Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

ARTICLE IN PRESS

Physica B 405 (2010) 430–434

Contents lists available at ScienceDirect

Physica B
journal homepage: www.elsevier.com/locate/physb

Influence of ball milling parameters on the particle size of barium titanate


nanocrystalline powders
A.K. Nath, Chongtham Jiten, K. Chandramani Singh 
Department of Physics, Sri Venkateswara College, University of Delhi, New Delhi 110021, India

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

Article history: Barium titanate (BaTiO3 or BT) powder was synthesized from BaCO3–TiO2 reaction at high temperature.
Received 25 July 2009 This powder was then used to produce BT nanocrystalline powders using high-energy ball milling under
Accepted 28 August 2009 different milling conditions. All the milled powders were examined by XRD and TEM. The particle sizes
calculated by the Sherrer formula from the XRD patterns agree with the results of TEM images. The
PACS: particle size was investigated by varying (i) the speed of the mill from 200 to 400 rpm with regular
61.46.Hk interval of 50 rpm, fixing the milling time at 30 h, and (ii) the milling time from 15 to 40 h, at a fixed
75.50.Tt speed of 300 rpm. The increase in milling speed results in a gradual reduction in particle size till it
77.84.Je finally reaches a saturation value of about 18 nm. However, with increasing milling time, the particle
78.67.Br
size first decreases, attains a minimum value of about 16 nm and then increases. The involved
81.05.Je
mechanisms of the observed results are discussed in the present paper.
81.07.Wx
& 2009 Elsevier B.V. All rights reserved.
Keywords:
Barium titanate
High-energy milling
Milling energy
Milling time

1. Introduction particle size for ferroelectric behavior in dense BT nanocrystalline


ceramics is below 50 nm [14,15]. From these standpoints, produc-
Research on lead free piezoelectric ceramics [1,2] has been tion of nanoceramic powders having desired particle size has been
gaining importance due to the highly toxic nature of lead based gaining importance in recent time. There are other routes for the
piezoelectric ceramics which are otherwise far more superior in synthesis of nanoceramic powders such as chemical coprecipita-
piezoelectric properties. BT is one of the most studied lead free tions [16,17], sol–gel technique [18,19], hydrothermal synthesis
materials due to its potential application as multilayer ceramic [20,21]. However, the high-energy ball milling technique is
capacitors, PTC thermistors, piezoelectric transducers, actuators, regarded as a simple and cost effective method for large scale
high-e dielectrics, dynamic RAM and a great variety of electro- production of nanoceramic powders [4,22,23].
optical devices. The characteristics of such electronic ceramic are Planetary ball mills are widely used as a high-energy ball mill
markedly influenced by the particle size and morphology of the for producing nanometer scale powders of ferroelectric materials.
material [3]. A few workers have reported great improvement on In a planetary ball mill, the grinding jar, mounted on a rotating
the physical, electrical and piezoelectric properties of lead free BT support frame (called the sun wheel), also rotate around its own
ceramics synthesized from the nanoceramic powders [4–6]. axis. The centrifugal forces produced by these two superimposed
Dielectric constant is also reported to be strongly dependent on rotations act on the grinding balls and material inside the jar.
the particle size [7–9]. In ferroelectrics such as BT, it is reported As the jar and the supporting frame rotate in opposite directions,
that ferroelectricity decreases with decreasing particle size and the centrifugal forces act alternately in the same and opposite
disappears below certain critical size [10–13]. It is believed that directions. A combination of impact and frictional forces results
below the critical size the lattice changes from tetragonal to cubic between the balls and jar, which acts on the material inside the
and ferroelectricity is lost. For BT the possible average critical jar. The interplay between these forces is responsible for size
reduction of particles and at the same time the microstrains
produced in them.
 Corresponding author. Tel.: + 9111 27457625, +91 9868202566; The present work deals with the production of BT nanoceramic
fax: + 9111 24118535. powders with particle sizes in the range of about 16–80 nm by
E-mail address: kongbam@gmail.com (K.C. Singh). high-energy ball milling under different milling conditions.

0921-4526/$ - see front matter & 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.physb.2009.08.299
ARTICLE IN PRESS
A.K. Nath et al. / Physica B 405 (2010) 430–434 431

2. Experimental average particle sizes evaluated for these BT powders are plotted
as a function of the milling speed in Fig. 3. Fig. 3 shows that there
The starting ingredients were AR grade with 99.9% purity is a gradual decrease in the particle size from 80 to 16 nm as the
powders of BaCO3 and TiO2. The powders were weighed in milling speed increases from 200 to 400 rpm.
proportion to the stoichiometric ratio to yield BaTiO3 and then The average particle sizes were also determined from TEM
homogeneously mixed in isopropyl alcohol medium using ball micrographs of the BT powders milled for different durations of
milling in plastic bottle with zirconia balls. The mixture was dried 15, 20, 25, 30, 35 and 40 h, each at the speed of 300 rpm. Fig. 4
and then calcined at 1050 1C for 4 h. The calcined powder was then shows the TEM images of the powders produced under these
high-energy milled in the isopropyl alcohol medium using a milling conditions. The variation of the corresponding average
Retsch PM 100 planetary ball mill in which the sun wheel and particle sizes as a function of the milling time is shown in Fig. 5.
grinding jar rotate in opposite directions with speed ratio 1: 2. As seen in this figure, the particle size first gradually decreases,
Agate vial and balls were used. The milling was performed (i) at reaches a minimum and then increases with increase in the
the speeds of 200, 250, 300, 350 and 400 rpm of the sun wheel, milling time.
each for 30 h, and (ii) for different durations of 15, 20, 25, 30, 35 The energy supplied by the planetary ball mill during high-
and 40 h, fixing the speed at 300 rpm. During each high-energy energy milling of a powder is used in the rupture of interatomic
milling, a mass ratio of 1:5 for powder and balls was always bonds in the crystal and the formation of additional surface as a
maintained. The planetary ball mill was set to a rotational mode result of the cleavage of crystalline grains [24]. While, because of
that changes the rotational direction of the vial and the sun wheel many different process parameters and conditions, a complete
every 6 min after a rest interval of 2 min. modeling of the milling process would be an aim very difficult to
X-ray Diffractometer PW 3020 with monochromatic CuKa achieve, there have been attempts made in this direction by taking
radiation (l = 0.5405 nm) was used over a 2y angle from 201 to 801 into considerations the kinematics and thermodynamics involved.
to characterize the crystalline phase of the powders milled under According to an analytical model proposed by Gusev et al. [25],
different conditions and also to determine the crystallite sizes the dependence of post-milling particle size D of the powders on
from the FWHM of the XRD peaks by using the Debye Scherrer the milling energy Emill is expressed as
formula. The particle sizes of the milled powders were also
M½A þ BlnðDin =2bÞemax f1  expðctÞg
examined by using TEM (Morgagni 268). D¼ ð2Þ
Emill þ M½A þ BlnðDin =2bÞemax f1  expðctÞg=Din

