Tia Lite

You might also like

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

Materials Characterization 173 (2021) 110764

Contents lists available at ScienceDirect

Materials Characterization
journal homepage: www.elsevier.com/locate/matchar

Effect of nanostructuring on thermal stability and decomposition of


aluminium titanate (Al2TiO5): A phase transformation study
Abolfazl Azarniya a, *, Hamid Reza Madaah Hosseini a, *, Chinappan Amutha b,
Seeram Ramakrishna b
a
Department of Materials Science and Engineering, Sharif University of Technology, P.O. Box 11155-9466, Azadi Avenue, Tehran, Iran
b
Department of Mechanical Engineering, National University of Singapore, 9 Engineering Drive 1, Singapore 117576, Singapore

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

Keywords: As a refractory ceramic, the thermal instability of aluminium titanate (AT or Al2TiO5) has been in the crosshairs
Aluminium titanate of research works in recent years. The studies have indicated that different parameters such as atmosphere, grain
Al2TiO5 size, additives, oxygen pressure, and synthesis method can bilaterally affect the thermal decomposition of AT
Nanostructuring
ceramic and resultant physicomechanical properties. In the present study, the AT nanostructure was synthesized
Thermal decomposition
by a citrate sol-gel method and influence of nanostructuring nature on its thermal instability and reaction
pathway was explored in some details. It was shown that the reduction of grain size down to 30 nm can
destabilize AT, so that the reaction starts from the threshold temperature of around 300 ◦ C instead of 900 ◦ C. This
significant reduction in reaction temperature was attributed to the increase of interfaces density and crystallo­
graphic imperfections, serving as preferential sites for starting the thermal dissociation. However, the nano­
structuring is not able to change the temperature spectrum in which a vast majority of AT ceramic is dissociated
with a strong kinetics (i.e. 1100 ◦ C). It was confirmed that the thermal instability of AT nanostructure bears a
slow kinetics, so that a heat treatment at 1200 ◦ C for 5 h cannot completely decompose this ceramic.

1. Introduction porous point-of-use water-purification filters [15], photocatalyst for the


decomposition of organic pollutants [4,16], and many high-temperature
Aluminium titanate (AT or Al2TiO5) is a synthetic refractory ceramic applications as a desired thermostable fireproof material where low
with desirable properties for a broad spectrum of metallurgical and in­ thermal conductivity and high thermal shock are seriously needed [2,8].
dustrial applications. It bears an orthorhombic crystallographic system Table 1 provides a brief list of the main characteristics reported in the
in which every aluminium or titanium is neighbored by six oxygen literature for AT, SiC and Al2O3 to enable a comparative examination
atoms, so that AlO6 and TiO6 octahedra are formed [1]. Among the from mechanical, thermal and physical perspectives. As shown, the
remarkable properties of this ceramic, the following cases can be thermal conductivity, fracture toughness hardness, and bending
mentioned: (a) relatively low thermal expansion coefficient (1 × strength of AT are generally lower than those of SiC and Al2O3, but its
10− 6K− 1) [2], (b) high melting temperature (1860 ◦ C) [3], (c) and low thermal expansion coefficient is smaller than Al2O3 and comparative to
thermal conductivity (0.9–1.5 W⋅m− 1⋅K− 1) [3–5], (d) exceedingly high SiC.
thermal shock resistance (up to 500 Wm− 1) [6], (e) excellent corrosion A wide variety of fabrication technologies is commonly used to
resistance [7], and (f) non-infiltration to metals and molten glasses [8]. synthesize AT phase in both particle and bulk or thin film forms, among
These characteristics have rendered AT a good candidate for catalyst which the reaction sintering of mechanically alloyed Al2O3-TiO2 blend
supports, particle filters [3,9], thermal barrier coatings and exhaust [64] and chemical bottom-up synthesis approaches [65]. In first tech­
filter components for diesel engines [10], soot traps in diesel engines and nique, the powder blend is homogeneously mixed through a high-energy
filters for hot gas clean-up applications [11], chemical source for the ball milling and heat treats at a sufficiently high temperature for trig­
synthesis of hybrid in-situ aluminium matrix composites [2,12–14], gering the chemical reaction between Al2O3 and TiO2. The

* Corresponding authors.
E-mail addresses: Abolfazl.azarniya@u.nus.edu (A. Azarniya), madaah@sharif.edu (H.R.M. Hosseini), mpecam@nus.edu.sg (C. Amutha), seeram@nus.edu.sg
(S. Ramakrishna).

https://doi.org/10.1016/j.matchar.2020.110764
Received 17 April 2020; Received in revised form 14 October 2020; Accepted 12 November 2020
Available online 16 November 2020
1044-5803/© 2020 Published by Elsevier Inc.
A. Azarniya et al. Materials Characterization 173 (2021) 110764

