You are on page 1of 46

Author’s Accepted Manuscript

Properties and microstructures of lightweight


alumina containing different types of nano-alumina

Lvping Fu, Ao Huang, Huazhi Gu, Hongwei Ni

www.elsevier.com/locate/ceri

PII: S0272-8842(18)31708-5
DOI: https://doi.org/10.1016/j.ceramint.2018.06.261
Reference: CERI18705
To appear in: Ceramics International
Received date: 4 April 2018
Revised date: 15 June 2018
Accepted date: 29 June 2018
Cite this article as: Lvping Fu, Ao Huang, Huazhi Gu and Hongwei Ni,
Properties and microstructures of lightweight alumina containing different types
of nano-alumina, Ceramics International,
https://doi.org/10.1016/j.ceramint.2018.06.261
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Properties and microstructures of lightweight alumina

containing different types of nano-alumina


Lvping Fu*1, Ao Huang, Huazhi Gu, Hongwei Ni

The State Key Laboratory of Refractories and Metallurgy, Wuhan University of Science and Technology, Wuhan,
Hubei 430081, China
*Corresponding author: Tel: +862768862160; Fax:+862768862160; E-mail: flpwust@163.com

Abstract

The key challenge in designing lightweight wear linings for industrial furnaces lies in the

fabrication of porous materials with the smallest possible open porosity and pore size. In this study,

in order to increase the migration velocity of the grain boundary, alumina sol and nanosized

α-Al2O3 were introduced for their high temperature grain boundary superplasticity. Investigations

on the effect of various types of nano-alumina additives on pore repartition and some properties of

sintered lightweight alumina were carried out. The introduction of nano-alumina increased the

surface stress and accelerated the division of pores. The addition of alumina sol and nanosized

α-Al2O3 showed similar effects on the physical properties of lightweight alumina. Compared to

samples without addition of nano-alumina, samples containing alumina sol and nanosized α-Al2O3

exhibited the increased closed porosities (from 2.8% to 7.1% and 9.2%, respectively), decreased

open porosities (from 12.4% to 10.5% and 5.2%, respectively), lower thermal conductivities

(decreasing amplitudes of 16% and 9%, respectively) and increased intracrystalline porosity.

Furthermore, the addition of different types of nano-alumina led to microstructural differences in

the samples. Abnormally grown grains, with a size of approximately 100-200grain μm, containing

numerous smaller intracrystalline pores inside were observed in the sample added alumina sol. The

1 Permanent address: No. 947 Heping Avenue, Qingshan District, Wuhan, Hubei, P.R. China, Zip: 430081
sample containing nanosized α-Al2O3 exhibited a homogeneous distribution of grain sizes. The

different grain growth rates, accompanying with various types of nano-alumina, were responsible

for the differences in microstructure.

Keywords: lightweight alumina; nano-alumina; closed-pores; microstructures; properties

1 Introduction

Refractories constitute essential materials for basic industries such as iron/steel making,

nonferrous metallurgy, inorganic materials production (cement, glass, ceramic…), and chemical

engineering. The improvements in the energy efficiency and cleanliness of industrial processes are

often linked to new furnace technologies involving high-performance refractories. Therefore,

refractories are no longer considered low-cost, consumable materials, but functional and structural

advanced materials for high-temperature applications. The energy preservation and emission

reduction of industrial furnaces has recently attained prime importance.

From the hotter surface, refractory linings are constituted of several layers, which are in charge

of resisting high-temperature attack (wear lining) and protecting the metallic casing from heat

(insulation lining). The literature has reported that the heat insulation materials exhibited a better

heat insulation function when positioned as close as possible to the wear lining [1]. Therefore, the

exploration of lightweight porous materials that could be used directly in wear lining in the steel

industry has attracted increasing attention. However, as wear lining, increasing porosity is well

known to affect the slag corrosion, thermal shock, and wear resistances against molten metals. Open

porosity constitutes a preferential way for slag penetration and hence promotes corrosion of the

refractory. Moreover, large pores have a detrimental effect on mechanical strength and slag

corrosion resistance. Thus, the target of this work is to process an alumina-based lightweight

refractory with a high porosity with preferentially closed and as small pores as possible, which is

also beneficial for diminishing the thermal conductivity [2,3].


Two methods are currently used to obtain a controlled porosity in sintered materials:

Dispersing particles with controlled granulometry into the green material. The elimination of

these particles during the heat treatment leaves pores that remain in the material after sintering.

Some examples are the use of pore-forming agents [4–7], in-situ decomposition of organic or

inorganic matter [8–11], direct foaming [12–16], and the freeze-drying method [17–21].

Introducing some additives or fine particles or adapting the process or the initial composition

can facilitate the persistence of pores in the sintered material. Partial sintering [22–25] or

reaction-bonding [26–28] methods are often used.

In the first method, the evacuation of decomposed matter out of the green bulk requires the

creation of channels, promoting the creation of open porosity. Reducing the diffusion rate during

sintering is the main action of the second technique. However, slower diffusion also results in

slower grain boundary migration, promoting the persistence of open porosity. Thus, the fabrication

of porous materials, characterized by a mainly closed, fine, and intragranular porosity, remains a

difficult technical issue for ceramics and refractories.

Interaction between pores and grain boundaries are of prime importance in densification during

sintering [29]. Pores, which act as obstacles for grain-boundary migrations, can inhibit grain growth.

However, if grain-boundary migration is rapid enough to overlap the pores, abnormal grain growth

will occur and pores will be trapped in grains. Therefore, the increase in grain-boundary migration

velocity is of interest to promote closed, intergranular porosity. Abnormal grain growth is a widely

observed effect in doped or impure alumina. For example, alumina materials doped with Y2O3 [30],

TiO2 [31,32], SiO2 [30–32] or CaO [33] were proved to be sensitive to abnormal grain growth.

However, in these cases, the creation of intracrystalline pores is generally limited.

Superplasticity is an important characteristic of nanosized materials [34,35]. This feature

enables them to deform, by means of grain-boundary sliding, under the application of stress. In our

previously published work, nanosized alumina sol was introduced during the fabrication of

microporous alumina aggregates in order to increase the boundary-migration velocity. The


nano/micro double-scale effect created partial regions in which the alumina sol collected, which

showed a larger surface stress, which could increase the boundary migration velocities [36]. Thus,

the introduction of alumina sol increased the closed porosity and decreased the bulk density of

materials. Although the formation of closed pores by adding nano-alumina has been proposed, the

division behaviour of closed pores, which results in decreased pore size, remains poorly understood.