where M is the mass of the initial powder, Din their initial size, A
3. Results and discussions and B are some constants characteristic of the material, b is the
magnitude of Burgers vector associated with a disordered network
It was observed from XRD analysis that all the milled powders of edge dislocations in the grains, e = emax[1  exp(ct)] with c o0 is
crystallize into single phase perovskite structure. Fig. 1 shows a an empirical function describing the dependence of the micro-
typical XRD pattern for the BT powder which was high-energy strains e on the milling time t.
milled at the speed of 300 rpm for 30 h. The average particle size The energy Emill consumed in milling of the powder is
‘‘t’’ is calculated using the Debye Scherrer method from the proportional to the cube of the angular rotation speed, o3, and
broadening of the diffraction line using the expression the milling time t, that is
0:9l Emill ¼ ko3 t ð3Þ
t¼ ð1Þ
b cos y
where k is a constant parameter of the milling system.
where l is the wavelength of the CuKa radiation, b is the FWHM
Using Eq. (3), the relation (2) can be written as
of the diffraction peak and y is the Bragg diffraction angle. The
average particle size of all the powders obtained ranges from 20 to M½A þBlnðDin =2bÞemax f1  expðctÞg
D¼ ð4Þ
90 nm (calculations were done taking care of the broadening due ko3 t þM½A þBlnðDin =2bÞemax f1  expðctÞg=Din
to the instrument and microstrains). Eq. (4) describes how an increase in the angular rotation speed
The particle size and morphology of the powders milled at o of the ball mill for a fixed milling time t leads to a gradual
different speeds were examined by TEM. TEM images of the decrease in the particle size D of the milled powder to a saturation
powders produced by milling at the speed of 200, 250, 300, 350 value. This prediction is strongly supported by the observed
and 400 rpm, each for 30 h are shown in Fig. 2. The corresponding decrease in the particle size with increasing rpm, as shown in Fig.
3. With increasing rpm the milling energy increases which favors
the cleavage of particles producing new finer particles. The
saturation in the particle size can be understood in terms of the
decreasing impact pressure on the particle. As the particle size of
the powder decreases, the number of particles in the zone of the
impact point contact with a grinding ball increases leading to the
decrease in pressure P imparted to a particle and the gradual
cessation of the grinding action. This explains the nature of the
experimental plot in Fig. 3 which also shows that the particle size
saturates at about 18 nm corresponding to an rpm of 400.
According to Eq. (4), for a fixed o, an increase in the milling
time should lead to a gradual decrease in the particle size D of the
milled powder to a saturation value. Such a decreasing trend of
crystal size evolution with the milling time t has also been
predicted to be represented by an exponential function of the type
[26,27]:
Fig. 1. X-ray diffraction pattern of a typical BaTiO3 powder milled at 300 rpm for
30 h. D ¼ Df þ ðDin  Df Þeðc1 tÞ ð5Þ
ARTICLE IN PRESS
432 A.K. Nath et al. / Physica B 405 (2010) 430–434

Fig. 2. TEM micrograph of BaTiO3 nanoparticles produced by milling for 30 h at the speed of (a) 200, (b) 250, (c) 300, (d) 350 and (e) 400 rpm.

where D should be considered representative of the crystal size,


being Din the initial, Df the steady-state crystal size, and c1 an
empirical constant depending on the process parameter.
The observed decrease (Fig. 4) in average particle size from
48 nm for powder milled for 15 h to 16 nm for milling time of 25 h
agrees with these predictions.
Existing theories, however, could not explain the observed
increase in particle size from a minimum of 16 nm at 25 h
to 38 nm at 40 h of milling. The increase in particle size with
milling time could be a complex process involving surface energy
and microstrains of the particles. As the particle size goes down to
nanometer range the surface to volume ratio of the particle and
hence their specific surface energy increases significantly.
On the other hand, the microstrains produced in the particles
increases as the milling time of the powder increases. Both these
processes lead to enhance instability in the powder particles. A
critical stage of instability is reached in the system with prolonged
grinding when the particles start coalescing to form bigger
particles by way of releasing the excess energy. Such an increase
of average particle size with increasing ball milling time has
also been observed in lead zirconate titanate (PZT) aqueous
Fig. 3. Variation of average particle size as a function of rpm.
suspension [28].
ARTICLE IN PRESS
A.K. Nath et al. / Physica B 405 (2010) 430–434 433

Fig. 4. TEM micrographs of BaTiO3 nanoparticles produced by milling at the speed of 300 rpm for (a) 15, (b) 20, (c) 25, (d) 30, (e) 35 and (f) 40 h.

4. Conclusion

Lead free BT nanocrystalline powders with average particle


size as small as 16 nm have been produced by high-energy ball
milling. With increasing speed of the ball mill from 200 to
400 rpm at a fixed milling time of 30 h, the particle size of the
milled powders decreases gradually till it reaches a saturation
value. This observed result agrees with a mathematical model that
describes the dependence of post-milling particle size on the rpm
of a mill. An increase in milling time from 15 to 25 h at a fixed
speed of 300 rpm yields powders with particle size decreasing to a
certain minimum and then showing an increase on further
increasing the milling time. The increasing trend of particle size
with increasing milling time suggests a mechanism in the system
that attempts to reverse the ever increasing microstrains and
specific surface energy of the grains.