Table 1 erential sites for the AT nucleation [73]. This mechanism is controlled
A literature review on some main properties of AT, SiC, and Al2O3. by the diffusion of Al atoms at high temperatures and depends on the
Property Al2TiO5 SiC Al2O3 starting precursors and applied heat treatment [74]. One can model the
AT decomposition process through the following Eq. [13]:
Thermal 2 [17] 150–325 [18] 30 [19]
conductivity 140 [20] In porous Ln (1 − α) = − kt− n
(1)
(W/m.k) form: 0.22
[21]
72–92 [22] 12–33 [23]
where k and n are temperature-dependent constants, t is the decompo­
55 [24] 3.76–12.32 sition time, and α is the reaction degree. This model can be manipulated
[25] by the grain size, applied atmosphere, sintering procedure, and additives
270 [26,27] 25.9–32.8 [28] [13]. It has been shown that the thermal degradation rate is accelerated
In porous form: 0.68–1.03 15–40 [30]
as the grain size is declined [75] or a reducing atmosphere is applied
[29]
Phase 950–1280 - order-disorder γ-Al2O3 to [76].
transformation [17] transformation under SPS: δ-Al2O3 at Another influencing factor is the nanoscale grain size. To the best of
temperature at 2200 ◦ C under 40 MPa 750 ◦ C our knowledge, no effort has been made to examine the influence of
(◦ C) - decomposition δ-Al2O3 to nanostructuring on the thermal decomposition of AT ceramic and its
temperature: between 2700 θ-Al2O3 at
and 2800 ◦ C at atmospheric 900 ◦ C
relationship with microstructure evolution. Recently, Azarniya et al.
pressure [31] θ-Al2O3 to [65] have synthesized an AT-based nanostructure with the average size
α-Al2O3 at of 50 nm by a citrate sol-gel method and showed that the thermal
1200 ◦ C [32] instability of this material starts approximately at 250 ◦ C in a single
1280 1400 [35] γ-Al2O3 to
kinetic step, so that the maximum possible dissociation occurs at 650 ◦ C.
[33,34] α-Al2O3 at
1250 ◦ C [36] This deviation from the thermodynamic predictions has been attributed
Thermal 0− 3× 3.5×10− 6
[37] 14×10− 6 [38] to the increased content of grain boundaries and population of crystal­
expansion 10− 6 [17] line defects, which act as preferred sites for the formation of Al2O3 and
coefficient (K− 1) 4.1–8.3 3.3 ×10 − 6
[40] 8.3×10− 6
[41] TiO2 phases [2]. To fill this gap, the present study strives to shed light on
×10− 6 [39]
1.5 ×10− 6 3.2 ×10− 6
[43]
the thermal dissociation of AT ceramic from the phase transformation
[42] and microstructure points of view. Additionally, the influence of nano­
Tensile strength 106 [44] 2.8 [45] 350 [46,47] structuring on the thermal instability of this ceramic will be discussed.
(MPa) 50–250 3.08 [35] 260–410 [30] Another purpose is to study on how the grain morphology and geometry
[39]
may change during the thermal decomposition.
Young Modulus 5–30 [17] 260 [45] 3.4 [48]
(GPa) 43–202 398.5, 385.3 [49] 57–141 [50]
[39] 2. Experimental procedure
8 [51]
14 [42] 2.1. Synthesis and characterization of AT-based nanostructure
Bending strength 10–50 [17] 395 [52] 290 [53]
(MPa) 61–230 280, 390 [54] 860 [55]
[39] 300 [56] The as-received nanostructured AT-based particles were synthesized
Fracture 0.45 [51] 4 [57] 4.1 [58] by a citrate sol-gel method under a protective humidity-free atmosphere.
toughness (KIC) 3.2 [59] 4.5 [52] 3.25 [46,47] The starting materials included Titanium Tetraisopropoxide (TTIP or
2.1–2.7 [60]
C16H36O4Ti), absolute ethanol, anhydrous aluminium chloride (AlCl3),
3.2–5.2 [61]
Flexural strength 9, 30 [62] 5.6 [57] 578 [58] and citric acid (C6H8O7). While TTIP and AlCl3 were chosen as the
(MPa) 15 [63] chemical source for entering Ti and Al ions to the solution, citric acid
25 [42] was used as the chelating agent persuading the nucleation of uncalcined
Vickers hardness 10.7 [59] 28.5 [52] 20–22 [60] AT clusters. All materials were supplied by Merk Company.
(GPa) 1.4 [63] 16–21 [61]
Fig. 1 shows a schematic flow chart of the used synthesis procedure.
As shown, TTIP and AlCl3 were added to the ethanol in the stoichio­
microstructure in this route is generally coarse due to exceedingly high- metric molar ratio. In this synthesis procedure, 2 g of anhydrous AlCl3
temperature heat treatments (e.g. 1400–1500 ◦ C) [64,66]. In contrast, and 2.6 g of TTIP were dissolved in 40 ml absolute ethanol and stirred
the solution-based synthesis methods such as the sol-gel method [67] for 30 min under a controllable flow of nitrogen gas at ambient tem­
and hydrothermal synthesis [68] enable producing well-engineered fine perature (solution 1). Then, 1.5 g of anhydrous citric acid was gradually
particles whose size distribution and morphology can be tailored by added to the solution 1 and stirred for 10 min to be completely dissolved
modulating the processing parameters. (solution 2). The chamber was then heated up to 80 min for 45 min to
Overall, there are two main drawbacks which restrict the industrial form a transparent and homogenous honey-like gel. Finally, the gel was
applications of AT-based ceramics: (a) the thermal decomposition of AT- dried at 100 ◦ C in an oven overnight to yield highly porous yellow
based ceramics to its parent phases i.e. rutile and corundum in the clusters. The dried gel was calcined at 900 ◦ C for 2 h to form white
temperature range of 800–1280 ◦ C; and (b) low mechanical strength and crystalline particles. The calcination parameters have been suggested by
high sensitivity to microcrack formation due to their intrinsic tendency Azarniya et al. [65] as optimal conditions.
to crack propagation and significant thermal anisotropy [13,65,69–71]. The x-ray diffraction (XRD) analysis was carried out to characterize
The microcracks often arise from the evolution of complex residual the existing phases in the as-synthesized AT (Philips Model PW3710,
stresses during the cooling step at grain boundaries and inside grains Netherlands). Also, field-emission scanning electron microscopy (FE-
[1], so that their formation culminates in alleviating the stress level SEM, Philips Model XL30, Netherlands) was used to monitor the
[72]. microstructure evolution and the distribution of the formed phases. To
Although a broad spectrum of practical studies have been carried out verify which type of Al2TiO5, TiO2 and Al2O3 phases may be evolved in
to explore the thermal decomposition of AT ceramic, its precise mech­ the synthesized powder, Raman spectroscopy was employed. Raman
anism is not completely identified. However, the experimental results studies were performed using Takram P50C0R10 spectrometer in 100×
suggest a nucleation and growth-controlled mechanism whose rate magnification with 638 nm laser excitation wavelength. To confirm the
peaks at 1100 ◦ C and the formed alumina particles can serve as pref­ Raman, XRD and FESEM results in the phase characterization, high-

2
A. Azarniya et al. Materials Characterization 173 (2021) 110764

Fig. 1. Schematic flow chart of the used synthesis procedure for nanostructured AT.

Fig. 2. (a–c) FESEM images of as-synthesized AT nanostructure in different magnifications; The particle size distribution of AT (d), Al2O3 (e) and TiO2 (f) phases.

resolution transmission electron microscopy (HR-TEM, JEOL JEM- 3. Results and discussion
2100F) was employed.
3.1. Characterization of as-received AT powder

2.2. Characterization of decomposed AT nanostructure Fig. 2 shows the FESEM images of the as-received AT. As can be seen,
a vast majority of the powder includes seemingly dense particles with a
To identify the temperature range in which AT nanostructure can be uniform size distribution and a diameter between 200 and 300 μm
decomposed to its parent phases, differential scanning calorimetry (DSC, (Fig. 2a). However, high-magnification images confirm the presence of
TA, Q600, USA) was employed in air at the heating rate of 10 ◦ C/min. two types of particles: The first type benefits a fully dense structure and
Two types of samples were heat treated at arbitrary temperatures to second one bears a relative porous grain configuration. While the grain
characterize which reactions are more likely to happen after the DTA size in the fully dense particles falls in the range of 55 nm, this value is
peak. One sample was heat treated at 1000 ◦ C for 33 h and the second around 600 nm for porous particles. As evidenced by Azarniya et al.
one was heated up to 1200 ◦ C for 5 min in air atmosphere. The phase [13], fully dense particles are composed of AT in β phase. Although it
content, phase transformation, and microstructure evolution of the forms a limited population of synthetic particles, relatively porous
samples were fully characterized through FESEM, HRTEM, Raman particles seem to be a composite blend of AT parent phases, i.e. Al2O3
analysis and XRD.

