Raman Line-Shape Analysis of Nano-Structural Evolution in Cation-Ordered Zrtio - Based Dielectrics

You might also like

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

Solid State Communications 127 (2003) 433437

www.elsevier.com/locate/ssc

Raman line-shape analysis of nano-structural evolution in


cation-ordered ZrTiO4-based dielectrics
Young K. Kim, Hyun M. Jang*
Department of Materials Science and Engineering, National Research Laboratory (NRL) for Ferroelectric Phase Transitions, Pohang
University of Science and Technology (POSTECH), Pohang 790-784, South Korea
Received 18 February 2003; accepted 26 May 2003 by H. Akai

Abstract
It is known that the microwave dielectric characteristics of ZrTiO4 are greatly influenced by processing conditions such as the
degree of Sn-modification and the cooling rate after sintering. The effects of these processing variables on the nano-structural
characteristics of ZrTiO4-based dielectrics were studied by analyzing the Raman line-shape parameters using the phononconfinement model. It was shown that the nano-structural shape of the cation-ordered domains underwent a sequential change
from thin-slab form to platelet shape, and finally to spherical shape with increasing Sn-content or cooling rate.
q 2003 Elsevier Ltd. All rights reserved.
PACS: 78.30.Ly; 77.84.Dy; 64.70.Rh; 64.60.Cn
Keywords: A. Nanostructure; A. Oxides; C. Raman spectroscopy; D. Phonons

1. Introduction
Dielectric resonator now becomes the key element in
modern microwave integrated devices. Dielectric materials
for the resonator should possess high dielectric permittivity
1r ; small temperature coefficient of resonant frequency
tcf ; and low dielectric loss tan d [1]. Zirconium titanate
(ZrTiO4)-based dielectrics meet all of these requirements
(1r < 35; tcf < 2 ppm=8C; tan d < 1:0 1024 ) and have
been widely used as dielectric filters, oscillators, and MIC
substrates [2]. The microwave dielectric properties of
ZrTiO4-based materials are very susceptible to processing
conditions. Among these processing variables, the degree of
Sn-modification and the cooling rate are known to be the
most important parameters [3]. Accordingly, extensive
efforts have been made to establish processing-structureproperty relationships [3 9].
It is known that the high-temperature phase of ZrTiO4
polymorphs has the orthorhombic a-PbO2 structure (space
* Corresponding author. Tel.: 82-54-279-2138; fax: 82-54279-2399.
E-mail address: hmjang@postech.ac.kr (H.M. Jang).

group Pbcn) in which two distinct cations, Zr and Ti,


randomly distribute over the crystallographically equivalent
sites [4]. It undergoes a transition to an incommensurately
ordered state at 1125 8C on quasi-static cooling [5,6].
According to the XRD study done by Yamamoto et al. [7],
the incommensurately ordered structure can be viewed as a
stacking of the 1:1 cation-ordered thin slabs with the
interface modulation along the crystallographic a-axis by
faulted boundaries. However, the transition to the incommensurate (IC) phase is initiated by the formation of nanoscale cation-ordered clusters in the disordered matrix. Thus,
for the study of nano-scale structural features in the
beginning stage of the transformation to the IC phase, one
has to employ an alternative method because XRD is not
able to give information on the short-ranged local structure.
The main purpose of the present study is to quantitatively
characterize nano-scale structural evolutions associated
with the formation of the IC phase from the hightemperature disordered phase by applying the phononconfinement model to the analysis of asymmetric Raman
line shape. We have particularly paid attention to the
morphological evolution of the nano-scale ordered domains
with the variation of the cooling rate and the Sn-content that

0038-1098/03/$ - see front matter q 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0038-1098(03)00463-0

434

Y.K. Kim, H.M. Jang / Solid State Communications 127 (2003) 433437

are considered to be the two most effective parameters for


controlling dielectric properties of ZrTiO4-based materials.
We have demonstrated that the analysis of the Raman lineshape parameters based on the phonon-confinement model
is a promising new approach to analyzing the degree of
nano-scale atomic positional ordering.

