Sie sind auf Seite 1von 11

The effect of grain and particle size on the microwave properties of barium titanate (

BaTiO 3 )
Mark P. McNeal, Sei-Joo Jang, and Robert E. Newnham
Citation: Journal of Applied Physics 83, 3288 (1998); doi: 10.1063/1.367097
View online: http://dx.doi.org/10.1063/1.367097
View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/83/6?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
Correlation between high ionic conductivity and twin structure of La 0.95 Sr 0.05 Ga 0.9 Mg 0.1 O 3
J. Appl. Phys. 100, 014107 (2006); 10.1063/1.2211308
Frequency-temperature response of ferroelectromagnetic Pb ( Fe 1 2 Nb 1 2 ) O 3 ceramics obtained by
different precursors. III. Dielectric relaxation near the transition temperature
J. Appl. Phys. 99, 124101 (2006); 10.1063/1.2201853
Dielectric relaxation in 91 % Pb ( Zn 1 3 Nb 2 3 ) O 3 9 % Pb Ti O 3 single crystal at low temperature
Appl. Phys. Lett. 84, 5317 (2004); 10.1063/1.1766081
Broadband dielectric spectroscopy of (1x) BiScO 3 x PbTiO 3 piezoelectrics
Appl. Phys. Lett. 83, 1605 (2003); 10.1063/1.1604945
Impact of domain wall displacements on the dielectric permittivity of epitaxial Ba 0.5 Sr 0.5 TiO 3 films
Appl. Phys. Lett. 79, 2052 (2001); 10.1063/1.1405147

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

JOURNAL OF APPLIED PHYSICS

VOLUME 83, NUMBER 6

15 MARCH 1998

The effect of grain and particle size on the microwave properties of barium
titanate BaTiO3
Mark P. McNeal, Sei-Joo Jang, and Robert E. Newnham
The Pennsylvania State University, Intercollege Materials Research Laboratory, University Park,
Pennsylvania 16802

~Received 27 October 1997; accepted for publication 25 November 1997!


The use of ferroelectric ceramics and thin films in microwave devices requires that they possess
frequency-stable, low-loss dielectric properties. At microwave frequencies, ferroelectric
polycrystalline ceramic materials typically exhibit a large dielectric relaxation, characterized by a
decrease in the relative permittivity ( e r ) and a peak in the dielectric loss (tan d). Mechanisms
attributed to the relaxation phenomenon include piezoelectric resonance of grains and domains,
inertia to domain wall movement, and the emission of gigahertz shear waves from ferroelastic
domain walls. As a result, the relaxation phenomenon appears to be intimately linked to the domain
state of the ferroelectric. The domain state of a ferroelectric is, in part, dependent upon its
microstructure. In this study, the microwave dielectric properties of ferroelectric barium titanate
were measured as a function of grain and particle size. Polycrystalline ceramic ferroelectric BaTiO3
~having average grain sizes of 14.4, 2.14, and 0.26 mm! and BaTiO3 powder-polymer matrix
composites ~possessing average particle sizes of 1.33 mm, 0.19 mm, and ;66 nm! were employed.
The composite samples were used to decouple resonances between adjacent grains as well as reduce
the three dimensional clamping experienced by grains in ceramic. Characterization studies were
performed to determine the effects of grain size and particle size on the crystal structure and degree
of tetragonality. Microwave dielectric measurements through 6 GHz were carried out using lumped
impedance, cavity perturbation, and post resonance experiments. All samples exhibited evidence of
relaxation or resonance phenomena in their dielectric spectra. Except for the 0.26 mm grain size
ceramics and the 66 nm particle size composites, all other samples exhibited relaxation in their
dielectric spectra. The 0.26 mm ceramic and 66 nm composite showed evidence of resonance in their
dielectric spectra. This work clearly shows the potential to tune the microwave properties of
ferroelectrics through control of grain/particle size and the domain state. In general, relaxation
frequencies increased and loss tangents decreased with decreasing grain/particle size. The relaxation
mechanisms were identified and correlated with the material characterization results and theoretical
models. Relaxation frequencies were generally governed by the smallest resonant width, i.e., the
domain width. 1998 American Institute of Physics. @S0021-8979~98!04106-1#

I. INTRODUCTION

There is an increasing demand for the use of highpermittivity ferroelectric materials in microwave devices.
Currently, the majority of microwave applications are related
to high speed microelectronics, radar, and communication
systems. High-permittivity ferroelectric materials will assist
in the material selection for decoupling capacitors currently
utilized in computer packaging. Decoupling capacitors are
used to neutralize line inductance by placing the decoupling
capacitors near the silicon chip to reduce spurious switching
in signal lines.13
Interest in ferroelectric materials for nonvolatile, high
speed random access memories ~FRAMs! and dynamic random access memories ~DRAMs! has also increased in recent
years. Ferroelectric thin films @with lead zirconate titanate
~PZT! receiving considerable attention# deposited via sol gel
or sputtering, offer the advantage of high polarizations which
lead to high charge storage densities.46
Phased array radar involves planar arrays of thousands
of closely spaced, individual antenna radiators whose com-