3
A. Azarniya et al. Materials Characterization 173 (2021) 110764

Fig. 3. (a) X-ray diffraction pattern and (b) Raman analysis of as-synthesized AT nanostructure; (c) XRD patterns of AT nanostructure after the heat treatment at (c1)
1000 ◦ C for 30 h and (c2) 1200 ◦ C for 5 h; and (d) Raman analysis of AT nanostructure after the heat treatment at (d1) 1000 ◦ C for 33 h, and (d2) 1200 ◦ C for 5 h.

and TiO2. Fig. 2(d–f) indicates the particle size distribution of these rooted in the partial hydration of aluminium ions during the sol-gel
phases. As shown, the particle size of Al2O3 phase falls in the range of process, non-stoichiometric measurement of starting materials (i.e.
100 nm, while this value is around 600 nm for TiO2 grains. Similar AlCl3 and TTIP), oxidation of present ions during the gelation through
morphologies have been reported for AT and Al2O3-TiO2 composite the exposure to the oxygen impurity or the slow formation of AT phase
blend by Azarniya et al. [65], Keyvani et al. [2] and Veettil Mani et al. from its parent phases. Based on the practical experience, some ap­
[77]. This fact is verified by XRD and Raman results in Fig. 3(a, b). As proaches may restrict the formation of Al2O3 and TiO2 during the syn­
can be seen, the synthetic powder is mainly composed of AT phase in β thesis of AT particles: (i) The use of exceedingly pure ethanol and
form, but there exist Al2O3 in corundum phase and TiO2 in anatase form nitrogen gas with no humidity, (ii) The maintenance of anhydrous AlCl3
as the synthesis impurities. The presence of Al2O3-TiO2 particles may be powder from the humidity, (iii) The use of an appropriate neutral

Fig. 4. HRTEM images of as-synthesized AT, showing nanoparticles of several nanometers in diameter [2].

4
A. Azarniya et al. Materials Characterization 173 (2021) 110764

Table 2 610, 644, and 893, respectively. Similar results have been reported in
A brief list of report temperature range for AT thermal dissociation. the literature for anatase, corundum and as-synthesized aluminium
Reported temperature range for AT thermal dissociation Ref. titanate particles. Accordingly, the broad bands at 236, 313, 441, 610,
and 893 cm− 1 may be attributed to the presence of AT nanostructure
700–1280 [9]
750–1280 [89–91] [78,79], while the peaks centered at 143 and 513 cm− 1 can be ascribed
750–1300 [66,92,93] to the existence of TiO2 particles in anatase form [80–83]. Besides,
780–1280 [94–96] relatively weak peaks at 253, 418 and 644 cm− 1 are wholly related to
800–1200 [77,97] Al2O3 in corundum form [79,84,85].
800–1280 [87,88,98]
800–1300 [99–101]
Fig. 4 indicates the size distribution of as-synthesized AT nano­
900–1280 [102–105] structure and its crystallographic planes by high-magnification HRTEM
images. As seen, the nanostructure is fragmented into nanoscale parti­
cles of about 20–30 nm in diameter, proving that the grain boundaries
between AT nanoparticles are not strong enough to preserve the total
cornerstone of AT nanostructure during the dispersion inside DI water
through the 3-h ultrasonication. The inter-planar distance measured by
the TEM image is around 3.36 nm. This value corresponds to the (101)
plane in β-Al2TiO5.

3.2. Characterization of decomposed AT nanostructure

One of the main disadvantages of AT-based ceramics which restricts


their industrial applications is the thermal instability within a wide
temperature range of around 700–1300 ◦ C. This range is controversial
about which a large population of empirical reports has been released.
Table 2 provides a brief list of the reported values. As seen, AT can be
entirely decomposed into Al2O3 and TiO2 from at least 700 to maximum
1300 ◦ C. To the best of our knowledge, there is only one paper which has
reported 250 ◦ C as the threshold temperature for the thermal decom­
Fig. 5. DTA-TGA plot of as-synthesized AT nanostructure at the heating rate of position of mechanically alloyed AT-based ceramics. In a systematic
10 ◦ C/min. study, Keyvani et al. [2] synthesized nanostructured AT through the sol-
gel method and harsh ball milling with the average grain size of around
atmosphere to prevent entering H2O molecules in the synthesis 20–30 nm and asserted that the thermal decomposition starts from
container, and (iv) The selection of an appropriate and optimized tem­ 250 ◦ C and happens through three subsequent stages. They attributed
perature for the calcination of AT gel. such an eminent temperature reduction to the nanostructuring nature of
The XRD results are in good agreement with the obtained Raman the powder due to the large content of interfaces and crystallographic
spectrum, as shown in Fig. 3b. As can been clearly seen, some intense imperfections, which act as preferred sites for triggering the thermal
peaks and some weak broad ones have emerged in the Raman analysis degradation. In our study, the DTA-TGA analysis was carried out to
results. They are centered at around 143, 236, 253, 313, 418, 441, 513, evaluate the temperature range in which AT nanostructure can be
decomposed into its parent phases. The results are given in Fig. 5. As can

Fig. 6. FESEM images of AT nanostructure after the heat treatment at 1000 ◦ C for 30 h: (a) distribution of as-heat treated AT nanostructure; (b–d) AT particles with
coarsened grains; and (e–g) Al2O3-TiO2 composite particles between coarsened unreacted AT grains.