Moreover, alumina is known to exist in a number of metastable polymorphs in addition to the

thermodynamically stable α-Al2O3. Although those metastable polymorphs are reported to

transform into α-Al2O3 during annealing, both phases still exhibit different sintering kinetics.

Therefore, in this study, lightweight alumina was fabricated by adding different types of

nano-alumina into α-Al2O3 micropowder. Microstructural mechanisms were proposed to explain the

effect of nano-alumina on the division of closed pores during sintering and the influence of the

types of nano-alumina on the microstructures and properties of fabricated lightweight alumina.

2 Experimental

2.1 Raw materials

α-Al2O3 micropowder (D50=2.36 μm, specific surface area=4.32 m2·g-1, Kaifeng Special

Refractory Co., Ltd., China), nanosized α-Al2O3 (mean particle diameter=30.3 nm, specific surface

area=205.39 m2·g-1, Shanghai Macklin Biochemical Co., Ltd., China), and alumina sol with 25 wt%

solid loading (mean particle diameter=26.4 nm, specific surface area= 315.52 m2·g-1, Zibo Jinqi

Chemical Technology Co., Ltd., China) were selected as raw materials. Corn starch (D50=14.287

μm, Shandong Hengren Industrial Trade Co., Ltd., China) was used as a binder and pore former.

The particle size distributions of α-Al2O3 micropowder and corn starch were measured by a

laser particle size analyser (Mastersizer 2000, Malvern Instruments Co., Ltd., UK), and the particle

size distributions of nanosized α-Al2O3 and alumina sol were tested by a nanoparticle analyser

(SZ-100, Horiba Scientific, Germany); the measurement results are shown in Fig. 1.

The microstructures of nanosized α-Al2O3 and alumina sol were observed by a transmission
electron microscope (TEM, 2100 UHR STEM/EDS JEOL, Japan). As shown in Fig. 2, the vast

majority of particle sizes of two nano-aluminas were both around 30 nm.

The phase compositions of α-Al2O3 micropowder and nanosized α-Al2O3 were tested by an

X-ray powder diffractometer (XRD, Xpert Pro MPD, PANalytical B.V., Holland) using Ni-filtered

CuKα radiation (λ=0.1542 nm) at a scan speed of 5°/min and temperature of 25 °C. As shown in Fig.

3, only the α-Al2O3 (corundum) phase has been detected in both raw materials. However, the

intensity of the corundum peaks in α-Al2O3 micropowder is much stronger, which indicates a better

crystallinity.

In order to confirm the existence form of alumina sol during the heat treatment, it was dried at

110 °C for 48 h, and XRD and thermal analyses were carried out on the obtained dried powder. The

phase composition was tested by an X-ray powder diffractometer at a scan speed of 5°/min and

temperature of 25–1200 °C. Thermogravimetric (TG) and differential scanning calorimetry (DSC)

traces of dried powder were measured by a thermal analyser (STA449F3, Netzsch, Germany) with a

heating rate of 10 °C/min in flowing air.

The main phase of the dried powder is AlOOH (boehmite) is shown in Fig. 4 (a). The peaks of

the γ-Al2O3 phase were detected when the temperature was increased to 400 °C, which indicates

that the transformation of boehmite into γ-Al2O3 starts around this temperature. At 1000 °C, in

addition to strong γ-Al2O3 peaks, obvious peaks of θ-Al2O3 and α-Al2O3 (corundum) were observed.

When the temperature increased to 1200 °C, the peak intensity of α-Al2O3 increased, whereas that

of γ-Al2O3 and θ-Al2O3 decreased.

Typical TG and DSC traces of dried powder are shown in Fig. 4 (b). During the heat treatment,

there are three weight-loss regions. The first is at room temperature to around 100 °C, which is

attributed to the removal of physically absorbed water. The second is 100–480 °C, which may be

due to the removal of attached hydroxyls in AlOOH. The last one is 480–1400 °C; a slight weight
loss was observed in that region. Two endothermic peaks, one around 80.2 °C and the other around

445.2 °C, were detected, which correspond to the removal of absorbed water and the transformation

of boehmite into γ-Al2O3, respectively. The formation of α-Al2O3 results in an exothermic peak

around 1199.2 °C.

A number of studies reported that, during heat treatment, the boehmite would transform in an

orderly manner into γ-Al2O3, δ-Al2O3, θ-Al2O3, and finally α-Al2O3 at around 1200 °C [37]. Our

results are in good agreement with the literatures, except for the detection of δ-Al2O3. This may be

related to the selected temperatures in the test, as the δ-Al2O3 to θ-Al2O3 and α-Al2O3 transitions

may have been completed at 1000 °C.

2.2 Experimental procedure

The four tested compositions are summarized in Table 1. The introduced solid contents of

nano-alumina were chosen to be 3 wt%, which partly substitutes to α-Al2O3 micropowder. Since

solute was inevitably introduced with alumina sol, the water additions were adjusted to set the total

liquid addition to the same value. The whole components were mixed with a planetary ball mill

(QM-BP, Nanjing Nanda Instrument Plant, China) for 30 min to produce a slurry, that was poured

into plastic mould, of the size of which was Φ40×50 mm. The slurry was consolidated by the starch

thermogelation after being baked at 80 °C for 24 h. After being demoulded, the samples were dried

at 110 °C for 24 h and sintered in an electric furnace under air atmosphere to produce lightweight

alumina. The sintering cycle was heating-up at 5 °C/min from room temperature to 1780 °C, and

plateauing 3 h at 1780 °C, and then allowed to cool to room temperature at a rate of 10 °C/min.

Table 1 Initial compositions of tested lightweight aluminas (wt%)

Sample no. S0 S1 S2
α-Al2O3 micropowder 100 97 97
Nanosized α-Al2O3 - - 3
Solid content in alumina sol - 3 -
Corn starch 10 10 10
Water 40 31 40
2.3 Testing and characterization

The density of green body and linear shrinkage of the sample were determined using the

following equations:

4000m
g  , (1)
 d 2h

hs  h
Ls  100 , (2)
h

where ρg is the density of the green body (g·cm-3), m is the weight of the green body (g), d is the

diameter of the green body (mm), h and hs are respectively the height of sample before and after

sintering (mm), and Ls is the linear shrinkage induced by sintering (%).

The bulk density and open porosity of the sintered samples were determined using a

gravimetric method based on Archimedes’ Principle with water, according to ISO 5017:1998

standard. The true density was measured using an automatic true density analyser (ACCUPYC 1330,

Micromeritics Instrument Corporation, Norcross, USA) using the 325 mesh-sieved powders from

the milled samples. The respective total and closed porosities of samples were calculated from the

following equations:

t  b
 to  ×100% , (3)
t

   t  o ,
c
(4)

where πto, ρt, ρb, πo, and πc represent the total porosity, true density, bulk density, open porosity, and

closed porosity, respectively.

The pore size distribution and average pore diameter of different alumina aggregates were

measured by mercury-intrusion porosimetry measurements (AutoPore IV 9500, Micromeritics

Instrument Corporation, Norcross, USA).

The samples were bored, cut, and polished to a size of approximately 12.5 mm in diameter and

2 mm in thickness. The thermal conductivities of the samples were measured by a laser flash

thermal analyser (Flashline TM 5000, Anter Corporation, USA). Here is the simple principle of the
instrumentation for the thermal conductivity measurement. In carrying out a measurement, the

lower surface of a plane-parallel sample is first heated by a short energy pulse. The resulting

temperature change on the upper surface of the sample is then measured with an infrared detector.

The higher the sample’s thermal diffusivity, the steeper the signal increase. The specific heat of

solids can be determined using the signal height compared to the signal height of a reference

material. Using the half-time and sample thickness, the thermal diffusivity and finally the thermal

conductivity can be calculated by means of the following formulas.

s2
a  0.1388 , (5)
t1/2

1
Tmax  , (6)
m  cp

 (T)  a(T)   (T)  c p (T) , (7)

where a is the thermal diffusivity, s is the sample thickness, t1/2 is the time value at half the signal

height, ΔTmax is the signal height, m is weight of sample, cp is the specific heat, and λ is the thermal

conductivity.

The thermal diffusivities and heat capacities, with the additional knowledge of density, of the

samples were tested by the laser flash thermal analyser (Flashline TM 5000, Anter Corporation,

USA) using the comparison method. The thermal conductivities of the samples could be

automatically computed from the measured values of thermal diffusivity and specific heat capacity.

The samples were fractured with a universal compression testing machine (ZCYE-S1000A,

Jinan Zhongchuang Industry Test Systems Co., Ltd., China); any loose sample fragments were

selected and ultrasonically cleaned in deionized water for 10 min, then dried at 110 °C for 24 h.

After that, all the samples were coated with gold via sputtering at 30 mA for 3 min. The

fractography of the samples was characterized by scanning electron microscopy (SEM, Quanta 400,

FEI Company, USA).


3 Results

3.1 Microstructure

Figure 5 shows the SEM images of fracture surfaces of various samples containing different

types of nanosized alumina.

For the sample without nano-alumina, the majority of the grain sizes were small,

approximately 20 μm. The pores were uniformly distributed in the sample; however, most were

located at grain boundaries, rather than within the grains (Fig. 5 (a)–(b)).

The addition of alumina sol led to obviously different microstructures. Abnormal growth of the

grains was observed in the sample containing 3 wt% nano-alumina. The vast majority of the grain

sizes were around 25–30 μm, but some large grains could reach approximately 100–200 μm (Fig. 5

(c)). Besides some intergranular pores (around 1–30 μm), numerous intragranular pores, with a

diameter of round 0.1–1 μm, were observed within the larger grains (Fig. 5 (d)).

However, in the sample containing 3 wt% nanosized α-Al2O3, no abnormally grown grains

were observed. The introduction of nanosized α-Al2O3 resulted in a slightly increased grain size, of

approximately 25–30 μm, and the quantity of intracrystalline pores was also increased (Fig. 5 (e)–

(f)).

Thus, in brief, the introduction of different types of nano-alumina both leads to an increase in

the number of intracrystalline pores and a decrease in pore size. However, the difference lies in the

grain size; abnormal growth of the alumina grains occurs when adding alumina sol, whereas the

grain size of the sample containing nanosized α-Al2O3 trends to be homogeneous.

3.2 Pore size distribution

Table 2 Median pore diameters of various samples (μm)

Sample No. Median pore diameter (μm)


S0 1.26
S1 0.69
S2 0.73
The pore size distributions (Fig. 6) of the samples reveals that two types of nano-alumina had

similar effect on the pore size distributions of samples. The introduction of nano-alumina resulted in

a narrowing of the pore size distributions and reduced median pore sizes (Table 2). For the samples

without the addition of nano-alumina, approximately 20% of the pores were below 1 μm. For the

samples containing 3 wt% alumina sol and nanosized α-Al2O3, the ratios of pores smaller than 1 μm

were 80% and 75%, respectively. Compared to samples containing nanosized α-Al2O3, samples

with added alumina sol samples exhibited smaller median pore sizes.

3.3 Physical properties

Table 3. Physical properties of the various alumina samples

Thermal
Green body Linear Bulk True Open Closed
conductivity at
Samples density shrinkage density density porosity porosity
800 °C
(g·cm-3) (%) (g·cm-3) (g·cm-3) (%) (%)
(W·m-1·K-1)
S0 1.75±0.01 18.2±0.20 3.32±0.01 3.96±0.008 12.4±0.1 2.8±0.1 7.53±0.025
S1 1.77±0.01 16.7±0.18 3.25±0.01 3.96±0.010 10.5±0.1 7.1±0.1 6.32±0.014
S2 1.77±0.01 18.2±0.23 3.36±0.01 3.96±0.007 5.2±0.3 9.2±0.2 6.85±0.004

The physical properties of samples containing different types of nano-alumina are shown in

Table 3. To guarantee the accuracy of the experimental results, for each sample, 12 specimens were

tested in this study, and the standard errors of each experimental result are also given in Table 3.