2. Experimental section
ZrTiO4 (ZTO) and Zr12xSnxTiO4 (ZST) polycrystalline
ceramic specimens used in this study were prepared
employing a high-purity sol gel processing route [3]. The
pressed pellets were sintered at 1500 8C for 4 h, and the
sintered specimens were cooled with an initial cooling rate
of 300 8C/h down to 1250 8C. Various cooling rates were
then employed for the subsequent cooling of ZTO for the
temperature range between 1250 and 1000 8C. This was
done to critically assess the effect of the cooling rate on the
degree of short-range ordering associated with the transition
to the IC phase. Except for one particular pellet that was
directly quenched by immersing it into water at 1250 8C, all
other ZTO specimens were then air-quenched immediately
after they had reached 1000 8C to isolate the IC phase that
was not thermodynamically stable at room temperature. In
case of the sintered ZST, the cooling rate between 1250 and
1000 8C was fixed at 3 8C/h.
Analysis of local composition, as examined using a FESEM/EDS, did not indicate any non-stoichiometric compositional deviation (, 1 at.%) for both interior grain and
grain-boundary regions. ICP (inductively coupled plasma)
analysis indicated that the total impurity level of ceramic
specimens was less than 0.5 at.%.
To obtain information on the mode symmetry of Raman
peaks, especially for the peak at near 810 cm21, we
separately prepared an incommensurately ordered singlecrystal of ZTO. It was grown by slowly cooling (1 8C/h) the
solution containing a Li2MoO4 MoO3 flux from 1300 to
900 8C after the mixing process at 1300 8C for 24 h. The
single-crystal of ZTO had well defined (001) and (110)
planes, and the crystallinity of the single crystal was
assessed by examining its XRD u-rocking curve. The
measured XRD line was very narrow with the FWHM value
less than 0.118, indicating a high degree of the crystallinity.
Raman spectra of various ZrTiO4-based ceramics and a
single crystal were obtained using a NRS2100 Raman
spectrometer (JASCO, Japan) equipped with a triple-grating
monochromator and a Coherent Innova 90C Ar-laser
operating at 514.5 nm with the power of 300 mW. It has a
spectral resolution of 1 cm21. The measurement was
performed with a micro-Raman option using a backscattering configuration. For the polarized scattering of the singlecrystal ZrTiO4, we employed ZYYZ and ZYXZ scattering
geometries, where X; Y; and Z are parallel to [100], [010],
and [001] directions, respectively.

3. Results and discussion


3.1. Mode symmetry of a peak at near 810 cm21
Fig. 1 shows room-temperature Raman spectra of a
series of ZTO and ZST polycrystalline ceramics prepared
employing various processing conditions. We have presented the Raman band at near 810 cm21 only because this
peak exhibits the most pronounced change in the degree of
asymmetric line broadening with the variation of processing
conditions. Thus, it is useful to clarify the mode symmetry
associated with this peak before we will examine the effects
of processing conditions on the degree of asymmetric line
broadening.
The high-temperature cation-disordered ZTO is represented by the a-PbO2 structure with the space group Pbcn
and the point group mmm D2h : The normal vibration
modes, as obtained by applying the factor group analysis,
predict that the high-temperature phase has 18 distinctive
Raman-active optical phonon modes [10,11]. These can be
summarized using the following non-degenerate irreducible
representations:

GRA 4Ag xx; yy; zz 5B1g xy 4B2g xz 5B3g yz 1


To obtain the mode symmetry associated with the peak at
810 cm21, we have performed polarized Raman scattering
on the (001) plane of the incommensurately ordered ZTO
single crystal using both ZYYZ and ZYXZ scattering
geometries. As shown in Fig. 2, the peak at near 810 cm21 is
active in the parallel ZYYZ scattering but is not active in
the cross ZYXZ scattering except for a broad peak caused
by a polarization leakage. According to Eq. (1) or to the
character table for mmm point group, the only normal mode
that is active in the parallel ZYYZ scattering but is inactive

