, J., Deer, B., Wang, X.D., Mackinnon, D.R.I., Xi, Y.F.,
Thermochim. Acta 167 (1), 139–143. 2022. Effects of vermiculite on in-situ thermal behaviour, microstructure, physical
Gualtieri, A., Bellotto, M., 1998. Modelling the structure of the metastable phases in the and mechanical properties of fired clay bricks. Constr. Build. Mater. 316, 125828.
reaction sequence kaolinite-mullite by X-ray scattering experiments. Phys. Chem. Weng, C.H., Lin, D.F., Chiang, P.C., 2003. Utilization of sludge as brick materials. Adv.
Miner. 25 (6), 442–452. Environ. Res. 7 (3), 679–685.
Hirsiger, W., Mullervonmoos, M., Wiedemann, H.G., 1975. Thermal-analysis of Whitney, D.L., Evans, B.W., 2010. Abbreviations for names of rock-forming minerals.
palygorskite. Thermochim. Acta 13 (2), 223–230. Am. Mineral. 95 (1), 185–187.
Ji, J., Zhou, S.Y., Lai, C.Y., Wang, B., Li, K., 2015. PVDF/palygorskite composite Wilson, M., Carter, M., Hoff, W., 1999. British Standard and RILEM water absorption
ultrafiltration membranes with enhanced abrasion resistance and flux. J. Membr. tests: a critical evaluation. Mater. Struct. 32 (8), 571–578.
Sci. 495, 91–100. Wu, X.P., Zhang, L., Wang, C., Liu, Y., Dai, J.J., Zhu, D., Wu, Y.C., Zhang, X.L., 2015.
Jimenez-Millan, J., Abad, I., Jimenez-Espinosa, R., Yebra-Rodriguez, A., 2018. Structural evolution and property of hot-pressed palygorskite clay. Trans. Indian
Assessment of solar panel waste glass in the manufacture of sepiolite based clay Ceram. Soc. 74 (3), 169–176.
bricks. Mater. Lett. 218, 346–348. Xi, Y., Sun, Z., Hreid, T., Ayoko, G.A., Frost, R.L., 2014. Bisphenol a degradation
Johari, I., Said, S., Hisham, B., Bakar, A., Ahmad, Z.A., 2010. Effect of the change of enhanced by air bubbles via advanced oxidation using in situ generated ferrous ions
firing temperature on microstructure and physical properties of clay bricks from from nano zero-valent iron/palygorskite composite materials. Chem. Eng. J. 247,
Beruas (Malaysia). Sci. Sinter. 42 (2), 245–254. 66–74.
Krakowiak, K.J., Lourenco, P.B., Ulm, F.J., 2011. Multitechnique investigation of Zhang, H.Y., He, H.H., 2014. Characterization of microstructure of FA ceramic brick
extruded clay brick microstructure. J. Am. Ceram. Soc. 94 (9), 3012–3022. using SEM. Mater. Sci. Technol. 849, 283–286.
Kulbicki, G., 1959. High temperature phases in sepiolite, attapulgite and saponite. Am. Zhang, Y.H., Yu, C.X., Hu, P., Tong, W.S., Lv, F.Z., Chu, P.K., Wang, H.L., 2016.
Mineral. 44 (7–8), 752–764. Mechanical and thermal properties of palygorskite poly(butylene succinate)
Liu, H.B., Chen, T.H., Chang, D.Y., Qing, C.S., Kong, D.J., Chen, D., Xie, J.J., Frost, R.L., nanocomposite. Appl. Clay Sci. 119, 96–102.
2013. Effect of rehydration on structure and surface properties of thermally treated Zhang, P.L., Tian, N., Zhang, J.P., Wang, A.Q., 2018a. Effects of modification of
palygorskite. J. Colloid Interface Sci. 393, 87–91. palygorskite on superamphiphobicity and microstructure of palygorskite@
Loutou, M., Taha, Y., Benzaazoua, M., Daafi, Y., Hakkou, R., 2019. Valorization of clay fluorinated polysiloxane superamphiphobic coatings. Appl. Clay Sci. 160, 144–152.
by-product from moroccan phosphate mines for the production of fired bricks. Zhang, Z.F., Wang, W.B., Tian, G.Y., Wang, Q., Wang, A.Q., 2018b. Solvothermal
J. Clean. Prod. 229, 169–179. evolution of red palygorskite in dimethyl sulfoxide/water. Appl. Clay Sci. 159,
Maithel, S., Ravi, A., Kuma, S., 2016. Market Assessment for Burnt Clay Resource 16–24.
Efficient Bricks (REBs). Greentech Knowledge Solutions Pvt. Ltd., New Delhi, India. Zhang, T., Li, Z.Q., Wang, W.B., Wang, Y., Gao, B.Y., Wang, Z.N., 2019a. Enhanced
McConville, C.J., Lee, W.E., 2005. Microstructural development on firing illite and antifouling and antimicrobial thin film nanocomposite membranes with
smectite clays compared with that in kaolinite. J. Am. Ceram. Soc. 88 (8), incorporation of Palygorskite/titanium dioxide hybrid material. J. Colloid Interface
2267–2276. Sci. 537, 1–10.
Mezencevova, A., Yeboah, N.N., Burns, S.E., Kahn, L.F., Kurtis, K.E., 2012. Utilization of Zhang, Y.D., Wang, L.J., Wang, F., Sun, J.F., Tang, Q.G., Liang, J.S., 2019b. Effects of
Savannah Harbor river sediment as the primary raw material in production of fired palygorskite on physical properties and mechanical performances of bone China.
brick. J. Environ. Manag. 113, 128–136. Appl. Clay Sci. 168, 287–294.
Mumpton, F.A., Roy, R., 1958. New data on sepiolite and attapulgite. Clay Clay Miner. 5, Zheng, L.Q., Schmid, A.W., Lambropoulos, J.C., 2007. Surface effects on Young's
136–143. modulus and hardness of fused silica by nanoindentation study. J. Mater. Sci. 42 (1),
Netinger, I., Vracevic, M., Ranogajec, J., Vucetic, S., 2014. Evaluation of brick resistance 191–198.
to freeze/thaw cycles according to indirect procedures. Gradevinar 66 (3), 197–209. Zhou, J.N., Zhang, X.Z., Wang, Y.Q., Larbot, A., Hu, X.B., 2010. Elaboration and
Nieminen, P., Romu, M., 1988. Porosity and frost-resistance of clay bricks. Bri. Blo. characterization of tubular macroporous ceramic support for membranes from
Mason. 1-3, 103–109. kaolin and dolomite. J. Porous. Mater. 17 (1), 1–9.
Nodari, L., Marcuz, E., Maritan, L., Mazzoli, C., Russo, U., 2007. Hematite nucleation and Zuo, X.C., Wang, D., Zhang, S.L., Liu, Q.F., Yang, H.M., 2018. Intercalation and
growth in the firing of carbonate-rich clay for pottery production. J. Eur. Ceram. Soc. exfoliation of kaolinite with sodium dodecyl sulfate. Minerals 8 (3), 112.
27 (16), 4665–4673.

12

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy