Anomalous Reduction in Surface Area During Mechanical Activation of Boehmite

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/241076094

Anomalous reduction in surface area during mechanical activation of


boehmite synthesized by thermal decomposition of gibbsite

Article  in  Powder Technology · March 2011


DOI: 10.1016/j.powtec.2010.12.010

CITATIONS READS

19 251

4 authors, including:

T.C. Alex Rakesh Kumar


National Metallurgical Laboratory National Metallurgical Laboratory
57 PUBLICATIONS   657 CITATIONS    110 PUBLICATIONS   1,803 CITATIONS   

SEE PROFILE SEE PROFILE

Sanat K Roy
Indian Institute of Technology Kharagpur
134 PUBLICATIONS   1,820 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Utilization of Iron & Steel Plant Wastes View project

Special topics in mechanochemistry and mechanical activation of solids View project

All content following this page was uploaded by Rakesh Kumar on 25 January 2018.

The user has requested enhancement of the downloaded file.


Powder Technology 208 (2011) 128–136

Contents lists available at ScienceDirect

Powder Technology
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p ow t e c

Anomalous reduction in surface area during mechanical activation of boehmite


synthesized by thermal decomposition of gibbsite
T.C. Alex a,⁎, Rakesh Kumar a, S.K. Roy b, S.P. Mehrotra c
a
National Metallurgical Laboratory, Jamshedpur-831 007, India
b
Indian Institute of Technology, Kharagpur-721 302, India
c
Indian Institute of Technology, Kanpur-208016, India

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

Article history: Mechanical activation of boehmite (γ-AlOOH), synthesized by thermal decomposition of gibbsite, has been
Received 15 April 2010 carried out in a planetary mill up to 240 min. After an initial decrease in particle size up to 15 min, the particle
Received in revised form 27 November 2010 size shows an increase with further milling; the median size (d50) has increased from 1.8 to 5 μm during 15 to
Accepted 11 December 2010
240 min of milling. Quite unexpectedly, the BET specific surface area of the sample (N2 adsorption method)
Available online 21 December 2010
decreases continuously from 264 m2/g to 67 m2/g with milling. A detailed analysis of N2 adsorption/
Keywords:
desorption isotherms has indicated that the decrease in surface area is associated with: (a) change in narrow
Boehmite slit like pores with microporosity to slit shaped pores originating from loose aggregate of platelet type
Mechanical activation particles; and (b) shift of maxima in pore size distribution plot at ~ 2 nm and ~ 4 nm to dominantly ~ 23 nm
Pore size distribution size pores. Scanning electron microscopy (SEM) studies have revealed that during milling, initial breakage is
Amorphisation followed by agglomeration/fusion of particles with consequent loss in porosity. Amorphisation, decrease in
Microstrain microcrystallite dimension (MCD) and increase in microstrain (ε) are indicated from a detailed analysis of X-
Reactivity ray powder diffraction patterns and Fourier Transform Infrared (FTIR) spectra. Reactivity of samples,
expressed in terms of increase in dissolution in alkali (in 8 M NaOH at 90 °C) and decrease in boehmite to γ-
Al2O3 transformation temperature, increases with milling time. The nature of correlations between reactivity
and physico-chemical changes during milling has been analyzed and discussed.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction peratures of boehmite to alumina phases. However, no results were


presented on particle size reduction and structural changes. Boehmite
Boehmite is a layered structure Al-monohydroxide (γ-AlOOH) was planetary ball milled for 60 min by MacKenzie et al. [6] and the
phase. Its structure consists of interlocking double layers of edge shared changes in structure and thermal reactions were monitored by X-ray
AlO6 octahedra that are stacked along [010] direction and connected powder diffraction (XRD), thermal analysis (TG/DTA) and 27Al MAS
with each other by hydrogen bond [1–3]. Reactivity of boehmite is of NMR. In the published studies, boehmite used was either hydrother-
interest during alkali leaching of boehmitic bauxites since it requires mally synthesized [8] or obtained from chemical suppliers without
higher temperature and alkali concentration compared to its tri-hydrate any information on its method of synthesis [5,6]. Kano et al. [5] and
counterpart, gibbsite [Al(OH)3] [4]. Boehmite is used extensively as a MacKenzie et al. [6] did not report any particle size data. In general,
precursor to prepare transition alumina through thermal transforma- mechanical activation results in a decrease in particle size (or increase
tion. Mechanical activation of boehmite has been studied to alter its in surface area) [9,10] as reported by Tsuchida and Horigome [8] and
reactivity and, consequently, reduce the thermal transformation Yong and Wang [11]. In our study on planetary milling of boehmite,
temperatures [5–8]. Tsuchida and Horigome [8] reported that during synthesized by thermal decomposition of gibbsite, an increase in
planetary milling of hydrothermally synthesized boehmite, median particle size and a decrease in the BET surface area have been observed.
particle size was reduced steeply during the first 30 min and it remained The synthesized boehmite was unique in terms of very high surface
unchanged during further milling up to 8 h. It was also reported that area (264 m2/g) which was attributed to channels of porosity formed
intensity of X-ray peaks in diffraction patterns decreased monoto- during the synthesis process. The anomalous change in the surface
nously except for (020) reflection. Kano et al. [5] found that the use of area has prompted a detailed study of the mechanical activation of
α-alumina as a seed during milling decreased transformation tem- the sample and the results on its characterization and reactivity are
reported in this paper. The emphasis is on characterization in terms of
particle size, specific surface area and porosity distribution, morpho-
⁎ Corresponding author. Tel.: +91 657 2345042; fax: +91 657 2345213. logical changes and structural changes. Reactivity is measured in terms
E-mail address: tc_alex@yahoo.com (T.C. Alex). of changes in thermal transformation temperatures and dissolution

0032-5910/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.powtec.2010.12.010
T.C. Alex et al. / Powder Technology 208 (2011) 128–136 129

behaviour. Attempt has also been made to correlate reactivity with tion. The full widths at half maxima (FWHM) were determined using
physico-chemical characteristics of mechanically activated samples. XRDA software [15]. The instrumental broadening was corrected
using quartz as an external standard. The corrected width (β) was
2. Materials and methods used to determine average micro-crystallite dimension (MCD) and
microstrain (ε) using Williamson–Hall equation [16,17]:
2.1. Synthesis of boehmite
0:9λ
The boehmite used in the study was prepared by thermal decom- β⋅ cos θ = + 4ε ⋅ sin θ ð2Þ
MCD
position of a coarse grained (d50 = 120 μm) gibbsite. The gibbsite
sample was supplied by M/s National Aluminium Company (NALCO),
Bhubaneswar, India. Thermal decomposition of gibbsite was carried MCD and ε were obtained from the intercept and slope of straight
at 350 °C for 2 h in a muffle furnace under static air atmosphere to line fitted to β cosθ vs. sinθ plot for eight strongest peaks. θ and λ
obtain phase pure boehmite. The temperature and time of decompo- correspond to the Bragg angle and wavelength of the X-ray radiation
sition were determined based on exploratory studies involving TG/DTA used.
studies, heating experiments and phase characterization. Fourier Transform Infrared (FTIR) spectra of the boehmite
samples were recorded on a Nicolet 5700 Spectrometer in the region
2.2. Mechanical activation experiments

Mechanical milling of boehmite was carried out in a batch type


planetary mill (Pulverisette P6, Fritsch GmbH, Germany) at ambient
conditions. The amount of sample taken was 30 g. The ball to powder
ratio was 10:1 and a rotational speed of 400 rpm was employed in all
the experiments. The direction of rotation was reversed at every
5 min. To ensure uniform grinding and avoid overheating of the mill
and the charge, the mill was stopped at every 15 min. The material
sticking to the walls of the grinding bowl was scrapped off and mixed
thoroughly before resuming the milling operation. The milling was
continued up to 240 min. The samples milled for different times,
ranging from 0 to 240 min, were used for characterization.

2.3. Characterization techniques

Particle size distribution was determined using a laser diffraction


particle size analyzer (Model: Mastersizer, Malvern, U.K.). To ensure
proper dispersion, sodium hexa metaphosphate was used as a dis-
persant and the sample was dispersed in water for 2 min using an
ultrasonic bath.
N2 adsorption method was used to determine the BET surface area
and pore size distribution. A Micromeritics (USA) surface area analyser
(Model: ASAP2020) was used for this purpose. High purity N2 (XL
Grade, BOC, India) was used as adsorbent. The adsorption–desorption
isotherms were recorded at liquid N2 temperature (77 K) after de-
gassing the sample. After the initial degassing at 90 °C, the temperature
was raised to 200 °C at a heating rate of 10 °C/min and continued
the degassing for 5 h under vacuum. The amount of N2 adsorbed (na) at
liquid N2 temperatures at various relative pressure (p/p0) was recorded.
The adsorption data in the relative pressure (p/p0) range of 0.05–0.3
was used to determine monolayer capacity (nam) using the BET
equation [12,13]

p 1 ðC−1Þ p
= a + a ð1Þ
na ⋅ðp0 −pÞ nm ⋅C nm ⋅C ⋅ p0

(where C is a constant which is a measure of the adsorbent–adsorbate


interaction). The specific surface area was calculated from nam. The
software feature of the equipment was used to determine pore size
distribution based on Barret–Joyner–Halenda (BJH) method [14].
The morphological features of the milled boehmite samples were
examined using a scanning electron microscope (SEM) (Model: 840A,
JEOL, Japan). The samples were coated with gold to avoid charging.
Powder X-ray diffraction (XRD) patterns were recorded in the 2θ
range of 10–80° with a Siemens diffractometer (Model: D500) using
CuKα radiation at a scan rate of 0.5°/min. Peak intensity was measured
from the area under the peak after making background correction. The
ratio of peak intensity (Ihkl) of mechanically activated sample and that Fig. 1. SEM micrographs of boehmite showing (a) uniform size, (b) agglomerate
of un-milled sample was used as a measure of degree of amorphisa- structure, and (c) slit like pores.
130 T.C. Alex et al. / Powder Technology 208 (2011) 128–136

of 4000–400 cm− 1. The samples were mixed with KBr and pressed in
the form of pellets to record the spectra.

2.4. Evaluation of reactivity

Reactivity of the milled samples was evaluated in terms of: (a) its
thermal transformation temperature to γ-Al2O3; and (b) leachability
in alkali under specified condition. A simultaneous TG–DTA appara-
tus (model: SDT Q600, TA Instruments, USA) was used to study the
thermal transformations. Thermal analysis of the milled samples was
carried out at heating rate of 10 °C/min and under a constant flow of
high purity argon gas. The DTA peak temperature corresponding to
boehmite ➝ γ-Al2O3 transition was taken as a measure of reactivity.
Lower the temperature of transformation, higher will be the reac-
tivity. Alkali leaching tests were carried out under following con-
dition: temperature—90 °C, NaOH concentration—8 M, duration—30
min and stirring speed—300 rpm. Reactivity in this case was
expressed in terms of percentage dissolution.
Fig. 2. Size distribution and characteristic diameters of boehmite milled for different
durations.

2.5. Correlations
3.2. Mechanical milling and characterization
Binary correlation (rxy) between reactivity (y) and physicochem-
ical characteristics of the samples (x), namely, median diameter (d50), 3.2.1. Particle size distribution
geometrical specific surface area (SSAgeo), BET surface area (SSABET), The particle size distribution and characteristic diameters, d90, d50
microcrystalline dimension (MCD) and microstrain (ε) were calcu- d10 of the milled and original samples are given in Fig. 2. The d90, d50,
lated to understand the nature of correlation. The correlation was and d10 values for the original sample were 153.4, 110.2 and 77.0 μm
calculated by dividing covariance between x–y with variance x and y respectively. The variation of characteristic diameters with time
[18]. indicates that the particle size decreases during initial stages of
milling (~15 min) and with further milling an increase, especially in
d90 and d50, is observed. Between 15 and 240 min, d50 shows an
3. Results and discussion increase from 1.8 μm to 5 μm and d90 increases from ~30 μm to 42 μm.
These results are in sharp contrast with the particle size results
3.1. Boehmite reported by Tsuchida and Horigome [8] which showed an initial
decrease in particle size followed by a steady size. As reported by
SEM micrographs in Fig. 1 show the morphology of the boehmite Juhasz and Opoczky [20,21], it is likely that the fine particles undergo
synthesized after thermal decomposition of gibbsite. The boehmite aggregation and agglomeration during prolonged milling which leads
particles were of uniform size (Fig. 1(a)) and preserved the agglomerate to an increase in particle size.
structure (Fig. 1(b)) of the precursor gibbsite. The particles showed
large number of slit shaped pores (Fig. 1(c)). These slit shaped pores 3.2.2. Surface area and porosity
are unique to the sample. The pores are developed during thermal The BET surface area of the boehmite samples obtained after
decomposition as a result of water removal preferentially along the different durations of milling is given in Fig. 3. Fig. 3 also shows the
interlayer regions [1,2,19]. The sample showed a very high BET surface geometrical surface area as determined from particle size analysis
area (264 m2/g) which is attributed to the unique pore structure present assuming spherical geometry for the particles. It is striking to note
in the sample. that the BET surface area decreases exponentially from 264 m2/g, by

Fig. 3. Variation of specific surface area (SSABET and SSAgeo) with milling.
T.C. Alex et al. / Powder Technology 208 (2011) 128–136 131

[13]) and conform to type IV isotherm as per IUPAC classification [13].


The hysteresis loop of type H4 of the unmilled sample, according to
the IUPAC nomenclature [13], indicates the presence of narrow slit
like pores and microporosity. Milling has altered the nature of hysteresis
loop to type H3 which signifies slit-shaped pores originating from
loose aggregates of plate like particles and without any microporosity.
Fig. 5(a) and (b) shows pore size distribution and cumulative pore
area derived from the adsorption data using BJH method [14] as a
function of pore diameter. The results clearly indicate change in porosity
distribution due to milling. Pore area vs. pore diameter plot for the
unmilled sample shows two maxima at ~2 and 4 nm and majority of
the pore area (~95%), is attributed to pores which are below 5 nm. The
average pore diameter is found to be around 3 nm. After 240 min of
milling, the maxima at ~2 and 4 nm have been drastically suppressed
and a new maxima appears at 22.5 nm. Consequently, the average pore
diameter has increased to ~13 nm. In comparison with the unmilled
sample, the contribution of smaller pores (b5 nm) to surface area is
small in the 240 min milled sample. This indicates that the decrease in
specific surface area during milling results from the annihilation of pores
Fig. 4. N2 adsorption–desorption isotherms for the unmilled and the 240 min milled having size b5 nm.
boehmite samples.
3.2.3. Morphological changes
The boehmite sample before milling consisted of agglomerated
nearly 75%, to 67 m2/g after 240 min of milling. Decrease in surface platelets of uniform size (Fig. 1). SEM micrographs in Fig. 6 show the
area, after an initial increase, has been reported by Fu et al. [22] for morphological features of the sample after 3, 60 and 240 min of
CaO prepared by thermal decomposition of CaCO3, and Dellisanti et al. milling. Morphology of the samples has undergone changes during
[23] for talc [a layered structure alumino-silicate mineral, Mg3Si4O10 milling. Under the influence of planetary milling, the disjointing of
(OH)2] after prolonged milling. The decrease in specific surface area platelets from the agglomerate and breakage of these platelets occur
has been attributed by these researchers to aggregation of particles during early stages. As seen for the 3 min milled sample, the broken
during prolonged milling. Similar particle aggregation, described in particles still retain the platelet structure of the original sample.
Section 3.2.1, during the 15–240 min milling of boehmite caused the The platelets are irregularly shaped and loosely packed (Fig. 6(a)). At
geometric surface area (SSAgeo) to decrease from 2.71 to 2.14 m2/g. higher magnification, the same sample shows smaller particles
The contribution of particle aggregation, thus, to the observed sticking to larger particles (Fig. 6(b)). As the milling time is increased
decrease in the BET surface area (SSABET) is nominal. to 60 min, the platelets show signs of fusing with one another leading
A large difference between the BET surface area and geometrical to aggregation and rounding of shape (Fig. 6(c)). The smaller particle
surface area in Fig. 3 indicates porous nature of particles. In order to sticking to bigger particles diminishes almost completely and aggrega-
understand the nature of porosity, N2 adsorption and desorption tion of smaller elongated particles is seen in the sample (Fig. 6(d)). In
isotherms were examined in detail at 77 K. Typical isotherms are the 240 min milled samples, greater fusion of particles and rounding
shown in Fig. 4 for the un-milled sample and the sample after 240 min of shape are observed (Fig. 6(e)) as compared to the 60 min sample
milling. The isotherms exhibited hysteresis loop which is a character- (Fig. 6(c)). The smaller particles show a decrease in their elongated
istic of mesoporous materials (typical pore diameter range 2–50 nm character, increased fusion and possibly growth in size (Fig. 6(f)). The
SEM micrographs in Fig. 6 suggest that the broken particles reassemble
through fusion during prolonged milling. Karagedov and Lyakhov [24]
have suggested that during mechanical milling of inorganic oxides/
hydroxides, the fusion of particles may be assisted by the presence of
surface water. The fusion of particles may be responsible for the increase
in particle size (Fig. 2) and alteration in the nature of surface area and
porosity (Figs. 3–5).

3.2.4. Powder X-ray diffraction studies


Fig. 7 shows the XRD patterns of the samples before and after
milling for different durations up to 240 min. The XRD pattern of the
un-milled sample conforms to boehmite [JCPDS File No: 83-2384] and
the absence of any extra peaks after milling indicates that no new
phase is formed. The peak characteristics, however, have changed;
decrease in peak intensity and broadening of peaks occur as the
milling progressed.
The variation of peak intensity ratio (Imilled/Iunmilled) with milling
time is given in Fig. 8 for a few prominent reflections. The intensity
ratio differs significantly for different reflections during initial stages;
the difference being highest for the peak corresponding to (020)
reflection. The intensity ratios progressively converge with milling.
This observed variation of intensity ratio with milling time may be
attributed to the presence of texture in the initial sample which
Fig. 5. (a) Pore size distribution and (b) cumulative pore area for the unmilled and the disappears during milling. The decrease in the intensity ratio of each
240 min milled boehmite samples. reflection with milling time indicates disordering of the structure.
132 T.C. Alex et al. / Powder Technology 208 (2011) 128–136

Fig. 6. Typical SEM micrographs of boehmites milled for different durations: (a), (b) 3 min; (c), (d) 60 min; (e), (f) 240 min.

Fig. 9 shows the variation of micro-crystallite dimension (MCD) bending mode and is a characteristic of physically adsorbed water
and microstrain (ε) with milling time. It is observed that MCD [27,29].
decreases exponentially with milling time. The microstrain (ε) shows Milling has induced changes in IR spectrum. IR peaks for Al–O–H,
a reverse trend. The microcrystallite boundaries indicate high energy especially in the region of 3000–3500 cm−1, show a decrease in
regions. A decrease in MCD and an increase in ε signify an increase in intensity and become less and less resolved with milling time. This is
the stored energy of the material [25]. indicative of decrease in crystallite size and amorphisation [28]. This
observation supports X-ray results (Figs. 7–9). IR peaks, at 475, 597 and
3.2.5. FTIR spectra 744 cm−1, corresponding to AlO6 are also affected. In the initial stages
Fig. 10 gives IR spectra of boehmite before and after milling for of milling (after 15 min), these peaks are better resolved and have
different durations, up to 240 min. As indicated in Table 1, all the IR become equally prominent. With further milling, the peaks at 475 and
peaks in the spectra of the unmilled sample have been assigned to 744 cm−1 decrease in intensity and almost merge with the central peak
boehmite, based on published literature [26–30], and confirmed its at 597 cm−1. It is also observed that the peak at 475 cm−1 progressively
purity. The spectrum has essentially three parts. IR peaks at 3430, shifts to 518 cm−1 after 240 min of milling. These changes in the AlO6
3300, 3090, 1150, 1070 and 744 cm−1 can be assigned to various Al– IR peaks result from a combination of phenomena, including loss of
O–H vibration modes. IR peaks at 475, 597, and 744 cm−1 correspond crystallinity, microstrain and amorphisation as indicated by the analysis
to vibration modes of AlO6. The peak at 1640 corresponds to H–O–H of XRD results (Figs. 7–9).
T.C. Alex et al. / Powder Technology 208 (2011) 128–136 133

The intensity of IR peak can be correlated with the amount of specific


species [27,31]. A decrease in the intensity of peak at 1640 cm−1
suggests that the amount of physically adsorbed water decreases
with milling time. Bokhimi et al. [32] have reported that water
becomes more tightly bound with a decrease in crystallite size of
boehmite. Since XRD results indicated a decrease in the MCD of
boehmite (Fig. 9), it can be inferred that the decrease in the intensity
of peak at 1640 cm−1 and MCD are linked.

4. Effect of mechanical milling on reactivity

TG–DTA results show that the boehmite to γ-Al2O3 transformation


temperature, T(AlOOH ➝ γ-Al2O3), for the unmilled sample is 522 °C.
This value is close to the reported γ-Al2O3 transformation tempera-
ture (517 °C) from a well crystallized boehmite [2]. The percentage of
boehmite dissolved (%R) in 8 M NaOH solution at 90 °C after 30 min is
found to be 44%. Figs. 11 and 12 show the variation of T(AlOOH ➝ γ-
Al2O3) and %R, with milling time. It is evident from Figs. 11 and 12 that
Fig. 7. XRD pattern of boehmites after milling for 0, 15, 30, 60, 90, 120, 180 and 240 min. milling results in an increase in reactivity which is manifested by
lowering of transformation temperature (T(AlOOH ➝ γ-Al2O3)) and
an increase in the boehmite dissolution (%R).

5. Correlations between reactivity and physicochemical


characteristics

In order to study the nature of correlation between reactivity and


physicochemical characteristics of the sample, reactivity is defined as:
(a) a decrease in the boehmite to γ-Al2O3 transformation temperature
(ΔT(AlOOH ➝ γ-Al2O3)); and (b) percentage increase in the boehmite
dissolved (ΔR) with reference to the unmilled sample. Binary
correlations (rxy) were calculated between reactivity (i.e. y = ΔT
(AlOOH ➝ γ-Al2O3) or ΔR) and physicochemical characteristics (x) of
the samples, namely, d50, SSAgeo, SSABET, MCD and ε, and the values
are given in Table 2.
The value of correlation coefficients in Table 2 indicates that there
appears to be a very strong negative correlation between reactivity
(ΔT(AlOOH ➝ γ-Al2O3) or ΔR) and SSABET. In general, solid reactivity
correlates positively with surface area and this observation is
intriguing. While there is a decrease in surface area with milling
Fig. 8. Variation of Imill/Iunmilled for different reflections (hkl) of boehmite with milling. time, reactivity increases continuously with time suggesting that the
BET surface area has a limited role in determining the reactivity.

Fig. 9. Variation of microcrystallite dimension (MCD) and microstrain (ε) with milling time.
134 T.C. Alex et al. / Powder Technology 208 (2011) 128–136

Fig. 11. Variation of boehmite to γ-Al2O3 transformation temperature (T(AlOOH to


γ-Al2O3)) °C with milling.

Fig. 10. FTIR spectra milled for different durations 0, 15, 30, 60, 90, 120, 180 and 240 min.

Reactivity shows strong negative and strong positive correlations


with MCD and strain (ε), respectively. A decrease in MCD indicates
creations of more grain boundaries or high energy regions having high
reactivity. Similarly, higher microstrain is indicative of greater
concentration of defects which is expected to enhance reactivity
since defects are associated with higher energy or source of energy
stored during milling [25]. The correlation between reactivity and d50
and SSAgeo is moderate. This is not unexpected since these parameters
are based on the size/shape of the solid and do not provide an
adequate description of activated state of the solid samples.
The pore structure of the boehmite also undergoes changes
during milling which may influence reactivity. While there is a
decrease in surface area during milling, pores having size b5 nm are
annihilated and average pore diameter increases from 3 to 13 nm. Fig. 12. Variation of boehmite dissolution in caustic soda with milling (8 M NaOH, temp.
90 °C, time 30 min).
The mass transfer in mesoporous materials is a complex phenome-
non [33,34]. However, in general an increase in mass transfer with
an increase in average pore diameter is observed. Mass transfer of Thus, the enhanced reactivity due to milling may result from a
reactant/products, i.e. H2O (g) in the case of thermal decomposition, combined effect of change in nature of porosity and stored energy
and NaOH/NaAlO2 during alkali dissolution, is expected to become resulting from a decrease in microcrystallite dimension and an in-
increasingly favorable with an increase in pore diameter. crease in microstrain.

Table 1 Table 2
Assignment of various IR peaks in the spectra of the unmilled boehmite. Peak Binary correlations (rxy) between reactivity (y) and physicochemical characteristics (x)
assignment is based on published literature [26–30]. of the samples.

IR peak Assignment y x rxy

475 (s) AlO6 vibration ΔT(AlOOH → γ-Al2O3) d50 −0.73


597 (m) AlO6 vibration SSAgeo 0.56
744 (m) AlO6 vibration/Al–OH bending SSABET −0.98
1070 (s) Al–OH symmetric bending MCD −0.98
1150 (sh) Al–OH asymmetric bending ε 0.98
1640 (s) H–O–H bending mode ΔR d50 −0.65
3090 (m) Al–OH symmetric stretching SSAgeo 0.47
3300 (sh) Al–OH asymmetric stretching SSABET −0.99
3430 (s) Al–OH asymmetric stretching H–O–H stretching MCD −0.94
ε 0.96
s—strong, m—medium, w—weak, sh—shoulder.
T.C. Alex et al. / Powder Technology 208 (2011) 128–136 135

6. Conclusion [10] K. Tkacova, Mechanical Activation of Minerals, Elsevier, Amsterdam, 1989.


[11] C.C. Yong, J. Wang, Mechanical activation triggered gibbsite–boehmite transition and
activation derived alumina powders, J. Am. Ceram. Soc. 84 (6) (2001) 1225–1230.
Results on physicochemical characteristics and reactivity of mechan- [12] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular layers,
ically activated boehmite, prepared by thermal decomposition of gibbsite, J. Am. Chem. Soc. 60 (1938) 309–319.
[13] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouquerol, T.
are presented in the paper. The boehmite used in the study was Siemieniewska, Reporting physisorption data for gas/solid systems, Pure Appl.
characterized by very high specific area and unique slit-like pore Chem. 57 (4) (1985) 603–619.
structure. Following are the major conclusions that follow from this study: [14] E.P. Barret, L.G. Joyner, P.P. Halenda, The determination of pore volume and area
distributions in porous substances. I. Computations from nitrogen isotherms,
1. BET surface area (N2 adsorption) of the boehmite sample shows an J. Am. Chem. Soc. 73 (1951) 373–380.
[15] S. Desgreniers, K. Lagarec, XRDA: a program for energy-dispersive X-ray
anomalous decrease with milling time; approximately 75% de- diffraction analysis on a PC, Appl. Cryst. 27 (1994) 432–434.
crease is observed after 240 min of milling. [16] G.K. Williamson, W.H. Hall, X-ray line broadening from filed aluminium and
2. The decrease in surface area is attributed to aggregation of particles wolfram, Acta Metall. 1 (1953) 22–32.
[17] H.G. Jiang, M. Ruhle, E.J. Lavernia, On the applicability of X-ray diffraction line
and changes in nature of pore structure as described below: profile analysis in extracting grain size and microstrain in nanocrystalline
(a) Narrow slit like pores with microporosity changes to slit shaped materials, J. Mater. Res. 14 (2) (1999) 549–559.
[18] J.C. Davis, Statistics and Data Analysis in Geology, John Wiley & Sons, New York, 1973.
pores originating from loose aggregate of platelet type particles. [19] M.H. Stacey, Kinetics of decomposition of gibbsite and boehmite and the
(b) Pores below the size of b5 nm are progressively annihilated, characterization of the porous products, Langmuir 3 (5) (1987) 681–686.
and average pore diameter increases from 3 to 13 nm after [20] Z. Juhasz, L. Opoczky, Mechanical Activation of Minerals by Grinding: Pulverizing
and Morphology of Particles, Akademiai kiado, Budapest, 1990.
240 min, with milling.
[21] L. Opoczky, Fine grinding and agglomeration of silicates, Powder Technol. 17
3. Change in surface area and nature of porosity is accompanied by (1977) 1–7.
simultaneous changes in the structure of the boehmite sample. XRD [22] Z.Y. Fu, S.L. Wei, Mechanochemical activation of calcium oxide powder, Powder
Technol. 87 (1996) 249–254.
studies revealed that the sample undergoes amorphisation together [23] F. Dellisanti, G. Valdrè, M. Mondonico, Changes of the main physical and technological
with a decrease in microcrystallite size and an increase in microstrain. properties of talc due to mechanical strain, Appl. Clay Sci. 42 (2009) 398–404.
The results of XRD analysis are supported by FTIR spectra of the samples. [24] G.R. Karagedov, N.Z. Lyakhov, Mechanochemical grinding of inorganic oxides,
KONA 21 (2003) 76–86.
FTIR also indicated a decrease in the amount of the physically adsorbed [25] S. Srikanth, C. Sasikumar, Rakesh Kumar, Energetics and effects of mechanical
water as the water gets more tightly bound with the progress of milling. activation of materials, Proc. European Metallurgical Conference (EMC2007), 11–
4. The reactivity of the sample increases with milling as manifested 14 June, 2007, Dusseldorf, Germany, GDMB Medienverlag, Clausthal-Zellerfeld,
Germany, 2007, pp. 2015–2030.
by a decrease in boehmite to γ-Al2O3 transformation temperature, [26] A.B. Kiss, G. Keresztury, L. Farkas, Raman and IR Spectra and structure of boehmite
and an increase in the amount of boehmite dissolved in alkali (AlOOH), evidence for recently discarded D17 2h space group, Spectrochim. Acta 36A
under identical leaching conditions. (1980) 653–658.
[27] S.L. Wang, C.T. Johnston, D.L. Bish, J.L. White, S.L. Hem, Water adsorption and
5. The reactivity of milled boehmite increases in spite of a decrease in
surface area measurement of poorly crystalline boehmite, J. Colloid Interface Sci.
surface area with milling. This may be due to additional stored energy 260 (2003) 26–35.
resulting from a decrease in crystallite size and an increase in [28] S. Music, D. Dragcevic, S. Popovic, Hydrothermal crystallization of boehmite
microstrain. The enhancement of reactivity may also result due to the freshly precipitated aluminium hydroxide, Mater. Lett. 40 (1999) 269–274.
[29] Y. Lu, W.C. Lu, L.M. Zhang, B.H. Yue, X.F. Shang, J.P. Ni, Synthesis, characterization
formation of bigger pores which may facilitate mass transfer for and growth mechanism of core/shell AlOOH microspheres, Acta Phys. Chim. Sin.
reactant/product species, i.e. H2O (g) in the case of thermal 25 (7) (2009) 1391–1396.
decomposition, and NaOH/NaAlO2 during alkali dissolution. [30] Y. Noel, R. Demichelis, F. Pascale, P. Ugliengo, R. Orlando, R. Dovesi, Ab initio
quantum mechanical study of γ-AlOOH boehmite: structure and vibrational
spectrum, Phys. Chem. Miner. 36 (2009) 47–59.
Acknowledgement [31] P.F. McMillan, A.M. Hofmeister, Infrared and Raman spectroscopy, in: F.C. Hawthorne
(Ed.), Reviews in Mineralogy 18 (Spectroscopic Methods in Mineralogy and
Geology), Mineralogical Soc. America, 1988, pp. 99–159.
The gibbsite sample used in this study for the synthesis of [32] X. Bokhimi, J.A. Toledo-Antonio, M.L. Guzman-Castillo, F. Hernandez-Beltran,
boehmite was very kindly provided by M/s National Aluminium Relationship between crystallite size and bond lengths in boehmite, J. Solid State
Chem. 159 (2001) 32–40.
Company (NALCO), Bhubaneswar, India. Valuable support from Drs. T.
[33] J. Karger, D. Freude, Mass transfer in micro and mesoporous materials, Chem. Eng.
Mishra, D. Mishra, S. K. Das, B. Ravi Kumar, C. Sasi Kumar and A. J. Technol. 25 (8) (2002) 769–778.
Kailath towards FTIR, SEM, XRD and TG–DTA studies, respectively, is [34] C.E. Salmas, G.P. Androutsopoulos, A novel pore structure tortuosity concept based
gratefully acknowledged. The authors would also like to express their on nitrogen sorption hysteresis data, Ind. Eng. Chem. Res. 40 (2001) 721–730.

gratitude to the Director, National Metallurgical Laboratory (NML) for


Thomas C. Alex, Scientist, National Metallurgical Laboratory
granting permission to publish this paper. (NML), Jamshedpur (India) obtained his M.Sc degree in
Physics from University of Kerala and M. Tech in Surface
and Colloid Science from Regional Institute of Technology,
References
Jamshedpur. He specialises in mineral processing and
[1] B.C. Lippens, J.H. De Boer, Study of phase transformation during calcinations of extractive metallurgy. His current research centers around
mechanical activation of minerals and ores with emphasis on
aluminium hydroxide by selected area electron diffraction, Acta Crystallogr. 17
(1964) 1312–1321. waste utilization and minimisation.
[2] K. Wefers, C. Misra, Oxides and Hydroxides of Aluminum, Alcoa Technical Paper
19, Revised, Aluminum Company of America, Pennsylvania, 1987.
[3] S. Brühne, S. Gottlieb, W. Assmus, E. Alig, M.U. Schmidt, Atomic structure analysis
of nanocrystalline Boehmite AlO(OH), Cryst. Growth Des. 8 (2) (2008) 489–493.
[4] J.J. Kotte, Bayer Digestion and Predigestion Desilication Reactor Design, in: G.M.
Bell (Ed.), Light Metals, The Met. Soc. AIME, Warrendale, 1981, pp. 45–77.
Dr. Rakesh Kumar, Dy. Director, National Metallurgical
[5] J. Kano, S. Saeki, F. Saito, M. Tanjo, S. Yamazaki, Application of dry grinding to
Laboratory (NML), Jamshedpur (India), obtained his B.Sc
reduction in transformation temperature of aluminium hydroxides, Int. J. Miner.
(Eng) degree in Metallurgical Engineering from National
Process. 60 (2000) 91–100.
Institute of Technology, Rourkela (India) and M.Tech and Ph.
[6] K.J.D. MacKenzie, J. Temuujin, M.E. Smith, P. Angerer, Y. Kameshima, Effect of
D degree from Indian Institute of Technology (IIT), Kanpur
mechanical activation on thermal reactions of boehmite (γ-AlOOH) and γ-Al2O3,
(India). His research interests include hydrometallurgy,
Thermochim. Acta 359 (2000) 87–94.
mechanochemistry, geopolymers and blended cements. He
[7] Tonejc, M. Stubicar, A.M. Tonejc, K. Kosanovic, B. Subotic, I. Smit, Transformation
leads the Mechanical Activation and Reactivity of Solids
of γ-AlOOH (boehmite) and Al(OH)3 (gibbsite) to α-Al2O3 (corundum) induced
(MARS) group at NML.
by high energy ball milling, J. Mater. Sci. Lett. 13 (1994) 519–520.
[8] T. Tsuchida, K. Horigome, The effect of grinding on the thermal decomposition of
alumina monohydrates, α and β-Al2O3·H2O, Thermochim. Acta 254 (1995) 359–370.
[9] P. Balaz, Extractive Metallurgy of Activated Minerals, Elsevier, Amsterdam, 2000.
136 T.C. Alex et al. / Powder Technology 208 (2011) 128–136

Prof. Sanat K Roy, Head, Metallurgical & Materials Engineer- Prof. Surya Pratap Mehrotra, Professor, Indian Institute
ing at Indian Institute of Technology (IIT), Kharagpur (India) of Technology (IIT), Kanpur (India) obtained his B.Tech and
holds bachelor and doctoral degrees in Metallurgical En- P.hD in Metallurgical Engineering from IIT, Kanpur. He has
gineering from IIT, Kharagpur, and has 40 years of teaching over three decades of teaching and research experience.
and research experience. His main areas of interest include During the period, 2002–2008, he served as a Director at
Mineral Processing, Extractive Metallurgy, Environmental National Metallurgical Laboratory, Jamshedpur (India). His
Degradation of Materials, Laser Processing of Materials and research interest includes mineral processing and extractive
Nano-ceramics. metallurgy, process design & development and mechanical
activation of solids.

View publication stats

You might also like