Fig. 1. Raman spectra of ZrTiO4-based dielectrics are shown. (a)


Raman spectra of pure ZrTiO4 (ZTO) prepared employing various
cooling rates between 1250 and 1000 8C: (i) 1 8C/h, (ii) 5 8C/h, (iii)
10 8C/h, (iv) 50 8C/h, (v) 100 8C/h, (vi) 300 8C/h, and (vii) rapidly
quenched. (b) Raman spectra of Zr12xSnxTiO4 (ZST) prepared with
various Sn-contents: (i) x 0:01; (ii) x 0:03; (iii) x 0:05; (iv)
x 0:07; (v) x 0:12; (vi) x 0:15; (vii) x 0:20; and (viii) x
0:40:

Y.K. Kim, H.M. Jang / Solid State Communications 127 (2003) 433437

435

namely:
Iv

BZ

lCq0 0; ql2 dq
with
v 2 v0 q2 G0 =22

lC0; ql2 exp2q2 L2 =4

Fig. 2. Polarized Raman spectra of the incommensurately ordered


ZTO single crystal in ZYYZ and ZYXZ scattering geometries.

in the cross ZYXZ scattering is the Ag normal mode. This


clearly indicates that the peak at 810 cm21 is not related to
Bg -type normal modes (i.e. B1g ; B2g ; and B3g ) but is caused
by the Ag mode that is symmetric to all eight distinctive
symmetric operations of mmm point group.

where v0 q denotes the phonon dispersion relation, G0 is


the intrinsic Raman line width, and L represents the phonon
correlation length.
As described in Section 1, the incommensurately ordered
ZTO can be viewed as a stacking of the 1:1 cation-ordered
thin slabs with the faulted boundaries along the a-axis. Thus,
the phonon will be one-dimensionally confined within a
cation-ordered thin-slab region. However, structural modifications either by increasing cooling rate or by increasing
Sn-content would gradually decrease the phonon correlation
parallel to the faulted boundary, eventually leading to the
formation of spherical domains. In this case, the phonon will
be three-dimensionally confined within an isolated nanoscaled sphere. Applying this concept to Eq. (2), one can
obtain the following equation for the line-shape function:
Iv

3.2. Raman line-shape analysis


Having identified the mode symmetry corresponding to
the peak at 810 cm21, we now examine the effects of
processing variables on the line shape of this peak. As
shown in Fig. 1, both the degree of the line broadening and
the asymmetry in the line shape steadily increase, in
addition to the downward shift of the mode frequency, with
increasing cooling rate for ZTO or with the degree of Snmodification for ZST. It is well known that the structural
characteristics of ordered domains in ZrTiO4-based dielectrics are greatly influenced by the cooling rate and by the
degree of Sn-modification [6,8,12]. Thus, the observed
variations of the line-shape parameters are closely related to
the structural variations of the cation-ordered domains.
To establish quantitative correlations between the
Raman line-shape parameters and the structural characteristics of the nano-scale domains, we will apply a concept of
phonon-confinement [13] to the analysis of spectral line
shape. This concept was used to quantitatively explain the
broad asymmetric line shapes observed in Si and GaAsbased semiconductors [14,15]. The asymmetry is not
expected in a single-crystal having a perfect translational
symmetry. In this case, only the Brillouin zone-center
modes are Raman-active because of the conservation of
crystal momentum. However, the introduction of defects
that limit the spatial correlation of phonons then gives rise to
a relaxation of the q 0 selection rule. A Gaussian spatial
correlation function has been used to account for the qvector relaxation related to finite-size effects and potential
fluctuations in alloy semiconductors [13 16]. The resulting
Raman line shape, Iv, is a superposition of the weighted
Lorentzian contributions over the first Brillouin zone,