Acknowledgments

The financial support from University Grants Commission


Fig. 5. Variation of average particle size as a function of milling time.
of India under the Major Research Project Scheme is gratefully
ARTICLE IN PRESS
434 A.K. Nath et al. / Physica B 405 (2010) 430–434

acknowledged. We also acknowledge the Research Section of [14] L. Mitoseriu, C. Harnagea, P. Nanni, A. Testino, M.T. Buscaglia, V. Buscaglia,
AIIMS for providing the TEM facility. M. Viviani, Z. Zhao, M. Nygren, Appl. Phys. Lett. 84 (13) (2004) 2418.
[15] Z. Zhao, V. Buscaglia, M. Viviani, M.T. Buscaglia, L. Mitoseriu, A. Testino,
M. Nygren, M. Johnsson, P. Nanni, Phys. Rev. B 70 (2004) 024107.
References [16] A. Testino, T. Buscaglia, M. Viviani, M. Buscaglia, P. Nani, J. Amer. Ceram. Soc.
87 (2004) 79.
[17] M. Buscaglia, Z. Zhao, V. Buscaglia, M. Viviani, M.T. Buscaglia, L. Mitoseriu,
[1] Y. Saito, H. Takao, T. Tani, T. Nonoyama, K. Takatori, T. Homma, T. Nagaya, A. Testino, Nanotechnology 15 (2004) 1113.
M. Nakamura, Nature 432 (2004) 84. [18] D. Kang, M. Han, S. Lee, S. Song, J. European Ceram. Soc. 23 (2003) 515.
[2] M.D. Maeder, D. Damjanovic, N. Setter, J. Electroceram. 13 (2004) 385. [19] J. Cho, M. Kubawara, J. European Ceram. Soc. 24 (2004) 2959.
[3] V. Prasad, L. Kumar, Ferroelectrics 102 (1990) 141. [20] H. Xu, L. Gao, J. Guo, J. European Ceram. Soc. 22 (2002) 1163.
[4] L.B. Kong, T.S. Zhang, J. Ma, F. Boey, Progr. Mater. Sci. 53 (2008) 207. [21] H. Xu, L. Gao, Mater. Lett. 58 (2004) 1582.
[5] T. Karaki, K. Yan, T. Miyamoto, M. Adachi, Japan. J. Appl. Phys. 46 (2007) L97. [22] T. Venugopal, K.P. Rao, B.S. Murty, J. Nanosci. Nanotechnol. 7 (2007) 2376.
[6] H. Takahasi, Y. Numamoto, J. Taaani, K. Maatsuta, J. Qiu, S. Tsurekawa, Japan. J. [23] Y. Wang, Y. Li, C. Rong, P. Liu, J. Nanotechnol. 18 (2007) 465701.
Appl. Phys. 45 (2006) L30. [24] P.Y. Butyagin, Chem. Rev. 23 (1998) 91.
[7] A. Bell, A. Moulson, Ferroelectrics 54 (1984) 147. [25] A.I. Gusev, A.S. Kurlov, Nanotechnology 19 (2008) 1.
[8] G. Arlt, D. Hennings, G. Dewith, J. Appl. Phys. 58 (1985) 1619. [26] F. Delogu, M. Monagheddu, G. Mulas, L. Schiffini, G. Cocco, J. Non-Crys. Solids
[9] K. Wu, W. Schltze, J. Amer. Ceram. Soc. 75 (1992) 3390. 232-234 (1998) 383.
[10] K. Ishikawa, K. Yoshikawa, N. Okada, Phys. Rev. B 37 (1988) 5852. [27] F. Delogu, G. Cocco, Mater. Sci. Eng. A 343 (2003) 314.
[11] M.H. Frey, D.A. Payne, Phys. Rev. B 54 (1996) 3158. [28] N. Traiphol, J. Ceram. Processing Res. 8 (2007) 137.
[12] S. Wada, T. Suzuki, T. Noma, J. Ceram. Soc. Japan 104 (1996) 383.
[13] D. McCauley, R.E. Newnham, C.A. Randall, J. Amer. Ceram. Soc. 81 (1998) 979.

You might also like