Due to their small particle size, the nano-alumina particles tended to fill the free spaces

between α-Al2O3 micropowder particles. Hence, the density of the green body increased with the

addition of nano-alumina. As for linear shrinkage, no difference was observed for the sample

containing nanosized α-Al2O3; however, the introduction of alumina sol resulted in decreased linear

shrinkage.

Sieved powders from the milled samples used for the measurement of true density inevitably

contained some small pores. Thus, the measured true densities were slightly lower than the

theoretical density of corundum.


The introduction of two types of nano-alumina had similar effects on the porosities of the

samples. With the introduction of nano-alumina, the open porosity decreased and the closed

porosity increased. However, the introduction of nanosized α-Al2O3 had little effect on the total

porosity and bulk density of the prepared samples, whereas the sample containing alumina sol

exhibited a remarkably higher total porosity and lower bulk density than that of the sample without

addition of nano-alumina. As shown in Fig. 4 (a), the boehmite would transform in an orderly

manner into γ-Al2O3, θ-Al2O3, and finally α-Al2O3. It is reported that during the reconstructive

transformation from θ-Al2O3 to α-Al2O3, there is a specific volume reduction due to the increase in

density (from 3.6 to 3.986 g·cm-3) [38,39]. That would result in an increase in the space among

alumina particles and the formation of pore channels. This may the reason that sample S1 exhibited

a lower bulk density and higher total porosity.

The introduction of nano-alumina resulted in reduced thermal conductivity. As shown in Eq.

(8), for a porous material, the effective thermal conductivity is controlled by the thermal

conductivity of the solid phase and gas in pores and the convective heat transfer and radiative

transfer through pores.

e  f (ss , gp , rp , cp ) , (8)

where λe is the effective thermal conductivity of porous material, λss is the thermal conductivity of

solid phase, λgp is the thermal conductivity of gas in pores, λrp is the radiation heat transfer

coefficient, and λcp is the convective heat transfer coefficient.

Generally, the effective thermal conductivity of a material is related to its composition and

microstructure, especially the pore parameters. Because the compositions of the sintered lightweight

alumina samples are both α-Al2O3, the differences in the thermal conductivities of the lightweight

alumina samples are caused by the differences in their pore parameters (such as porosity, pore size,

and form of pores).

As shown in Eq. (9), the thermal conductivity of solid phase is directly related to the free path
of a phonon [40].

1
ss  CVl , (9)
3

where C is the volumetric heat capacity of phonon, V is the mean velocity of phonon, and l is the

mean free path of phonon.

The introduction of nano-alumina led to an increase in the total porosity, which would lead to

an increase in the gas–solid interface amount; the phonon scattering during the heat transfer through

the solid phase was consequently increased, which results in a decrease in thermal conductivity.

The effective thermal conductivity contributed by the pores can be described using the

following equations [41]:

dp
gp  g , (10)
lg  d p

rp  4G d pT 3 , (11)

cp  f (d p PrGrT ) , (12)

where λg is the thermal conductivity of free gas, d p is the mean diameter of the pores, lg is the

mean free path of a free gas molecule, G is a geometrical factor of the pores, ε is the emissivity, σ is

the Stefan–Boltzmann constant, T is the temperature in the pores, and Pr and Gr are the Prandtl

number and Grashof number of the gas in the pores.

As shown in Table 2, the nano-alumina addition also resulted in a smaller pore size, and the

thermal conductivity of gas in pores and the radiation heat transfer coefficient were reduced (seen in

Eqs. (10)–(11)). Moreover, with the decreasing pore size, the motion space and velocity of gas

molecules were reduced; hence, the convective heat transfer coefficient was decreased. Additionally,

the geometrical factor of the pores, G, in Eq. (10) is affected by the form of the pores in material.

The samples containing nano-alumina exhibited higher closed porosities, which are favourable to

reducing the radiative transfer.


Thus, in brief, the samples with the nano-alumina addition showed lower thermal

conductivities.

4 Discussion

In this study, different types of nano-alumina were introduced into α-Al2O3 micropowder, and

their effect on the final microstructure was investigated. In the following discussion, we attempt to

explain how the pore parameters and microstructures of the samples change with the addition of

nano-alumina. First, the surface stress of the materials with or without the addition of nano-alumina

is calculated to study its effect on the formation of closed pores. Then, the division time of closed

pores for different compositions is considered to investigate how the pore size was affected by the

introduction of nano-alumina. Finally, the grain growth of various nano-alumina is investigated to

explain the differences in the microstructures and properties of lightweight materials containing

different types of nano-alumina.

4.1 Effect of nano-alumina on the formation of closed pores

The formation and elimination of pores during sintering is strongly affected by the migration

velocities of grain boundaries and pores. Usually, in the sintering process, the migration velocity of

pores is greater than that of grain boundaries, and the migration of boundaries is restricted by the

pores. In this case, after heat treatment, the pores are eliminated or opened. Closed pores would be

formed only when the grain boundary velocity is greater than the pore velocity; in this case, the

grain boundaries are liberated from the restriction of the pores and move freely. Thus, the pores,

before being eliminated, could be closed by the moving grain boundaries, thus forming closed pores.

Because the migration of grain boundaries is mainly driven by the surface stress among particles,

the surface stresses of materials with or without added nano-alumina is calculated in the following

section.

As reported in reference [36], the surface stress in the neck between two semi-liquid particles

can be calculated by the following equation:


1  2
P  , (13)
1 2

where P is the surface stress, γ1 and γ2 are the surface energies, and ρ1 and ρ2 are the neck curvature

radii of each particle.

The curvature radius of the neck (ρ) is related to the particle radius (r) and the central angle of

the neck (θ) [36]:


  r[sin 1 ( )  1] . (14)
2

The surface energy of nano-alumina and that of α-Al2O3 micropowder follow the relationship

 n rm
 , (15)
 m rn

where γn and γm are respectively the surface energies and rn and rm are the radii of

nano-alumina and α-Al2O3 micropowder, respectively.

Therefore, according to Eqs. (13)–(15), the surface stress in the neck between two α-Al2O3

micropowder particles and in the neck between one α-Al2O3 micropowder particle and one

nano-alumina particle would follow the relationship:

P2 rn2  rm2
 , (16)
P1 2rn2

where P1 is the surface stress in the neck between two α-Al2O3 micropowder particles (Pa), and P2

is the surface stress in the neck between one α-Al2O3 micropowder particle and one nano-alumina

particle (Pa).

Introducing the particle sizes of nano- and microsized alumina raw materials into Eq. (16), it

follows that the nano/micro double scales provide a larger surface stress, which is 3.096×103 times

that in the sintering of the sole micropowder.

As shown above, with the introduction of nano-alumina, larger surface stresses are created in

local areas containing nano-alumina that accelerate the grain-boundary migration. The further rapid

migration of grain boundaries tends to increase the grain size and trap some intergranular pores
within the larger grains. Therefore, the samples containing nano-alumina exhibited a higher closed

porosity and lower open porosity (Table 3).

4.2 Effect of nano-alumina on the division of closed pores

In addition to promoting the creation of a closed porosity, the addition of nano-alumina tends

to decrease the average size of pores. During the sintering process, some large unstable pores would

be divided into stable ones with a smaller size, owing the mass diffusion and transfer [42–44].

Hence, the effect of nano-alumina addition on pore division will be discussed in the following

section.

During the sintering process, mass diffusion and transfer result in changes in the pore shapes.

The pore shape evolution is a consequence of energy reduction, which tends to transform

thermodynamically unstable shapes into stable ones. The pore shape change is considered to be

controlled by surface diffusion, and surface-attachment-limited kinetics was widely used to describe

the pore shape evolution [42–44]. Among the various shapes of pores evolving during sintering,

spherical shapes were proven to be the most stable, and cylindrical ones were the most unstable.

Gupta et al. [45] reported that cylindrical voids in alumina were unstable and could undergo various

configurational changes during annealing prior to their ultimate breakup into a configuration of

rows of spherical voids.

A dynamic model was developed to describe the evolution of cylindrical pores into spheres by

Stüwe et al. [46]. A similar occurrence was observed in our work, as illustrated by the SEM image

represented in Fig. 7. The arrangement of intracrystalline pores was accorded in a row. Therefore,

the formation of these intracrystalline spherical pores is assumed to be a consequence of the

intragranular trapping of cylindrical pores by rapid grain-boundary displacement, resulting in their

further disintegrations into small spheres arranged in rows. However, at a larger scale (i.e., within a

whole grain), these local arrangements in rows appeared randomly distributed, probably because of

the random distribution of the original cylindrical pores. Another reason could be that all cylindrical

pores have not been trapped at the same time or by a single grain boundary.
Therefore, in order to explain how the pore size decreases with the introduction of

nano-alumina, the division time of cylindrical closed pores in the samples with or without

nano-alumina addition was calculated in the following section based on Stüwe’s model.

According to reference [46], the time required for the division of the cylindrical closed pores, τ,

is given by the equation:

kT  4 1 
  2 ( ) ln , (17)
Ds  Z A0

where v and ξ are respectively the mean diffusion path factor and the diffusing atoms number factor,

which can be roughly estimated as 1/3 [46]; k is the Boltzmann constant; T is the temperature; Ds is

the surface diffusion coefficient of alumina; γ is the surface energy of alumina; ρ is the radius of the

cylindrical pore; ψ is the diameter of alumina molecule; Z is a calculation parameter, can be roughly

estimated as 1/(16π2) [46]; and A0 is the original fluctuation amplitude, which is equal to ψ.

From reference [47], the ratio between the surface diffusion coefficients of two powders with R

and R’ radii can be expressed as

Ds ' R 2 ( R ' + C )
 , (18)
Ds R '2 ( R  C )

where λ is the average free path of gas molecules and C is the Cunningham correction factor, equal

to 1.26.

The surface diffusion coefficient of α-Al2O3 micropowder is provided by reference [43], and

the surface diffusion coefficient of nano-alumina can be calculated with Eq. (18).

The parameters required for calculation are listed in Table 4. Introducing these parameters into

equation (17), the time that is necessary to divide cylindrical pores with or without nano-alumina

addition can be obtained. Without the addition of nano-alumina, the division time is approximately
1.58×104 hours, which is impossible to reach practically. In contrast, only 1.18 hours are necessary

to divide the pores in nano-alumina-added materials. That is to say, the introduction of

nano-alumina lead to the decreased division time of pores; the large pores can be divided rapidly

into smaller ones, which results in a reduction in pore size.

Table 4 Set of parameters used for calculation

Parameters Values
Mean diffusion path factor, v 1/3 [46]
Diffusing atoms number factor, ξ 1/3 [46]
Boltzmann constant, k (J∙K-1) 1.23×10-23
Temperature, T (K) 2053
Radius of the cylindrical pore, ρ (m) 10-7 [46]
Diameter of alumina molecule, ψ (m) 2×10-10 [48]
Calculation parameter, Z 1/(16π2) [46]
Original fluctuation amplitude, A0 (m) 2×10-10 [46]
Surface diffusion coefficient of α-Al2O3 micropowder, Ds (m2·s-1) 10-14 [47]

Surface diffusion coefficient of nano-alumina, Ds (m2·s-1) 8.518×10-13 [47]
Surface energy of α-Al2O3 micropowder, γ (N·m-1) 0.678 [36]
Surface energy of nano-alumina, γ’ (N·m-1) 53.359

4.3 Effect of various types of nano-alumina on the sintering behaviour of lightweight

alumina

Due to its small particle size, the introduced nano-alumina would fill in the pores between

α-Al2O3 micropowder particles and form an accumulation of nano/micro double scales. As

discussed above, on the one hand, the nano/micro double-scale effect induces a greater surface

stress, which can activate the diffusion and increase the grain boundary migration velocities. The

pores can be closed before being eliminated and transform into intracrystalline pores. On the other

hand, the introduction of nano-alumina can shorten the division time of closed pores. The large

intracrystalline pores can be divided rapidly into smaller ones, which results in a reduction in pore

size. Hence, the introduction of the appropriate amount of nano-alumina would result in an

increased closed porosity and decreased pore size of lightweight alumina.

However, the samples containing different types of nano-alumina still exhibited some
differences in the microstructures, especially in the grain size and amount of intracrystalline pores.

Therefore, we attempt to explain the causes of the microstructural differences in the following

section.

In this study, nano-alumina was introduced to increase the grain-boundary migration. As

shown in Eqs. (16)–(18), in the samples with the addition of nano-alumina, the surface stress and

division time of closed pores are closely related to the grain size of the introduced nano-alumina.

During sintering, the grain growth would inevitably occur, resulting in an increased average grain

size. However, a nano-alumina type that exhibits a slower grain growth rate may act for a longer

time and lead to a final material with better properties. Therefore, in order to compare the growth

rate of different types of nano-alumina, the microstructures of samples prepared by different

nano-alumina powders were observed to explore the reason for the differences in the properties of

samples containing different types of nano-alumina.

Considering that alumina sol, in a liquid state, is not easy to be shaped to a solid sample,

nanosized γ-Al2O3 was used to investigated the grain growth of alumina sol, since the alumina sol

would transform into γ-Al2O3 at 400 °C (Fig. 4(a)).

Nanosized α-Al2O3 (with the same information as the materials mentioned in Section 2.1) and

γ-Al2O3 (mean particle diameter=26.0 nm, specific surface area= 276.49 m2·g-1, Shanghai Macklin

Biochemical Co., Ltd., China) were selected as raw materials in this study. The particle size

distribution and TEM image of nanosized γ-Al2O3 are given in Fig. 8.

As a binding agent, 0.8 wt% polyvinyl alcohol was mixed with the nano-alumina powders

using a planetary ball mill (QM-BP, Nanjing Nanda Instrument Plant, China), rotated at 365 rpm for

10 min. The powders were then pressed into green samples under a uniaxial pressure of 150 MPa.

After drying at 100 °C for 24 hours, the green samples were heated in an electric furnace under air

atmosphere. The sintering cycle was heated at a rate of 2 °C/min from room temperature to 550 °C,

at 8 °C/min from 550 °C to 1300 °C, and plateaued for 1 h at 1300 °C, and then was allowed to cool
to room temperature at a rate of 10 °C/min.

The fracture surfaces of different nano-alumina samples are given in Fig. 9. The grain sizes of

α-Al2O3 were significantly larger than those of γ-Al2O3 sample. The activation energy of

thermodynamically stable α-Al2O3, especially when nanosized, is rather low, and results in a

relatively low sintering temperature and facilitates grain growth [49]. However, in the case of

γ-Al2O3 starting powder, it does not promote the sintering process until the γ–α phase

transformation is fully accomplished [50–52]. Therefore, γ-Al2O3 powder exhibits a low sintering

rate at temperatures lower than 1200 °C [52]. In brief, compared to α-Al2O3, the α-Al2O3 derived

from γ-Al2O3 participated in the sintering process later, limiting the grain growth.

As shown in Eqs. (16)–(18), in the samples with the addition of nano-alumina, the surface

stress and division time of closed pores are closely related to the grain size of the introduced

nano-alumina. If the grain size of nano-alumina remains in the nanoscale, due to the large surface

stress, the grain boundaries possess a strong driving force to move across the pores, leading to a

large grain with intracrystalline pores inside. However, with sintering, the grain growth of

nano-alumina would inevitably occur, resulting in increased grain size. The surface stress, as well as

the driving force of grain-boundary migration, would be reduced, and the grain sizes reach their

limits.

For the samples with the addition of nanosized α-Al2O3, the grain growth of nano-alumina is

more likely to occur, which leads to a decreased surface stress and increased division time of closed

pores. Hence, those samples exhibit small grain sizes and large pore sizes. By comparison, the

alumina sol shows a slower grain growth rate, and the large surface stress and short division time

could remain for a longer time; thus, abnormally grown grains with numerous smaller

intracrystalline pores inside are observed.

5 Conclusions

(1) By adding nano-alumina into α-Al2O3 micropowder, an accumulation of nano/micro double


scales was formed. Compared to the sintering of pure micropowder, the introduction of

nano-alumina provides a larger surface stress and shortens the division tome of pores. During the

sintering process, the large surface stress would activate the superplasticity of nano-alumina, the

boundary migration velocities are increased, and the pores are consequently closed inside of the

grains. In addition, the large intercrystalline pores can be divided rapidly into smaller ones, which

results in a reduction in pore size.

(2) Two types of nano-alumina (nanosized α-Al2O3 and alumina sol) have a similar effect on

the properties of lightweight alumina. The introduction of nano-alumina leads to an increase in

closed porosities of the samples, a decrease in the bulk densities and open porosities, a reduction in

pore sizes and thermal conductivities, and an increase in the amount of intracrystalline pores.

(3) The grain growth of nano-alumina is more likely to occur when adding nanosized α-Al2O3;

hence, the samples exhibit small grain sizes and large pore sizes. By comparison, the alumina sol

shows a slower grain growth rate, and its large surface stress and short division time could remain

for a longer time; thus, abnormally grown grains with numerous smaller intracrystalline pores inside

are observed.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China

(Grant No. 51474165 and 51772222), the China Postdoctoral Science Foundation (Grant No.

2017M622535), the International Science & Technology Cooperation Program of Hubei Province

of China (Grant No. 2016AHB026), the Fund for the Excellent Creative Research Groups of Higher

Education of Hubei Province of China (Grant No. T201602), and the Foreign Experts Project of the

State Administration of Foreign Experts Affairs of China (Grant no. GDW20174200160).

References
[1] L.P. Fu, H.Z. Gu, A. Huang, M.J. Zhang, X.Q. Hong, L.W. Jin, Possible improvements of

alumina-magnesia castable by lightweight microporous aggregates, Ceram. Int. 41 (2015)

1263-1270.
[2] L.P. Fu, H.Z. Gu, A. Huang, M.J. Zhang, Z.K. Li, Slag resistance mechanism of lightweight

microporous corundum aggregate, J. Am. Ceram. Soc. 98 (2015) 1658-1663.

[3] L.P. Fu, A. Huang, P.F. Lian, H.Z. Gu, Isolation or corrosion of microporous alumina in contact

with various CaO-Al2O3-SiO2 slags, Corros. Sci. 120 (2017) 211-218.

[4] H. Nishijima, R. Maki, Y. Suzuki, Microstructural control of porous Al2TiO5 by using potato

starch as pore-forming agent, J. Ceram. Soc. Jpn. 121 (2013) 730-733.

[5] R. Liu, C.A. Wang, Effects of mono-dispersed PMMA micro-balls as pore-forming agent on the

properties of porous YSZ ceramics, J. Eur. Ceram. Soc. 33 (2013) 1859–1865.

[6] X.W. Yin, X.M. Li, L.T. Zhang, L.F. Cheng, Y.S. Liu, T.H. Pan, Microstructure and mechanical

properties of Lu2O3-doped porous silicon nitride ceramics using phenolic resin as pore-forming

agent, Int. J. Appl. Ceram. Tec. 7 (2010) 391–398.

[7] F. Chen, Q. Shen, F.Q. Yan, L.M. Zhang, Pressureless sintering of α-Si3N4 porous ceramics

using a H3PO4 pore-forming agent, J. Am. Ceram. Soc. 90 (2007) 2379–2383.

[8] Z.Y. Deng, T. Fukasawa, M. Ando, G.J. Zhang, T. Ohji, High-surface-area alumina ceramics

fabricated by the decomposition of Al(OH)3, J. Am. Ceram. Soc. 84 (2004) 485-491.

[9] S.J. Li, N. Li, Effects of composition and temperature on porosity and pore size distribution of

porous ceramics prepared from Al(OH)3 and kaolinite gangue, Ceram. Int. 33 (2007) 551-556.

[10] G.G. Xu, H.L. Zhao, H.Z. Cui, Z.H. Zhang, J. Li, Stability, microstructure and mechanical

properties of (Al,Fe)2TiO5 porous ceramic reinforced by in-situ mullite, J. Ceram. Soc. Jpn. 123

(2015) 156-159.

[11] P. Barreiro, P. Rey, A. Souto, F Guitián, Porous stabilized zirconia coatings on zircon using

volatility diagrams, J. Eur. Ceram. Soc. 29 (2009) 653-659.

[12] P. Sepulveda, J.G.P. Binner, Processing of cellular ceramics by foaming and in situ

polymerization of organic monomers, J. Eur. Ceram. Soc. 19 (1999) 2059-2066.

[13] J.G. Park, A. Pokhrel, S.D. Nam, W. Zhao, S.M. Park, I.J. Kim, Self-setting wet foams to

porous ceramics by direct forming, J. Ceram. Process. Res. 14 (2013) 508-512.


[14] W. Cao, X.D. Cheng, L.L. Gong, L. Ye, R.F. Zhang, H.P. Zhang, Thermal conductivity of

highly porous ceramic foams with different agar concentrations, Mater. Lett. 139 (2015) 66-69.

[15] S. Bhaskar, G.H. Cho, J.G. Park, S.W. Kim, H.T. Kim, I.J.Kim, Micro porous SiO2-SiC

ceramics from particle stabilized foams by direct foaming, J. Ceram. Soc. Jpn. 123 (2015) 378-382.

[16] N. Sarkar, J.G. Park, S. Mazumder, A. Pokhrel, C.G. Aneziris, I.J. Kim, Al2TiO5–mullite

porous ceramics from particle stabilized wet foam, Ceram. Int. 15 (2015) 6306-6311.

[17] W. Li, K. Lu, J.Y. Walz, Effects of solids loading on sintering and properties of freeze-cast

kaolinite–silica porous composites, J. Am. Ceram. Soc. 96 (2013) 1763–1771.

[18] X.B. Xu, H. He, Y.P. Zhang, D.F. Zhang, Z.D. Yang, Influence of position on the

microstructure of carbon black/polyvinyl alcohol composite obtained by the directional

freeze-drying process, J. Macromol. Sci. B. 53 (2014) 568-574.

[19] C. Tallóna, R. Moreno, I. Nieto, Shaping of porous alumina bodies by freeze casting, Adv.

Appl. Ceram. 108 (2009) 307-313.

[20] C. Pekor, I. Nettleship, The effect of the molecular weight of polyethylene glycol on the

microstructure of freeze-cast alumina, Ceram. Int. 40 (2014) 9171-9177.

[21] M. Fukushima, Y.I. Yoshizawa, T. Ohji, Macroporous ceramics by gelation–freezing route

using gelatin, Adv. Eng. Mater. 16 (2014) 607–620.

[22] G. Jean, V. Sciamanna, M. Demuynck, F. Cambier, M. Gonon, Macroporous ceramics: novel

route using partial sintering of alumina-powder agglomerates obtained by spray-drying, Ceram. Int.

40 (2014) 10197-10203.

[23] T. Isobe, A. Ooyama, S. Mai, A. Nakajima, Pore size control of Al2O3 ceramics using two-step

sintering, Ceram. Inter. 38 (2012) 787–793.

[24] A. Kalemtas, G. Topates, H. Özcoban, H. Mandal, F. Kara, R. Janssen, Mechanical

characterization of highly porous β-Si3N4 ceramics fabricated via partial sintering & starch addition,

J. Eur. Ceram. Soc. 33 (2012) 1507-1515.

[25] J. Yang, J.F. Yang, S.Y. Shan, J.Q. Gao, T. Ohji, Effect of sintering additives on microstructure
and mechanical properties of porous silicon nitride ceramics, J. Am. Ceram. Soc. 89 (2006)

3843-3845.

[26] S. Ding, S. Zhu, Y.P. Zeng, D. Jiang, Fabrication of mullite-bonded porous silicon carbide

ceramics by in situ reaction bonding, J. Eur. Ceram. Soc. 27 (2007) 2095-2102.

[27] R.Z. Liu, G.W. Liu, J.F. Yang, H.Y. Jin, Y. Lu, W.W. Gu, Effect of removal of silicon on

preparation of porous SiC ceramics following reaction bonding and recrystallization, Mater. Sci.

Eng. A. 552 (2012) 9-14.

[28] C.Y. Bai, Y. Li, Z.M. Liu, P.W. Liu, X.Y. Deng, J.B. Li, J. Yang, Fabrication and properties of

mullite-bonded porous SiC membrane supports using bauxite as aluminum source, Ceram. Inter. 41

(2015) 4391-4400.