5
A. Azarniya et al. Materials Characterization 173 (2021) 110764

Fig. 7. FESEM images of AT nanostructure after the heat treatment at 1200 ◦ C for 5 h: Whole AT particles are completely decomposed to bulky Al2O3 and lamellar
TiO2 phases with larger mean size.

be seen, the phase reaction slowly embarks on from a low temperature of decomposition is a nucleation-and-growth process, the heat treatment
around 300 ◦ C and reaches a peak within the temperature range of temperature plays an important role in the grain coarsening of newly
950–1200 ◦ C. This fact can be confirmed by the TGA curve. Accordingly, formed phases and reaction progress. Since heating AT particles for 5 h
the weight loss follows a linear function from 300 to 950 ◦ C, so that its at 1200 ◦ C is a relatively long-time process, TiO2 and Al2O3 particles are
slope is abruptly increased as the reaction is accelerated at 950–1200 ◦ C. comparatively larger than those formed at 1000 ◦ C for 33 h. This fact can
This fact indicates that the nanostructuring of AT ceramic can eminently be observed by comparing Fig. 6(f, g) and Fig. 7c).
reduce the decomposition temperature range from 900 down to 300C, To confirm the thermal decomposition of AT nanostructure and
but it is unable to change the temperature spectrum in which an over­ formation of TiO2 and Al2O3 phases, XRD analysis was conducted. The
whelming majority of AT ceramic is dissociated with a strong kinetics. results are given in Fig. 3c for the samples heat treated at 1000 ◦ C for 33
The literature have reported that the decomposition rate reaches a h and 1200 ◦ C for 5 h. In accordance with the results, AT is partially
maximum at around 1100 ◦ C [8,86–89]. decomposed to corundum and rutile, while a large content of AT par­
Fig. 6 shows the FE-SEM images of the AT nanostructure after the ticles are left intact. In contrast, as the reaction temperature increased,
heat treatment at 1000 ◦ C for 30 h at air. As can be seen, some significant more AT particles were thermally degraded, but not necessarily all of
changes occurred in the microstructure of the powder compared to the them. In means that the thermal dissociation of AT, even in the nano­
as-synthesized AT: (1) There are two types of microscale AT powders: structured form, proceeds with an exceedingly slow rate, so that AT will
the first type has very dense structure in which the nanoscale grains are be exhausted just provided a very long-time heat treatment at a suffi­
coarsened due to the long-time heat treatment (Fig. 6(b–d)); and (2) ciently high temperature is applied. The Raman spectra confirm the XRD
relatively porous particles in which very coarse micrometric grains can results in good approximation. The results are given in Fig. 3d. As
be observed beside remaining submicron AT particles (Fig. 6(e–g)). obviously seen, several vibrational bands including 143, 236, 253, 445,
Based on these facts, it seems that the AT particles have experienced two and 613 cm− 1 have emerged for the sample heat treated at 1000 ◦ C for
main phenomena: (1) The morphology and geometry of micrometric AT 33 h. The band located at 143 cm− 1 is related to anatase phase [80–83]
particles remained intact, but their constituent grains are significantly whose intensity is significantly diminished compared to as-synthesized
grown up to submicron level; and (2) The thermal decomposition has AT. On the other hand, the wide peak at 236 cm− 1 confirms the pres­
started in some AT particles so that large 2–3 μm Al2O3-TiO2 particles ence of unreacted AT. This peak is neighbored by a small shoulder at
have formed between remaining unreacted coarsened AT grains. These 253 cm− 1 which may be ascribed to the vibrational band of corundum
particles are either granular with flat surface or lamellar. Similar mor­ [78,79]. Finally, two subsequent sharp peaks have appeared at 445 and
phologies are reported by Sun Zhihua et al. [102] and Patel [106]. They 616 cm− 1 which are characteristic of rutile phase [80]. The aforemen­
have shown that nucleation-formed TiO2 particles may be distinguished tioned facts indicate that AT is partially decomposed to corundum and
by their lamellar structure. However, it is expected that all AT particles rutile after heat treating at 1000 ◦ C for 33 h, so that a large part of
are completely decomposed to bulky Al2O3 and lamellar TiO2 phases anatase is transformed to its high-temperature polymorph. In addition to
and all remaining AT grains get exhausted if the heating temperature is the mentioned peaks, some new peaks have been emerged for the sample
escalated. Fig. 7 shows the FE-SEM images of the AT nanostructure after heat treated at 1200 ◦ C for 5 h. While the intensity of two sharp peaks
the heat treatment at 1200 ◦ C for 5 h at air. As obviously seen, the related to the rutile have been sharpened, the intensity of the broad peak
microstructure is completely manipulated, so that there is no trace of AT related to AT has been reduced. Also, two new broad peaks have
nanoparticles in many particles which are studied by SEM. In other appeared at 730 and 870 cm− 1, which can be attributed to the formation
words, a large part of AT grains seems to be substituted with extraor­ of new corundum phase as a consequence of the reaction progress [107].
dinary lamellar TiO2 and bulky Al2O3 particles. Since the AT To further cast light on the phase transformation of AT

6
A. Azarniya et al. Materials Characterization 173 (2021) 110764

Fig. 8. HRTEM images of AT nanostructure after the heat treatment at (a) 1000 ◦ C for 30 h, and (b) 1200 ◦ C for 5 h, showing the formation of Al2O3 and TiO2 phases.