1
0

L =4

2 2

f qe2q


 dq
v 2 v0 q G0 =2 2
2

where f q 1 for a thin-slab while f q 4pq2 for an


isolated sphere. In Eq. (3), q is the magnitude of the
wavevector in the unit of 2p=a; where a is the lattice
constant.
Fig. 3 shows a fitting example of the Raman line shape
using an incommensurately ordered ZTO specimen prepared with the cooling rate of 1 8C/h so that the ordered
domains are composed of thin slabs [7]. Thus, f q 1 was
used for the fitting. To evaluate the intrinsic line-shape
parameters in the absence of the confinement, we have
employed the Raman spectrum of a commensurately
ordered Zr5Ti7O24 specimen that does not possess any
faulted boundary (i.e. L ! 1) and thus provides us a
symmetric reference line for the Ag mode at 810 cm21. It
was shown that both the mode frequency and the line width
for the Ag normal mode had a good linear correlation with

Fig. 3. Measured Raman spectrum (filled circles) and computed line


profiles (solid lines) of the incommensurately ordered ZTO. DY is a
plot of the fitting residuals as a function of the Raman shift.

436

Y.K. Kim, H.M. Jang / Solid State Communications 127 (2003) 433437

the lattice parameters b [12]. The lattice parameter b, as


.
deduced from the XRD pattern of Zr5Ti7O24, was 5.324 A
Then, the corresponding intrinsic mode frequency v0
and the line width G0 ; as estimated from the extrapolated
linear relationship for the Ag symmetry [12], were 815
and 29 cm21, respectively. The line width was corrected by
the instrumental line width obtained from an Ar -laser
line.
We then adopted the following one-dimensional periodic
relation to evaluate the dispersion relation appeared in Eq.
(3):

v0 q A B cosqp

Thus, the intrinsic mode frequency is composed of the two


unknown parameters, A and B: To obtain numerical values
for A and B; we first took the estimated value for the average
period of the faulted boundary in the incommensurately
ordered ZTO (7.46lal or 3.6 nm) as the phonon correlation
length for the thin-slab confinement. The correlation length
of 3.6 nm was estimated from the corresponding XRD
superlattice peak. The optimized values of A and B; as
obtained from the line-shape fitting of Fig. 3 using Eq. (3)
with L 3:6 nm and G0 29 cm21 ; were 735 cm21 and
80 cm21, respectively.
3.3. Nano-scale structural evolution
Fig. 4 presents the computed variations of the Raman
line-shape parameters for two extreme cases (i.e. onedimensional vs. three-dimensional confinement). Eqs. (3)
and (4) with the optimized values of A and B were used to
obtain the two curves. For any particular point on the
computed curve, there exists a unique corresponding value
of L: In addition to this model computation, we also have
plotted the experimental Raman shift and the full width at
half-maximum (FWHM) of ZTO and ZST prepared with
various processing conditions for the purpose of compari-

son. The Raman spectra given in Fig. 1 were used to extract


the experimental line-shape parameters, the Raman shift and
the FWHM.
The most outstanding difference in the computed lineshape parameters between the two-extreme confinements is
that the line width increases much more rapidly with the
downward shift of the mode frequency for the thin-slab
confinement. The data points encircled by the dotted line at
the lower-right corner of Fig. 4 correspond to ZTO
specimens prepared with slow cooling rates (D 1; 5, and
10 8C/h) and ZSTs with low Sn-modification levels
(x 0:01 and 0.03). Under these line-narrowing conditions,
the experimentally obtained line-shape parameters rather
closely follow the theoretical prediction of the thin-slab
confinement with L between 3 and 4 nm. Beyond this
region, however, the line-shape parameters for both ZTO
and ZST deviate significantly from the prediction of the
thin-slab model.
For the Sn-content higher than 20 at.% (i.e. x $ 0:2), the
line-shape parameters for ZST closely follow the prediction
of the spherical-confinement model. The two data points
encircled by the broken solid line at the upper-left corner of
Fig. 4 show this trend. The phonon correlation length L in
this region had values around 4.0 nm and was not
susceptible to the Sn-content. Because the correlation length
is essentially independent of the Sn-content for x $ 0:2; the
domain size parallel to [100]-direction (or a-axis) is little
affected by the Sn-content. Contrary to this, the size of the
ordered domain perpendicular to the a-axis must be strongly
affected by the Sn-content because of the observed
morphological change in the ordered domain from the thin
slab to the three-dimensionally confined nano-sphere with
increasing Sn-content [8].
As shown in Fig. 4, the line-shape parameters for ZTO
do not closely follow the prediction of the spherical
confinement even under the condition of rapid cooling.
Thus, one can conclude that the Sn-modification is a more
effective processing parameter than the cooling rate for the
stabilization of nano-spherical domain boundaries, which is
very important to improving microwave dielectric properties of ZrTiO4-based materials. On the basis of our model
computations and analysis of the Raman line-shape
parameters, we are now able to delineate the structural
evolutions of the nano-scale ordered domains with the
variations of the Sn-content and the cooling rate. These are
schematically represented in Fig. 5.

4. Conclusions
Fig. 4. Theoretical correlations of the Raman shift for the Ag band
with its line width (FWHM) for the two distinctive models of the
phonon confinement. The line-shape parameters of ZST (open
circles) and ZTO (filled triangles) prepared with various processing
conditions are also plotted for the purpose of comparison.

The effects of processing variables on the nano-structural


characteristics of ZrTiO4-based dielectrics were studied by
analyzing the Raman line-shape parameters using the
phonon-confinement model. The results indicate that the
shape of the nano-scale ordered domains undergo a

Y.K. Kim, H.M. Jang / Solid State Communications 127 (2003) 433437

437

References

Fig. 5. Schematic representation of the effects of the Sn-content and


the cooling rate on the nano-scale structural evolution of ZrTiO4based dielectrics.

sequential change from thin-slab form to platelet shape, and


finally to spherical shape with increasing Sn-content or
cooling rate.

Acknowledgements
This work was supported by the KISTEP of Korea
through the NRL program.

[1] S. Nomura, Ferroelectrics, 49 (1983) 61.


[2] See, for example,W. Wersing, in: B.C.H. Steele (Ed.),
Electronic Ceramics, Elsevier, London, 1991, Chapter 4.
[3] S.I. Hirano, T. Hayashi, A. Hattori, J. Am. Ceram. Soc. 74
(1991) 1320.
[4] P. Bordet, A. McHale, A. Santoro, R.S. Roth, J. Solid State
Chem. 64 (1986) 30.
[5] A.E. McFale, R.S. Roth, J. Am. Ceram. Soc. 66 (1983) C-18.
[6] R. Christoffersen, P.K. Davies, J. Am. Ceram. Soc. 75 (1992)
563.
[7] A. Yamamoto, T. Yamada, H. Ikawa, O. Fukunaga, K.
Tanaka, F. Marumo, Acta Crystallogr., Sect. C: Cryst. Struct.
Commun. 47 (1991) 1588.
[8] R. Christoffersen, P.K. Davies, J. Am. Ceram. Soc. 77 (1994)
1441.
[9] Y. Park, K.M. Knowles, J. Appl. Phys. 85 (1999) 6434.
[10] M.A. Krebs, R.A. Condrate Sr., J. Mater. Sci. Lett. 7 (1988)
1327.
[11] F. Azough, R. Freer, J. Petzelt, J. Mater. Sci. 28 (1993) 2273.
[12] Y.K. Kim, H.M. Jang, J. Appl. Phys. 89 (2001) 6349.
[13] H. Richter, Z.P. Wang, L. Ley, Solid State Commun. 39
(1981) 625.
[14] K.K. Tiong, P.M. Amirtharaj, F.H. Pollak, D.E. Aspnes, Appl.
Phys. Lett. 44 (1984) 122.
[15] P. Parayanthal, F.H. Pollak, Phys. Rev. Lett. 52 (1984) 1822.
[16] D. Bersani, P.P. Lottici, X.-Z. Ding, Appl. Phys. Lett. 71
(1998) 73.

You might also like