[29] L.P. Fu, H.Z. Gu, A. Huang, H.W. Ni, Correlations among processing parameters and porosity

of a lightweight alumina, Ceram. Int. 44 (2018) 14076-14081.

[30] I. Maclaren, R.M. Cannon, M.A. Gülgün, R. Voytovych, N.P. Pogrion, C. Scheu, U. Täffner, M.

Rühle, Abnormal grain growth in alumina: synergistic effects of yttria and silica, J. Am. Ceram. Soc.

86 (2003) 650–659.

[31] Y.M. Kim, S.H. Hong, D.Y. Kim, Anisotropic abnormal grain growth in TiO2/SiO2-doped

alumina, J. Am. Ceram. Soc. 83 (2000) 2809–2812.

[32] A. Kebbede, J. Parai, A.H. Carim, Anisotropic grain growth in α-Al2O3 with SiO2 and TiO2

additions, J. Am. Ceram. Soc. 83 (2000) 2845–2851.

[33] S.H. Hong, D.Y. Kim, Effect of liquid content on the abnormal grain growth of alumina, J. Am.

Ceram. Soc. 84 (2001) 1597–1600.

[34] B.N. Kim, K. Hiraga, K. Morita, Y. Sakka, Superplasticity in alumina enhanced by

co-dispersion of 10% zirconia and 10% spinel particles, Acta. Mater. 49 (2001) 887–895.

[35] X.Z. Zhou, D.M. Hulbert, J.D. Kuntz, R.K. Sadangai, Superplasticity of zirconia–alumina–

spinel nanoceramic composite by spark plasma sintering of plasma sprayed powders, Mater. Sci.

Eng. A. 394 (2005) 353–359.


[36] L.P. Fu, A. Huang, H.Z. Gu, D.H. Lu, P.F. Lian, Effect of nano-alumina sol on the sintering

properties and microstructure of microporous corundum, Mater. Design. 89 (2016) 21-26.

[37] R.S. Zhou, R.L. Snyder, Structures and transformation mechanisms of the η, γ and θ transition

aluminas, Acta. Crystallogr. 47 (1991) 617-630.

[38] S. Jagota, R. Raj, Model for the crystallization and sintering of unseeded and seeded boehmite

gels, J. Mater. Sci. 27 (1992) 2551-2557.

[39] H. Chen, L. Zhao, X. He, W. Fang, Z.X. Lei, H. Chen, The fabrication of porous corundum

spheres with core-shell structure for corundum-spinel castables, Mater. Des. 85 (2015) 574-581.

[40] K.T. Regner, D.P. Sellan, Z.H. Su, C.H. Amon, A.J.H. McGaughey, J.A. Malen, Broadband

phonon mean free path contributions to thermal conductivity measured using frequency domain

thermoreflectance, Nat. Commun. 4 (2013) 1640.

[41] W. Pabst, E. Gregorová, G. Tichá, Elasticity of porous ceramics—A critical study of

modulus—porosity relations, J. Eur. Ceram. Soc. 26 (2006) 1085-1097.

[42] M. Kitayama, A. M. Glaeser, The kinetics of pore shape evolution in alumina, J. Mater. Synth.

Proces, 6 (1998) 161-167.

[43] M. Kitayama, T. Narushima, W. C. Carter, R.M. Cannon, A. M. Glaeser, The wulff shape of

alumina: I, modeling the kinetics of morphological evolution, J. Am. Ceram. Soc. 83 (2000) 2561–

2571.

[44] M. Kitayama, T. Narushima, A. M. Glaeser, The wulff shape of alumina: II, experimental

measurements of pore shape evolution rates, J. Am. Ceram. Soc. 83 (2000) 2572–2583.

[45] T. K. Gupta, Instability of cylindrical voids in alumina, J. Am. Ceram. Soc. 61 (2010) 191-195.

[46] H.P. Stüwe, O. Kolednik, Shape instability of thin cylinders, Acta. Metall. 36 (1988)

1705-1708.

[47] V.V. Karasev, A.A. Onischuk, O.G. Glotov, A.M. Baklanov, A.G. Maryasov, V.E. Zarrko, V.N.

Panfilov, A.I. Levykin, K.K. Sabelfeld, Formation of charged aggregates of Al2O3, nanoparticles by

combustion of aluminum droplets in air, Combust. Flame. 138 (2004) 40-54.


[48] R.J. Brook, Pore-grain boundary interactions and grain growth, J. Am. Ceram. Soc. 52 (1969)

56–57.

[49] W.M. Zeng, L. Gao, L.H. Gui, J.K. Guo, Sintering kinetics of α-Al2O3 powder, Ceram. Int. 25

(1999) 723–726.

[50] Z.Q. Sun, B.Q. Li, P. Hu, F. Ding, F.L. Yuan, Alumina ceramics with uniform grains prepared

from Al2O3 nanospheres, J. Alloy. Compd. 688 (2016) 933–938.

[51] T.C. Chou, T.G. Nieh, Nucleation and concurrent anomalous grain growth of α-Al2O3 during

γ-α phase transformation, J. Am. Ceram. Soc. 74 (1991) 2270–2279

[52] R.B. Bagwell, G.L. Messing, Effect of seeding and water vapor on the nucleation and growth

of α-Al2O3 from γ-Al2O3, J. Am. Ceram. Soc. 82 (1999) 825–832.


Fig. 1 Particle-size distributions of raw materials

Fig. 2 TEM images of (a) nanosized α-Al2O3 and (b) alumina sol

Fig. 3 XRD patterns of nanosized α-Al2O3 and α-Al2O3 micropowder

Fig. 4 (a) XRD patterns and (b) TG/ DSC curves of dried power during the heat treatment

Fig. 5 SEM images of various samples: (a–b) without nanosized alumina; (c–d) containing 3 wt%

addition of alumina sol; (e–f) containing wt% addition of nanosized α-Al2O3

Fig. 6 Pore size distribution expressed as (a) discrete frequency distribution and (b) cumulative size

distribution of various samples

Fig. 7 Straight line-like distribution of intracrystalline pores

Fig. 8 (a) Particle size distribution and (b) TEM image of nanosized γ-Al2O3

Fig. 9 SEM images of samples with different raw materials: (a) nanosized α-Al2O3; (b) nanosized

γ-Al2O3

You might also like