nanostructure, the samples heat treated at 1000 ◦ C for 33 h and 1200 ◦ C nanostructure bears a slow kinetics, so that a heat treatment at 1200 ◦ C
for 5 h were examined by HRTEM. The results are given in Fig. 8. As for 5 h cannot completely decompose this ceramic.
seen, the heat treatment at high temperatures has resulted in the grain
coarsening of AT particles and their byproducts. This physical phe­ Declaration of Competing Interest
nomenon is more evident in the sample heat treated at 1200 ◦ C for 5 h,
where the average grain size is around 300 nm. For the sample heat The authors declare that they have no known competing financial
treated at 1000 ◦ C for 33 h, the vast majority of particles studied by interests or personal relationships that could have appeared to influence
HRTEM were AT particles. Based on Fig. 8c, the inter-planar distance the work reported in this paper.
measured by the TEM image for this sample is around 2.65 nm. This
value is characteristic of the (230) plane in β-Al2TiO5. In contrast, a large References
part of the particles studied by TEM for the sample heat treated at 1200C
for 5 h were rutile and corundum, albeit a limited population of AT [1] K. Kornaus, R. Lach, M. Szumera, K. Świerczek, A. Gubernat, Synthesis of
aluminium titanate by means of isostructural heterogeneous nucleation, J. Eur.
particles could be identified. As given in Fig. 8(f, g), the (110) crystal Ceram. Soc. 39 (2019) 2535–2544.
plane of rutile and (012) plane of corundum phase could be character­ [2] N. Keyvani, A. Azarniya, H.R.M. Hosseini, M. Abedi, D. Moskovskikh, Thermal
ized by the inter-planar distances (i.e. 3.24 nm for rutile and 3.47 nm for stability and strain sensitivity of nanostructured aluminum titanate (Al2TiO5),
Mater. Chem. Phys. 223 (2019) 202–208.
corundum).
[3] W. Chen, A. Shui, C. Wang, J. Li, J. Ma, W. Tian, T. Ota, X. Xi, Preparation of
aluminum titanate flexible ceramic by solid-phase sintering and its mechanical
4. Conclusions behavior, J. Alloys Compd. 777 (2019) 119–126.
[4] F. Bakhshandeh, A. Azarniya, H.R. Madaah Hosseini, S. Jafari, Are aluminium
titanate-based nanostructures new photocatalytic materials? Possibilities and
In the present study, aluminium titanate (AT or Al2TiO5) nano­ perspectives, J. Photochem. Photobiol. A Chem. 353 (2018), https://doi.org/
structure was synthesized by a citrate sol-gel method and its thermal 10.1016/j.jphotochem.2017.11.043.
instability was explored in some details. The main purpose was whether [5] M.A. Violini, M.F. Hernández, M. Gauna, G. Suárez, M.S. Conconi, N.
M. Rendtorff, Low (and negative) thermal expansion Al2TiO5 materials and
the nanostructuring nature of this ceramic can significantly affect its Al2TiO5-3Al2O3. 2SiO2-ZrTiO4 composite materials. Processing, initial zircon
thermal decomposition and formed byproducts. The results showed that proportion effect, and properties, Ceram. Int. 44 (2018) 21470–21477.
the AT nanostructure starts decomposing to corundum and rutile from a [6] C.C. Guedes-Silva, F.M. de S. Carvalho, T. dos S. Ferreira, L.A. Genova, Formation
of aluminum titanate with small additions of MgO and SiO2, Mater. Res. 19
low temperature of around 300 ◦ C and reaches a peak within the tem­ (2016) 384–388.
perature range of 950–1200 ◦ C. Such an abrupt reduction in the reaction [7] F. Hassan, H.J. Woo, N.A. Aziz, M.Z. Kufian, S.R. Majid, Synthesis of Al 2 TiO 5
temperature from 900 to 300 ◦ C was attributed to the increase of in­ and its effect on the properties of chitosan–NH 4 SCN polymer electrolytes, Ionics
(Kiel) 19 (2013) 483–489.
terfaces density and crystallographic imperfections, serving as prefer­ [8] G. Feng, F. Jiang, W. Jiang, J. Liu, Q. Zhang, Q. Wu, Q. Hu, L. Miao, J. Liang,
ential sites for starting the thermal dissociation. However, the Low-temperature preparation of novel stabilized aluminum titanate ceramic
nanostructuring is not able to change the temperature spectrum in fibers via nonhydrolytic sol-gel method through linear self-assembly of
precursors, Ceram. Int. 45 (2019) 18704–18709.
which a vast majority of AT ceramic is dissociated with a strong kinetics
[9] R. Naghizadeh, H.R. Rezaie, F. Golestani-Fard, The influence of composition,
(i.e. 1100 ◦ C). It was confirmed that the thermal instability of AT cooling rate and atmosphere on the synthesis and thermal stability of aluminum
titanate, Mater. Sci. Eng. B 157 (2009) 20–25.