posite beam can be shaped and spatially directed in microseconds. Beam steering is accomplished by varying the relative phase between radiating elements. The phase shifter has
the form of an electrical delay line, which causes a phase
shift by controlled variation of the group velocity of the microwave signal.7,8 The design of the ferroelectric phase
shifter material is based on the change in permittivity under
a dc bias field applied parallel to the polarization of the rf
energy, and normal to the direction of signal propagation.
Current trends towards more complexity, higher power,
smaller size, lighter weight, lower cost, and higher frequencies through the integration of transmission lines, passive
components ~resistors, capacitors, and inductors!, and active
devices ~diodes and transistors!, has led to the development
of microwave integrated circuits ~MICs!. Monolithic microwave integrated circuits ~MMICs! are a more recent development in which the active and passive circuit elements are
grown or implanted directly on a semiconductor substrate.
These technologies have already progressed to the point
where complete microwave subsystems ~receiver front ends,
radar transmit/receive modules, etc.! can be integrated on a

0021-8979/98/83(6)/3288/10/$15.00
3288
1998 American Institute of Physics
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

single chip a few square millimeters in size. There is an


increasing need for ferroelectric materials for use in these
microwave applications, as just discussed. The use of highpermittivity ferroelectric ceramics and films in most microwave devices requires that they possess frequency stable permittivities and low dielectric losses. A fundamental concern
with the implementation of ferroelectric materials in microwave technologies is that they typically undergo a marked
relaxation in their dielectric properties. This relaxation is
characterized by a decrease in the relative permittivity, ( e r ),
with frequency, accompanied by a peak in the dielectric loss
(tan d). The high losses lead to high insertion losses which
are undesirable in most high-power applications. The origins
of this relaxation phenomenon have been attributed to the
existence of domain structures, inherent to ferroelectrics.913
In addition, as microelectronic MMICs continue to
shrink in size, smaller, more compact modules are being developed, incorporating thick and thin film technologies.
These technologies introduce specific microstructural constraints which directly influence the domain configuration of
ferroelectrics. This article reports the microwave properties
of a prototypical ferroelectric: barium titanate (BaTiO3). The
objectives of this work were to measure the dielectric properties of BaTiO3 through its microwave relaxation, and to
investigate the importance of domain walls on the relaxation
behavior. BaTiO3 ceramics with small grain sizes and 03
powder-polymer matrix composites containing small BaTiO3
particles, were studied. The composite samples, with well
controlled particle size distributions embedded in a polymer
matrix, served to decouple resonances between adjacent
grains and reduce the clamping effects experienced by grains
in ceramic. By reducing this three dimensional clamping, it
is believed that the quasifree particles possessed a domain
configuration intrinsic to particle size, and not dominated by
stress relief mechanisms. The effect of grain/particle size,
and the subsequent modulation of the domain structure, on
the microwave dielectric properties were examined by considering three different grain size regimes: ;10, ;1.0, and
;0.1 m m; and three different particle size regimes: ;1,
;0.1, and ,0.1 m m.
II. BACKGROUND
A. Size effects in ferroelectrics

Ferroic crystals have movable domain walls which result


in hysteretic behavior. Domain twinning in these materials
effectively reduces the energy of destabilization fields,
whether they be magnetic, electric, or elastic. In terms of
particle size effects, ferromagnetic materials are by far the
best studied, and it is anticipated that ferroelectrics will follow the magnetic analog as size is reduced. At large particle
sizes, twinning lowers the volume energy of the particle,
with the appearance of complex domain structures possessing several types of walls. As size is reduced further, it becomes increasingly difficult to recover the wall energy from
the volume term, consequently, it is expected that the number of domains will decrease, at first by one type, and then
the others. The first transformation is then from a polydomain particle to a single domain particle. Upon further re-