7
A. Azarniya et al. Materials Characterization 173 (2021) 110764

[10] A. Borrell, M.D. Salvador, V.G. Rocha, A. Fernández, T. Molina, R. Moreno, [42] I.J. Kim, H. Supkwak, Thermal shock resistance and thermal expansion behaviour
Enhanced properties of alumina–aluminium titanate composites obtained by with composition and microstructure of Al2TiO5 ceramics, Can. Metall. Q. 39
spark plasma reaction-sintering of slip cast green bodies, Compos. Part B Eng. 47 (2000) 387–396.
(2013) 255–259. [43] Z. Li, R.C. Bradt, Thermal expansion of the cubic (3C) polytype of SiC, J. Mater.
[11] A. Shokuhfar, M.N. Samani, N. Naserifar, P. Heidary, G. Naderi, Prediction of Sci. 21 (1986) 4366–4368, https://doi.org/10.1007/BF01106557.
physical properties of Al2TiO5-based ceramics containing micro and nano size [44] R. Maki, Y. Suzuki, Microstructure and mechanical properties of MgO-doped
oxide additives by using artificial neural network. Vorhersage der physikalischen Al2TiO5 prepared by reactive sintering, J. Ceram. Soc. Japan 121 (2013)
Eigenschaften von Al2TiO5-basierenden Keramiken mit Mikro-und 568–571.
Nanoadditiven mittels künstlicher neuronaler Netze, Mater. Und [45] M. Takeda, Y. Imai, H. Ichikawa, N. Kasai, T. Seguchi, K. Okamura, Thermal
Werkstofftechnik Entwicklung, Fert. Prüfung, Eig. Und Anwendungen Tech. stability of SiC fiber prepared by an irradiation-curing process, Compos. Sci.
Werkstoffe 40 (2009) 169–177. Technol. 59 (1999) 793–799.
[12] A. Azarniya, H.R.M. Hosseini, A new method for fabrication of in situ Al/ [46] K. Niihara, New design concept of structural ceramics, J. Ceram. Soc. Japan 99
Al3Ti–Al2O3 nanocomposites based on thermal decomposition of nanostructured (1991) 974–982.
tialite, J. Alloys Compd. 643 (2015) 64–73. [47] L. Gao, H.Z. Wang, J.S. Hong, H. Miyamoto, K. Miyamoto, Y. Nishikawa, S. Torre,
[13] A. Azarniya, H.R.M. Hosseini, M. Jafari, N. Bagheri, Thermal decomposition of Mechanical properties and microstructure of nano-SiC–Al2O3 composites
nanostructured Aluminum Titanate in an active Al matrix: A novel approach to densified by spark plasma sintering, J. Eur. Ceram. Soc. 19 (1999) 609–613.
fabrication of in situ Al/Al<inf>2</inf>O<inf>3</inf>-Al<inf>3</inf>Ti [48] E.D. Case, J.R. Smyth, O. Hunter Jr., Microcracking in large-grain Al2O3, Mater.
composites, Mater. Des. 88 (2015), https://doi.org/10.1016/j. Sci. Eng. 51 (1981) 175–179.
matdes.2015.09.050. [49] N. Soga, O.L. ANDERSON, High-temperature elastic properties of polycrystalline
[14] M.S. Ahmadvand, A. Azarniya, H.R.M. Hosseini, Thermomechanical synthesis of MgO and Al2O3, J. Am. Ceram. Soc. 49 (1966) 355–359.
hybrid in-situ Al-(Al3Ti+ Al2O3) composites through nanoscale Al-Al2TiO5 [50] C. Gault, F. Platon, D. Le Bras, Ultrasonic measurements of Young’s modulus of
reactive system, J. Alloys Compd. 789 (2019) 493–505. Al2O3-based refractories at high temperatures, Mater. Sci. Eng. 74 (1985)
[15] T. Hono, N. Inoue, M. Morimoto, Y. Suzuki, Reactive sintering and microstructure 105–111.
of uniform, openly porous Al2TiO5, J. Asian Ceram. Soc. 1 (2013) 178–183. [51] C.H. Chen, H. Awaji, Mechanical properties of Al2TiO5 ceramics, in: Key Eng.
[16] N.D. Trung, H.C. Anh, N. Tri, P. PH, H.T. Cuong, A low temperature fabrication Mater, Trans Tech Publ, 2007, pp. 1417–1419.
and Photoactivity of Al2TiO5 in cinnamic acid degradation, Mater. Trans. 60 [52] Z.-H. Zhang, F.-C. Wang, J. Luo, S.-K. Lee, L. Wang, Processing and
(2019) 2022–2027. characterization of fine-grained monolithic SiC ceramic synthesized by spark
[17] B. Freudenberg, Aluminum titanate, in: Concise Encycl. Adv. Ceram. Mater, plasma sintering, Mater. Sci. Eng. A 527 (2010) 2099–2103.
Elsevier, 1991, pp. 20–22. [53] W. Dienst, Reduction of the mechanical strength of Al2O3, AlN and SiC under
[18] R. Barea, M. Belmonte, M.I. Osendi, P. Miranzo, Thermal conductivity of Al2O3/ neutron irradiation, J. Nucl. Mater. 191 (1992) 555–559.
SiC platelet composites, J. Eur. Ceram. Soc. 23 (2003) 1773–1778. [54] J.-S. Lee, M. Imai, T. Yano, Fabrication and mechanical properties of oriented SiC
[19] Y. Hu, G. Du, N. Chen, A novel approach for Al2O3/epoxy composites with high short-fiber-reinforced SiC composite by tape casting, Mater. Sci. Eng. A 339
strength and thermal conductivity, Compos. Sci. Technol. 124 (2016) 36–43. (2003) 90–95.
[20] D.J. Senor, D.J. Trimble, G.E. Youngblood, G.A. Newsome, C.E. Moore, J. [55] L. Gao, J.S. Hong, H. Miyamoto, S. Torre, Bending strength and microstructure of
J. Woods, Effects of neutron irradiation on thermal conductivity of SiC-based Al2O3 ceramics densified by spark plasma sintering, J. Eur. Ceram. Soc. 20
composites and monolithic ceramics, Fusion Technol. 30 (1996) 943–955. (2000) 2149–2152.
[21] Y.W. Lo, W.C.J. Wei, C.H. Hsueh, Low thermal conductivity of porous Al2O3 [56] S. Suyama, T. Kameda, Y. Itoh, Development of high-strength reaction-sintered
foams for SOFC insulation, Mater. Chem. Phys. 129 (2011) 326–330. silicon carbide, Diam. Relat. Mater. 12 (2003) 1201–1204.
[22] D.-M. Liu, B.-W. Lin, Thermal conductivity in hot-pressed silicon carbide, Ceram. [57] G.C. Wei, P.F. Becher, Improvements in mechanical properties in SiC by the
Int. 22 (1996) 407–414. addition of TiC particles, J. Am. Ceram. Soc. 67 (1984) 571–574.
[23] Z. Živcová, E. Gregorová, W. Pabst, D.S. Smith, A. Michot, C. Poulier, Thermal [58] X. Teng, H. Liu, C. Huang, Effect of Al2O3 particle size on the mechanical
conductivity of porous alumina ceramics prepared using starch as a pore-forming properties of alumina-based ceramics, Mater. Sci. Eng. A 452 (2007) 545–551.
agent, J. Eur. Ceram. Soc. 29 (2009) 347–353. [59] A.S. Ramos, M.A. de Souza, R. de O. Magnago, C. dos Santos, C.A.A. da Silva,
[24] E. Volz, A. Roosen, W. Hartung, A. Winnacker, Electrical and thermal B. de Almeida Fortes, Sintering of Al2O3-TiO2 Mixtures Obtained by High-Energy
conductivity of liquid phase sintered SiC, J. Eur. Ceram. Soc. 21 (2001) Ball Milling, in: Adv. Sci. Technol, Trans Tech Publ, 2014, pp. 157–161.
2089–2093. [60] M. Hotta, T. Goto, Densification and microstructure of Al2O3-cBN composites
[25] T. Shimizu, K. Matsuura, H. Furue, K. Matsuzak, Thermal conductivity of high prepared by spark plasma sintering, J. Ceram. Soc. Japan 116 (2008) 744–748.
porosity alumina refractory bricks made by a slurry gelation and foaming [61] Z. Shen, M. Johnsson, Z. Zhao, M. Nygren, Spark plasma sintering of alumina,
method, J. Eur. Ceram. Soc. 33 (2013) 3429–3435. J. Am. Ceram. Soc. 85 (2002) 1921–1927.
[26] H. Nakano, K. Watari, Y. Kinemuchi, K. Ishizaki, K. Urabe, Microstructural [62] T. Korim, I. Kotsis, Effect of additives on the properties of Al2TiO5 ceramics, in:
characterization of high-thermal-conductivity SiC ceramics, J. Eur. Ceram. Soc. Mater. Sci. Forum, Trans Tech Publ, 2003, pp. 117–120.
24 (2004) 3685–3690. [63] S.J. Kalita, V. Somani, Al2TiO5–Al2O3–TiO2 nanocomposite: structure,
[27] K. Watari, H. Nakano, K. Sato, K. Urabe, K. Ishizaki, S. Cao, K. Mori, Effect of mechanical property and bioactivity studies, Mater. Res. Bull. 45 (2010)
grain boundaries on thermal conductivity of silicon carbide ceramic at 5 to 1300 1803–1810.
K, J. Am. Ceram. Soc. 86 (2003) 1812–1814. [64] V. Somani, S.J. Kalita, Synthesis, densification, and phase evolution studies of
[28] D.S. Smith, S. Fayette, S. Grandjean, C. Martin, R. Telle, T. Tonnessen, Thermal Al2O3–Al2TiO5–TiO2 nanocomposites and measurement of their electrical
resistance of grain boundaries in alumina ceramics and refractories, J. Am. properties, J. Am. Ceram. Soc. 90 (2007) 2372–2378.
Ceram. Soc. 86 (2003) 105–111. [65] A. Azarniya, A. Azarniya, H.R.M. Hosseini, A. Simchi, Nanostructured aluminium
[29] X. Wu, H. Ma, X. Chen, Z. Li, J. Li, Thermal conductivity and microstructure titanate (Al2TiO5) particles and nanofibers: synthesis and mechanism of
properties of porous SiC ceramic derived from silicon carbide powder, 2013. microstructural evolution, Mater. Charact. 103 (2015), https://doi.org/10.1016/
[30] P. Auerkari, Mechanical and Physical Properties of Engineering Alumina j.matchar.2015.03.030.
Ceramics, Technical Research Centre of Finland Espoo, 1996. [66] D. Kim, H.-J. Kim, H.-T. Kim, J. Namkung, I. Kim, S.-C. Choi, S.-S. Ryu,
[31] E. Zapata-Solvas, S. Bonilla, P.R. Wilshaw, R.I. Todd, Preliminary investigation of Mechanical properties of Al2TiO5 ceramics for high temperature application,
flash sintering of SiC, J. Eur. Ceram. Soc. 33 (2013) 2811–2816. Curr. Nanosci. 10 (2014) 154–158.
[32] Y. Wang, C. Suryanarayana, L. An, Phase transformation in nanometer-sized [67] P. Innocenzi, A. Martucci, L. Armelao, S. Licoccia, M.L. Di Vona, E. Traversa,
γ-alumina by mechanical milling, J. Am. Ceram. Soc. 88 (2005) 780–783. Sol− gel synthesis of β-Al2TiO5 thin films at low temperature, Chem. Mater. 12
[33] E. Kato, K. Daimon, J. Takahashi, Decomposition temperature of β-Al2TiO5, (2000) 517–524.
J. Am. Ceram. Soc. 63 (1980) 355–356. [68] M. Zaharescu, M. Crisan, M. Preda, V. Fruth, S. Preda, Al2TiO5-based ceramics
[34] E. Volceanov, A.M. Gurban, A. Volceanov, P. Niţă, High thermal shock resistant obtained by hydrothermal process, J. Optoelectron. Adv. Mater. 5 (2003)
aluminium titanate type ceramics, in: Key Eng. Mater., Trans Tech Publ, 2004, 1411–1414.
pp. 993–996. [69] G.F. Zhao, Y. Bai, L.J. Qiao, Porous aluminum titanate ceramics prepared using
[35] T. Shimoo, K. Okamura, F. Toyoda, Thermal stability of SiO2-coated SiC fiber (Hi- graphite as pore former, in: Adv. Mater. Res, Trans Tech Publ, 2014, pp. 189–192.
Nicalon) under argon atmosphere, J. Mater. Sci. 35 (2000) 3811–3816. [70] N. Sarkar, J.G. Park, S. Mazumder, A. Pokhrel, C.G. Aneziris, I.J. Kim,
[36] Y. Wu, Y. Zhang, G. Pezzotti, J. Guo, Influence of AlF3 and ZnF2 on the phase Al2TiO5–mullite porous ceramics from particle stabilized wet foam, Ceram. Int.
transformation of gamma to alpha alumina, Mater. Lett. 52 (2002) 366–369. 41 (2015) 6306–6311.
[37] D.N. Talwar, J.C. Sherbondy, Thermal expansion coefficient of 3C–SiC, Appl. [71] G. Zhao, Y. Bai, L. Qiao, Aluminum titanate-calcium dialuminate composites with
Phys. Lett. 67 (1995) 3301–3303. low thermal expansion and high strength, J. Alloys Compd. 656 (2016) 1–4.
[38] M. Halvarsson, V. Langer, S. Vuorinen, Determination of the thermal expansion of [72] B. Dittert, M. Wiessner, P. Angerer, J.M. Lackner, H. Leichtfried, Tailored Al2O3-
κ-Al2O3 by high temperature XRD, Surf. Coat. Technol. 76 (1995) 358–362. Al2TiO5-TiO2 composite ceramics from different titanium precursors, Arch.
[39] S. Bueno, R. Moreno, C. Baudı ́n, Design and processing of Al2O3–Al2TiO5 layered Metall. Mater. 64 (2019).
structures, J. Eur. Ceram. Soc. 25 (2005) 847–856. [73] I. Low, Z. Oo, In situ diffraction study of self-recovery in aluminum titanate,
[40] Z. Li, R.C. Bradt, Thermal expansion of the hexagonal (4 H) polytype of SiC, J. Am. Ceram. Soc. 91 (2008) 1027–1029.
J. Appl. Phys. 60 (1986) 612–614. [74] M. Mirzaee, M.M. Amini, M. Sharbatdaran, Hydrothermal assisted sol-gel process
[41] O. Sarikaya, Effect of the substrate temperature on properties of plasma sprayed on binary mixtures of aluminum and titanium alkoxides: a novel route for the
Al2O3 coatings, Mater. Des. 26 (2005) 53–57. fabrication of single phase tialite, J. Adv. Mater. Process. 5 (2017) 73–81.