McNeal, Jang, and Newnham

3289

duction in size, the next expected transformation is from a


single domain crystallite to a superparaelectric phase. From
the magnetic analog, this superparaelectric phase would
then be expected to behave as an unpolarized, but highly
orientable single domain, possessing a high dielectric constant. Finally, since ferroelectricity is a cooperative phenomenon, it is reasonable to suppose that the system will be
forced to revert to a paraelectric state at a critical size too
small to sustain the cooperative interactions necessary for
ferroelectric behavior. In summary, the size dependence of
ferroelectric particles should exhibit four states, predicted
from the magnetic analog: multidomain, single domain, superparaelectric, and paraelectric.14
Intuitively, grain size effects are expected to differ from
particle size effects mainly because the boundary conditions
within a ceramic differ from those which govern a free particle. Most significantly, grains in ceramic are three dimensionally clamped by surrounding grains. In the case of
BaTiO3, as the ceramic cools through the cubic-tetragonal
phase transition, a tetragonal lattice distortion is introduced,
resulting in large strains in the individual grains. This leads
to rather complicated domain structures which strongly influence the electric permittivity. Typically, large grain
(.20 m m) polycrystalline ceramic BaTiO3 exhibits a relative permittivity of, e rt;1500 2000 at room temperature.
Various investigators, however, have observed e rt to increase
with decreasing grain size,1518 peak at some critical size,
and then decrease with continued size reduction.1922
The anomalous increase in e rt with decreasing grain size
has been attributed to both an increase in domain wall density with decreasing grain size ~down to ;1 m m!, as well as
to an increase in residual internal stress ~for submicron grain
sizes!. The formation of 90 ~ferroelastic! domains below the
cubic-tetragonal structural phase transition is the means by
which internal stress energy is minimized. Various attempts
have been made to correlate the 90 wall density with grain
size. It has been proposed that the width of 90 domains is
fixed at 1 mm, which suggests that no 90 domain wall can
exist in submicron BaTiO3 ceramics.23 Consequently, this
would result in an unrelieved high residual internal stress in
submicron BaTiO3 ceramics. Using a modified Devonshire
thermodynamic treatment, Buessem showed that this high
internal stress should result in an increase in e rt in accordance with observations.
Another more quantitative model has shown the twinning associated with stress relief to be rather
complicated.2022 For grain sizes greater than ;10 m m,
twinning through the formation of both 180 and 90 walls
accounts for a three dimensional correction, such that a tetragonal lattice distortion is completely accommodated
within a cubic grain. Domain width, d, was predicted to be
proportional to (g) 1/3, where g is grain size. At grain sizes
less than ;10 m m down to ;1 m m, it has been proposed
that 90 twinning allows only for a two dimensional correction; i.e., the cubic grain retains its gross shape in two dimensions ~square! by twinning; the third dimension is left
uncompensated. The condition for total elastic, electric, domain wall interface, and surface energy minimization requires that the domain width, d, be proportional to (g) 1/2,

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

3290

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

which has been shown to be valid to grain sizes down to


;1 m m. 22
Frey et al.,24 considered in detail the role of nonferroelectric interfaces in submicron, fine grain ceramic BaTiO3.
The grain size dependent CurieWeiss characteristics and
dielectric properties in the ferroelectric state were measured.
In this work, a significant enhancement in the apparent dielectric constant, e r , was observed upon decreasing grain
size from 1.7 to ;0.5 m m, where e r reached a maximum of
;4000. Below 0.5 mm, e r appeared to decrease, however,
remaining above 2000 for a grain size less than 100 nm.
These results were clearly interpreted within the context
of the internal stress model discussed previously,23 and by
treating the ceramic as a diphasic dielectric comprised of
isolated, highly stressed ferroelectric grains surrounded by a
continuously connected nonferroelectric grain boundary region. By applying a series dielectric mixing law, it was demonstrated that the measured decrease in e r with grain sizes
below ;0.5 m m was due exclusively to a series dilution arising from the grain boundary region. With decreasing grain
size, it was shown that the true dielectric constant of the
ferroelectric grains increased to ;4800 at a grain size of
;0.5 m m, and remained constant for all grain sizes less than
;0.5 m m down to ;70 nm. Frey et al., concluded that in
the submicron grain size region, domain twinning significantly decreased with an associated increase in internal residual stress, which acts to enhance e r . For grain sizes less
than 0.5 mm, it was proposed that the ceramics were comprised of substantially single-domain, highly stressed grains.

B. High frequency relaxation

Several research groups have investigated the high frequency dielectric properties of barium titanate (BaTiO3).
Previous work on both polycrystalline ceramics and polydomain single crystal suggest the presence of a large dielectric
relaxation in the low gigahertz ~GHz! frequency region.2530
The origins of this relaxation have been explained by several
investigators. Kittel9 suggested that the motion of a domain
wall has an inertial component, as the ions change position
slightly during dipole moment reorientation. The high frequency dielectric dispersion was attributed to the inertia of
the boundary, with a calculated resonant frequency of 2 GHz
for 180 domain boundaries. Devonshire10 attributed the relaxation to piezoelectric resonance of the domains; above the
acoustic resonance of the sample, polarized domains undergo
piezoelectric deformations in an alternating field, with an
effective resonant frequency approximately equal to that of a
single domain.
The frequency dependence of ferroelectricity, including
the apparent disappearance of the ferroelectric response in
the microwave region was considered by von Hippel.11 In a
ferroelectric material, such as BaTiO3, permanent electric
dipole moments are built into the crystal structure and are
therefore, firmly anchored in place, and not available for free
rotation. High frequency relaxation was attributed to the
change of the permanent net moment and with it, the creation
of a mechanical deformation that travels with the velocity of
sound. Since individual grains have dimensions on the order

McNeal, Jang, and Newnham

of 1023 cm, and since the velocity of sound is about


105 cm/s, resonance frequencies in the 100 MHz range were
predicted.
The effect of piezoelectric grain resonance in ceramics
was uniquely addressed by studying the dielectric response
of conventionally sintered and hot pressed lithium niobate
(LiNbO3) ceramics, up to 1 GHz.12 Because LiNbO3 possesses only 180 domains, with no 90 domains, the frequency of domain resonance could be related directly to
grain size. The high-frequency dielectric spectrum of
LiNbO3 was shown to have a distinct resonant character with
the relative permittivity, e r , increasing slightly, then decreasing before saturating out to its clamped value. The tan d
loss passed through a maximum in the same frequency
range.
Using an equivalent circuit model of grain resonance,
the dielectric spectrum of LiNbO3 was predicted. In the
model, a series branch ~L 1 , C 1 , and R 1 ! represented the
mechanical damping of vibration and a parallel branch, C 0 ,
represented the clamped high-frequency capacitance. The
value of the above parameters were related to grain size,
grain orientation, and material constants, in a first approximation, by assuming a cubic grain, such that, the resonant
frequency, f 0 , of a grain was given by Eq. ~1!:
f 05

1
2 p AL 1 C 1

1
2d Ar S E

~1!

In Eq. ~1!, r is the density, d is the grain size, and S E is the


mechanical compliance at a constant field.
Arlt et al.13,30 attributed the main origin of the relaxation
to the presence of ferroelastic or 90 domain walls in
BaTiO3. In tetragonally distorted perovskites, the minimization of free energy results in a characteristic domain configuration in which the domains form regular stacks with 90
domain walls between the laminar domains. In an alternating
field, above the acoustic resonance of the bulk sample and
individual grains, domains can still be deformed, but the
overall shape of the grain is preserved. The domain walls
shift, resulting in a gross shear of the stack, and the 90
domain walls were modeled as shear wave transducers. Arlt
demonstrated that the relaxation should be Debye-like
where the frequency dependence of the total permittivity is
governed by

e total5D 1 /E 1 5 e ` 1

De
,
11 j v t

~2!

with the relaxation step, De, given by,


D e 5 e 0D e r5

P 20
2c 55S 20

~3!

and the relaxation time, t, given by,

t5

2Z m S 20
k

d
.
2 n 55

~4!

In Eqs. ~2!~4!, D 1 is the dielectric displacement, E 1 is the


electric field, e ` is the intrinsic permittivity, v is the angular
frequency, P 0 is the spontaneous polarization, c 55 is the stiff-

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

McNeal, Jang, and Newnham

TABLE I. Grain size statistics of BaTiO3 ceramic samples.

Sample
Coarse grain ~CGBT!
Small grain ~SGBT!
Fine grain ~FGBT!

m gs
~mm!
14.45
2.14
0.26

TABLE II. Summary of PSD and specific surface area measurements.

s gs
~mm!
8.77
1.27
0.13

ness constant, S 0 is the spontaneous strain, and Z m is the


input acoustic impedance of the surrounding media. In arriving at Eqs. ~3! and ~4!, a force constant per unit domain wall
area, k54C 55 S 20 /d and the phase velocity, n 55 5 Ac 55 / r ,
have been used. The relaxation frequency, f r , is related to
the relaxation time, t, through f r 51/(2 p t ). Using Eq. ~4!, it
follows that
f r5

Ac 55 / r
pd

3291

~5!

The domain wall emits shear waves when its thickness is


much larger than the wavelength of the emitted acoustic
wave. However, when the domain wall diameter is smaller
than the wavelength of the emitted shear wave, the acoustic
input impedance, Z m , must be replaced by a complex impedance, which is strongly frequency dependent. Consequently,
for large emitters ~i.e., domain wall diameters.the sound
wavelength!, the relaxation is Debye-like, as suggested by
Eq. ~4!, however for small emitters ~i.e., domain wall
diameters,the sound wavelength!, the solution for e total
shows a decrease over a much broader frequency range.13,30
III. EXPERIMENTAL PROCEDURE

The ceramic and composite materials used in this study


were prepared from several sources of commercial powders.
Coarse grain ~1020 mm! BaTiO3 polycrystalline ceramics
~CGBT! were processed using conventional solid state sintering. Small and fine grain size ~gs! ceramics ~gs;1 mm and
gs;0.2 mm! were prepared using hydrothermal starting powders. Small grain ceramics of gs;1 mm ~SGBT! were prepared by dry pressing the powder into 1-in.-diam pellets. The
pellets were first uniaxially pressed at ;9 MPa followed by
cold isostatic pressing at ;280 MPa. The pellets were then
hot pressed at 53 MPa at 1150 C, for 30 min in an argon
atmosphere. The fine grain ceramics of gs;0.2 mm ~FGBT!
were prepared in nearly an identical manner, but the starting
powder was first acid washed in dilute HNO3 acid
(PH;4.0). The small and fine grain ceramic samples were
annealed in oxygen at 1100 C for 30 min. Scanning electron
microscopy ~SEM! was used to determine mean grain sizes.
The grain size statistics are summarized in Table I.
Thermal analysis was carried out on each of the powders
used in the composites to reveal the presence of residual
organics and absorbed hydroxyl groups.31,32 Thermogravimetric analysis ~TGA! on TICON HPB powder showed no
weight change up to 1000 C. Data collected on the hydrothermal powders, however, showed a precipitous decrease of
;1.4 wt % between 220 and 410 C. Another weight drop
occurred between 770 and 850 C of less than 0.3 wt %.

Powder

Mean particle
size
~mm!

Standard
deviation
~mm!

Specific surface
area
(m2/g)

TAM
BT-8
BT-16

1.33
0.19

0.77
0.17

1.37
8.26
14.81

Primary
particle size
0.72 mm
0.12 mm
66.9 nm

Consequently, the hydrothermal powders used for composite


fabrication were annealed at 500 C in oxygen for three
hours, prior to subsequent use. The annealing process had no
effect on the particle size. Although the TICON HPB
showed no weight loss, this powder was annealed in oxygen
at 1100 C for two hours. This was done to remove any
anomalous surface layers on the particles, possibly introduced during milling or grinding. Surface layers have been
shown to decrease the room temperature tetragonality, as
well as lead to remnant tetragonality above the ferroelectric
to paraelectric transition temperature, T c ; annealing under
the above conditions was shown to increase room temperature tetragonality and eliminate remnant tetragonality above
T c . 33,34
Characterization of the commercial powders used for
composite fabrication included a particle size distribution
~PSD! analysis, and specific surface area ~SSA! measurements obtained from Brunauer, Emmett, and Teller ~BET!
measurements. The particle size statistics are summarized in
Table II.
To determine the phase purity and lattice parameters of
the ceramics and annealed powders used for composite fabrication, x-ray diffraction measurements were performed using an automated diffractometer employing Ni-filtered Cu
K a radiation with a tube voltage and current of 40 kV and
35 mA, respectively. The diffractometer scan ranged between 20 and 70 2u at a rate of 0.02 2u/min.
Composite samples of 50 vol % powder loadings were
prepared using a CW Brabender electrically heated mixing
preparation center. Polypropylene was selected for the matrix material because of its thermal stability (T melt5176 C)
and its low loss at microwave frequencies ~e r 52.256, tan d
;1024 !.35 The composite material was then placed in heated
~180 C! dies and pressed at pressures not exceeding 5000
psi. Efforts were made to maintain constant pressure on all
samples as they cooled.
Relative permittivity and loss measurements from
104 Hz through 10 MHz were carried out using conventional
LCR bridge measurement techniques. The samples were polished and sputtered gold electrodes were applied to the disk
faces. Capacitance and loss were measured, and dielectric
constants were calculated from the sample dimensions.
A lumped impedance method was used mainly in the
frequency range of 50 MHz to 1 GHz. The method involved
placing a sample at the end of a shorted coaxial line and
measuring the magnitude and phase angle of the complex
reflection coefficient, G * . Figure 1 depicts the sample holder
and 50 V coaxial line. Prior to measurements, a calibration
was performed by placing known standards ~short, open, and
matched load! on the end of the coaxial air line. The dielectric properties were calculated from the following equations:

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

3292

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

McNeal, Jang, and Newnham

FIG. 2. Illustration of post resonance measurement configuration.

S D

ll 0
c
5
,
2L
vp

FIG. 1. Illustration of sample holder used for lumped impedance


measurements.

e r8 5

2G sin u
Cf
2
,
2
v C 0 Z 0 ~ G 12G cos u 11 ! C 0

~6!

e r9 5

12G 2
,
v C 0 Z 0 ~ G 12G cos u 11 ! )

~7!

where e r8 and e r9 are the real and imaginary part of the complex relative permittivity, e r* , respectively. G is the magnitude, and u is the phase angle of the reflection coefficient,
measured using an impedance analyzer. C f is a correction
made for the real part of the permittivity to account for fringing fields, and Z 0 is the characteristic impedance of the line.
The complex relative permittivity, e r* 5 e 8 2 j e 9 where
e 9 / e 8 5tan d.36
A post resonance method was used to measure dielectric
properties at frequencies greater than 1 GHz. For an isotropic
sample of radius a and length L placed between two infinite
parallel conducting plates resonating in the TE0nl mode, the
dielectric constant is calculated from the following transcendental equation:

J 0~ a !
K 0~ b !
52 b
,
J 1~ a !
K 1~ b !

~8!

where J 0 ( a ) and J 1 ( a ) are the Bessel functions of the first


kind of orders zero and one, respectively; K 0 ( b ) and K 1 ( b )
are the modified Bessel functions of the second kind of orders zero and one, respectively. The parameters a and b are
as follows:

F S DG
FS D G

a5

c
2pa
K2
l0
vp

b5

2pa
l0

c
vp

2 1/2

~9!

~10!

1/2

21

where c is the velocity of light, and n p is the phase velocity


in the dielectric medium and

~11!

where l is the number of longitudinal variations of the field


along the length of the sample and l 0 is the free space resonant wavelength. To each value of b, there is an infinite
number of solutions a n . Consequently, for a dielectric rod
resonating in the TE0nl mode, for each b l given by Eq. ~12!,
2pa
b l5
l0

FS D G
ll 0
2L

1/2

21

~12!

there corresponds an a n . Knowing the indices n and l,

e r5

S D S D

a n l 0 2 ll 0 2
1
.
2pa
2L

~13!

Dielectric loss for high e r materials can be approximated by


tan d '

1
,
Qu

~14!

where Q u is the unloaded quality factor. More precise determinations are provided in the referenced literature.3739
Samples were fabricated into cylindrical rods and placed
between two parallel conducting plates. Signal input and output coupling to the sample was obtained by 0.0859 semirigid
coaxial cables terminated with small circular loops. Figure 2
shows the measurement configuration. The frequency response of the power transmission coefficients (S 21) was measured and displayed on a Hewlett Packard 8510 network analyzer. The TE011 ~n51, l51! resonant mode/frequency was
identified and the relative permittivities and loss of the rods
were obtained from the measured values of the resonant frequency f 0 and the unloaded quality factor, Q u . Because this
method relied on identification of the resonant frequency,
lossy samples ~low Q! could not be measured using this
technique.
A cavity perturbation technique was utilized to measure
dielectric properties at 1.5, 3, and 5.6 GHz using brass cylindrical cavities resonating in the TM010 mode. In this case,
the sample is in the form of a long thin rod, the length of
which typically equals the cavity height. Another method
was utilized as well, in which the samples were less than the
height of the cavity. This facilitated dielectric measurements
of the high e r , lossy ceramic samples, which otherwise resulted in large field perturbations. Following Parkash et al.,40
the sample is treated as a dipole equal in length to the

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

McNeal, Jang, and Newnham

3293

sample. Using the method of images, the effects of the polarization of the sample and its image dipoles on the net
depolarizing field in the sample are considered, yielding an
effective depolarizing factor,
N e 5N

ph
ph
cot
,
2H
2H

~15!

where N is the depolarizing factor and depends on the axial


ratio of the specimen, 2H is the height of the cavity, and 2h
is the length of the sample. By considering the fields in both
the unperturbed and perturbed cavity, the dielectric parameters can be obtained from
Vc

e r8 215

V cd

e r9 5

G F S DG
G F S DG

dv
Vs
dv
1
2N e V c
2N e V 2c d
2
v 0 2J 1 ~ ka !
v0
2Q
2
Vs
dv
1 2
2N e V c
1 N eV cd
2
v0
2Q
2J 1 ~ ka !

S DF
F
1
2Q

Vs
2J 21 ~ ka !

,
FIG. 3. Cross section of TM010 cavity and sample used for cavity perturbation measurements.

~16!

S D
G
G F S DG

2N e V c

Vs
2N e V c
2
2J 1 ~ ka !

dv
dv
1
1N e V 2c
d
v0
v0
2Q

dv 2
1
1 N eV cd
v0
2Q

IV. RESULTS AND DISCUSSION


A. Grain size effects

~17!

where d (1/2Q) represents the difference (1/2) @ (1/Q 1 )


2(1/Q 0 ) # , Q 1 and Q 0 are the quality factors of the filled and
unfilled cavity, respectively, v 0 is the resonant frequency of
the unperturbed cavity, dv is the difference between resonant
frequencies for the unperturbed and perturbed cavity, and
(1/2)J 21 (ka) is 1.8552 for first-order mode excitation. When
the sample equals the height of the cavity, Eqs. ~16! and ~17!
reduce to,

e r8 50.539
e r9 5

Vc dv
11,
Vs v0

X-ray diffraction ~XRD! analysis was used to determine


lattice parameters and degree of tetragonality of each ceramic specimen. Lattice parameters were determined by the
Cohens least-square analytical method. Diffraction peaks
were obtained in the range of 2 u 520 to 80 and peak position was determined by analyzing the individual peaks using a Lorentzian curve fit function. Figure 4 shows the $200%
reflections obtained from the various ceramic samples. The
CGBT sample shows obvious tetragonality, as evident from
the distinct splitting of the $200% reflections. The SGBT
sample also shows a strong indication of tetragonality as

~18!

1
0.539 V c 1
2
.
2 e r8 V s Q Q 0

~19!

In this study, both the ceramic and composite samples


were varied in length between 1.1 and 1.6 cm. Crosssectional areas were on the order of 1023 and 1022 cm2 for
ceramic and composite samples, respectively. Samples were
held symmetrically along the axis of the cavity in a hole
drilled through a thin strip of mica. Figure 3 illustrates the
sample and cavity configuration. The shift in the resonant
frequency and the Q s of the filled and empty cavity were
measured, and e r8 and e r9 were calculated.
At each frequency, multiple samples were measured,
generally five or more whenever possible. The reported
value, therefore, is an average of the measured values, with
the error given by the maximum spread in the measured values. The measurement of multiple samples helped to reveal
the precision associated with the various measurement
techniques.

FIG. 4. $200% reflections from XRD patterns obtained on the various


ceramics.
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

3294

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

McNeal, Jang, and Newnham

TABLE III. Summary of calculated lattice parameters for ceramic BaTiO3.

Ceramic
CGBT
SGBT
FGBT

a lattice parameter
~!
3.992
3.991
4.015

c lattice parameter
~!
4.033
4.027
4.015

c/a ratio
~!
1.010
1.009
1

well. The structure of the FGBT ceramic, is at best, inconclusive. Figure 4 supports that the structure is tending towards cubic as the degree of splitting of the $200% reflections
is greatly reduced. However, it is expected that for a purely
cubic structure, the a 2 peak would be more clearly resolved.
Consequently, the FGBT sample can best be described as
pseudocubic. The decrease in tetragonality is attributed to the
effects of internal stress on the crystal structure. In the coarse
grain ceramic ~CGBT!, the large grain size allows for the
development of the banded domain structure which provides
a three dimensional compensation for homogenous stress.
The small grain ceramic ~SGBT! allows for partial internal
stress relief within the grains, with the development of inhomogeneous stress at the grain boundaries. It is plausible that
in the FGBT material, the grain size is too small for domain
twinning to occur, resulting in high internal stresses throughout the grain. This uncompensated stress is expected to reduce the tetragonal distortion, leading to a pseudocubic
phase. A summary of the calculated lattice parameters is provided in Table III.
Figure 5 compares the temperature dependence of the
dielectric constant observed in ceramic samples at 1 MHz. It
can be seen that above the ferroelectric-paraelectric transition temperature, T c , the dielectric properties of both the
CGBT and SGBT samples coincide, whereas the dielectric
properties of the FGBT differs dramatically from the larger
grain samples. The difference in behavior of the relative permittivity observed in the FGBT above T c relative to the other
samples, as well as the apparent broadening of the lower
temperature transitions, is again attributed to stress related
crystallographic changes. Table IV summarizes the measured
transition parameters.

TABLE IV. Summary of transition parameters in ceramic BaTiO3 samples.


Temperature Temperature
Temperature
~tetragonal
~orthorhombic
~cubic tetragonal! Curie constant orthorhombic! rhomboheral!
Sample
~C!
~C!
~C!
~C!
CGBT
SGBT
FGBT

127
124
119

132896
161361
156937

15
20
;30

290
276

Using the previously discussed measurement techniques,


the dielectric properties of the coarse grain ceramic BaTiO3
~CGBT! were measured through 6 GHz. The lumped impedance method was only useful to 500 MHz due to the limitation of the uniform field approximation as well as to the fact
that at higher frequencies, the reflections approached that of
a short, due to the high permittivity of the sample. Figure 6
shows the measured results. Clearly the dielectric properties
exhibit a large relaxation. The e r of approximately 1900
measured at 104 Hz saturates to 280 at 5.6 GHz, yielding a
D e r of approximately 1600. The peak in the loss occurs at
771 MHz. Through the use of Eq. ~5!, a domain width of
0.98 mm was calculated, in agreement with direct domain
width observations.
Using similar techniques, the dielectric properties of
both the 2 and 0.26 mm average grain size ceramics were
measured as well. Because of the high relative permittivity,
and lossy characteristics of the 2 mm grain size ceramic, the
measurement techniques were only useful up to 100 MHz. A
comparison of the measured dielectric spectra of the various
ceramic samples is shown if Fig. 7. From Fig. 7, it is obvious
that the SGBT sample exhibits the onset of the relaxation
phenomenon at the lowest frequency, ;100 kHz. At this frequency, the dielectric constant begins to decrease significantly and the loss increases with increasing frequency.
Above ;100 kHz, the SGBT exhibits the highest loss of all
the ceramic samples up to its highest measured frequency.
The very broad relaxation may be attributed to the fact that
the domain wall diameter is smaller than the wavelength of
the emitted shear wave, as discussed previously.
Up to ;10 MHz, the dielectric properties of the CGBT
and FGBT samples are virtually equivalent. Above 10 MHz,

FIG. 6. Spectrum of dielectric properties measured from CGBT ~14 mm!


FIG. 5. Comparison between the temperature dependence of the dielectric
ceramics using LCR, lumped impedance, and cavity perturbation
constant of ceramic BaTiO3 samples of different average grain size.
techniques.
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

McNeal, Jang, and Newnham

3295

TABLE V. Microwave properties of polypropylene.