8
A. Azarniya et al. Materials Characterization 173 (2021) 110764

[75] I.M. Low, W.K. Pang, In situ diffraction study of self-recovery in vacuum [91] H.R. Rezaie, M. Sobhani, R. Naghizadeh, Formation and decomposition of sol-gel
decomposed Al 2TiO5, J. Aust. Ceram. Soc. 49 (2013) 48–52. synthesized aluminum titanate nano powders at the presence of Fe2O3 additive,
[76] C. Chen, B.R. Müller, C. Prinz, J. Stroh, I. Feldmann, G. Bruno, The correlation in: Defect Diffus. Forum, Trans Tech Publ, 2008, pp. 549–553.
between porosity characteristics and the crystallographic texture in extruded [92] Y.H. Wang, G. Chen, Z.S. Wang, J.W. Liu, P.F. Luo, Improvement of microcracks
stabilized aluminium titanate for diesel particulate filter applications, J. Eur. resistance of porous aluminium titanate ceramic membrane support using
Ceram. Soc. 40 (2020) 1592–1601. attapulgite clay as additive, Ceram. Int. 44 (2018) 2077–2084.
[77] T.V. Mani, H.K. Varma, K.G. Warrier, A.D. Damodaran, Aluminum titanate [93] M. Özçatal, M. SerhatBaşpınar, The Effects of Temperature and Additives on the
powder synthesis via thermal decomposition of transparent gels, J. Am. Ceram. Microstructure of Al2TiO5, (n.d.).
Soc. 74 (1991) 1807–1810. [94] G. Xu, H. Zhao, H. Cui, Z. Zhang, J. Li, Stability, microstructure and mechanical
[78] L. Bonhomme-Coury, N. Lequeux, S. Mussotte, P. Boch, Preparation of Al 2 TiO 5- properties of (Al, Fe) 2TiO5 porous ceramic reinforced by in-situ mullite,
ZrO 2 mixed powders via sol-gel process, J. Sol-Gel Sci. Technol. 2 (1994) J. Ceram. Soc. Japan 123 (2015) 156–159.
371–375. [95] G.G. Xu, H.L. Zhao, H.Z. Cui, Y.H. Ma, J.X. Ding, Z.H. Zhang, Fabrication and
[79] S.J. Webb, N.M. Alford, S.J. Penn, A. Templeton, X. Wang, The use of Raman characterisation of Al2TiO5-mullite porous composite ceramics reinforced by
spectroscopy for characterisation of ceramic dielectrics, Nondestruct. Test. Eval. alumina platelets, Mater. Res. Innov. 18 (2014) 477–480.
17 (2001) 205–212. [96] G. Xu, Y. Ma, G. Ruan, H. Cui, Z. Zhang, B. Bai, Preparation of porous Al2TiO5
[80] A. Wypych, I. Bobowska, M. Tracz, A. Opasinska, S. Kadlubowski, A. Krzywania- ceramics reinforced by in situ formation of mullite whiskers, Mater. Des. 47
Kaliszewska, J. Grobelny, P. Wojciechowski, Dielectric properties and (2013) 57–60.
characterisation of titanium dioxide obtained by different chemistry methods, [97] R. Papitha, M.B. Suresh, D. Chakravarty, A. Swarnakar, D. Das, R. Johnson,
J. Nanomater. 2014 (2014). Eutectoid decomposition of aluminum titanate (Al2TiO5) ceramics under Spark
[81] T. Mazza, E. Barborini, P. Piseri, P. Milani, D. Cattaneo, A.L. Bassi, C.E. Bottani, Plasma (SPS) and Conventional (CRH) thermal treatments, Ceram. Int. 40 (2014)
C. Ducati, Raman spectroscopy characterization of Ti O 2 rutile nanocrystals, 659–666.
Phys. Rev. B 75 (2007) 45416. [98] M. Sharma, D.K. Singh, R.K. Upadhyay, M.S. Yadav, S.S. Amritphale, N. Chandra,
[82] M. Ocana, J.V. Garcia-Ramos, C.J. Serna, Low-temperature nucleation of rutile Novel approach for sol–gel synthesis of nanosize aluminium titanate, Mater. Res.
observed by Raman spectroscopy during crystallization of TiO2, J. Am. Ceram. Innov. 18 (2014) 235–240.
Soc. 75 (1992) 2010–2012. [99] L. Li, Q. Wang, G. Liao, K. Li, G. Ye, Densification behavior of mullite-Al2TiO5
[83] V. Swamy, B.C. Muddle, Q. Dai, Size-dependent modifications of the Raman composites by reaction sintering of natural andalusite and TiO2, Ceram. Int. 44
spectrum of rutile Ti O 2, Appl. Phys. Lett. 89 (2006) 163118. (2018) 3981–3986.
[84] A. Aminzadeh, Excitation frequency dependence and fluorescence in the Raman [100] N. Sarkar, K.S. Lee, J.G. Park, S. Mazumder, C.G. Aneziris, I.J. Kim, Mechanical
spectra of Al2O3, Appl. Spectrosc. 51 (1997) 817–819. and thermal properties of highly porous Al2TiO5–Mullite ceramics, Ceram. Int.
[85] K.R. Nagabhushana, B.N. Lakshminarasappa, F. Singh, Photoluminescence and 42 (2016) 3548–3555.
Raman studies in swift heavy ion irradiated polycrystalline aluminum oxide, Bull. [101] R. Papitha, M. Buchi Suresh, D. Dibakar, R. Johnson, High-temperature flexural
Mater. Sci. 32 (2009) 515. strength and thermal stability of near zero expanding doped aluminum titanate
[86] V. Buscaglia, P. Nanni, Decomposition of Al2TiO5 and Al2 (1-x) MgxTi (1+ x) O5 ceramics for diesel particulate filters applications, Int. J. Appl. Ceram. Technol.
ceramics, J. Am. Ceram. Soc. 81 (1998) 2645–2653. 11 (2014) 773–782.
[87] V. Singh, G. Jayarao, R. Mazumder, Phase evolution and stability of aluminium [102] 孙志华, 刘开平, 汪敏强, 刘民武, 纳米 Fe2O3 对钛酸铝陶瓷热稳定性能的影响, 硅
titanate prepared by solid-state route: effect of two different particle size of 酸盐学报 41, 2013, pp. 437–442.
Al2O3, Trans. Indian Ceram. Soc. 75 (2016) 170–174. [103] R.D. Skala, D. Li, I.M. Low, Diffraction, structure and phase stability studies on
[88] G. Feng, W.H. Jiang, J.M. Liu, Q. Zhang, Z. Hu, L.F. Miao, Q. Wu, Low- aluminium titanate, J. Eur. Ceram. Soc. 29 (2009) 67–75.
temperature synthesis of magnesium-stabilized aluminum titanate powder via [104] I.M. Low, Z. Oo, Reformation of phase composition in decomposed aluminium
non-hydrolytic sol-gel method, in: Mater. Sci. Forum, Trans Tech Publ, 2016, titanate, Mater. Chem. Phys. 111 (2008) 9–12.
pp. 319–323. [105] I.M. Low, D. Lawrence, R.I. Smith, Factors controlling the thermal stability of
[89] H.J. Li, X.S. Shao, K. Li, Z.H. Geng, P. Wu, Thermal stability of aluminum titanate aluminum titanate ceramics in vacuum, J. Am. Ceram. Soc. 88 (2005)
ceramics, in: Key Eng. Mater, Trans Tech Publ, 2016, pp. 137–140. 2957–2961.
[90] W.H. Jiang, Q.M. Jiang, J.M. Liu, Q.X. Zhu, Q. Zhang, Prepartion of stabilized [106] P. Patel, Nanomaterials antimicrobial titanium dioxide coatings activated by
aluminum titanate film via nonhydrolytic Sol-gel route, in: Adv. Mater. Res, Trans indoor light, Chem. Eng. News 97 (2019) 7.
Tech Publ, 2012, pp. 96–100. [107] D. Tuschel, Photoluminescence Spectroscopy Using a Raman Spectrometer, 2016.

You might also like