FIG. 7. Comparison of dielectric spectra measured from the various ceramics ~error bars not shown for clarity!.

the CGBT exhibits Debye-like relaxation, characterized by a


sharp peak in the loss at 771 MHz. The dielectric properties
of the FGBT samples differ markedly above 10 MHz. The
dielectric constant of the FGBT remains essentially flat,
showing a slight decrease up to ;400 MHz; above this frequency, the dielectric constant increases, and as supported by
the highest frequency measurements, traverses a maximum,
and then decreases. The FGBT samples also exhibited the
lowest loss of the ceramic samples, with its loss peak apparently shifted to a higher frequency, compared to that of the
CGBT samples. The loss peak of the FGBT appears to be
centered at ;1.6 GHz. Thus, the FGBT spectrum seems
closer to a true resonance than a relaxation. It is expected
that at higher frequencies, the relative permittivity saturates
out to its clamped value. The resonant behavior is attributed
to the coupled resonance of piezoelectric single domain
grains of different size and orientation, relative to the measuring field. If this is the case, the distinctive features in the
frequency dependence of the dielectric constant and dielectric loss, should be controlled, in large part, by the majority
of grains having a grain size close to the mean grain size
~0.26 mm!; that is, the controlling dimension of the fundamental resonating unit is expected to be ;0.26 m m. Because
of the resonant character exhibited in the dielectric properties
of the FGBT samples, it is tempting to use Eq. ~1!, which
was developed to describe the equivalent circuit of a grain
near its resonant frequency. The value of S E depends on the
orientation of the grain with respect to the field and the coupling mode. However, based on literature values of S E , 41 f 0
can be placed between 5.1 and 8.5 GHz. These values seem
slightly high when compared to the observed spectrum, and
may be explained by the fact that the effective length of the
resonant units is increased through intergrain coupling,
thereby lowering the effective resonant frequency.

Diameter
~cm!

Length
~cm!

f0
~GHz!

Qu

er

tan d

2.521
1.875
1.272

1.426
0.973
0.474

9.682
13.807
25.671

1889
2421
808

2.274
2.267
2.270

0.0005
0.0004
0.0012

were fabricated and used to measure their microwave dielectric properties by the post resonance method. Table V summarizes the measured parameters and calculated dielectric
properties. It can be seen that the e r is virtually frequency
independent, and the loss is on the order of 1024 .
X-ray diffraction ~XRD! analysis was used to determine
lattice parameters and degree of tetragonality of each powder
specimen. High angle reflections were carefully examined
for peak splitting, indicative of the tetragonally distorted perovskite structure. Figure 8 shows the $220% reflections obtained from the various powders. The TAM powder shows
obvious tetragonality, as evident from the distinct splitting of
the $220% reflection. The BT-8 shows a strong indication of
tetragonality as well, albeit, to a lesser degree than the TAM
powder. Finally, the structure of the BT-16 powder, is at
best, inconclusive. Figure 8 supports that the structure is
tending towards cubic, however, it is expected that for a
purely cubic structure, the a 2 peak would be more clearly
resolved. Figure 8 verifies the fact that with decreasing particle size there is a decrease in tetragonality. The BT-16
powder can best be described as pseudocubic, with perhaps
some fraction of the particles retaining ferroelectricity. A
summary of the measured lattice parameters is provided in
Table VI.

B. Particle size effects

As discussed previously, polypropylene was selected as


the matrix material for the composites because of its low
losses in the microwave frequency range. To confirm this,
cylindrical rods of polypropylene of various dimensions

FIG. 8. $220% reflections from XRD patterns obtained on the various


powders.
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

3296

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

McNeal, Jang, and Newnham

TABLE VI. Summary of calculated lattice parameters for powder BaTiO3.

Powder

a lattice parameter
~!

c lattice parameter
~!

c/a ratio
~!

TAM
BT-8
BT-16

3.988
3.998
4.005

4.029
4.012
4.005

1.010
1.004
1

BaTiO3 powder-polypropylene composites were prepared at 50 vol % powder loadings. Figure 9 shows the measured dielectric properties for the TAM HPB ~1.33 mm!
powder-polymer composites. Frequency dependent dispersion is again apparent. The low frequency e r of 37 decreases
to 27 at 5.6 GHz, with e r apparently saturating at its highfrequency value. The loss peak obtained by curve fitting occurs at 1.5 GHz, and by Eq. ~5!, a domain width of 0.48 mm
was calculated.
Similar measurements were conducted on the composites fabricated using the hydrothermal powders. Figure 10
compares the high frequency dielectric properties measured
from the TAM HPB (;1 m m), the Cabot BT-8
(;0.2 m m), and the Cabot BT-16 ~66 nm! composites, each
at 50 vol % powder loading. At the lowest measured frequencies ~not shown in Fig. 10!, interfacial polarization increased with the surface area of the powders. These effects
were manifested mainly in the dielectric losses, of which the
BT-16 composites had the highest, and the TAM composites
had the lowest. At higher frequencies, beginning at ;1 MHz
and shown in Fig. 10, the effects of interfacial polarization
are no longer apparent. The dielectric properties of the TAM
composites and BT-8 composites are virtually identical up to
;100 MHz. Above this frequency, the tan d of the TAM
composite material undergoes a more rapid increase. The
BT-8 composite material also begins to show a marginal
increase in tan d as well. It appears that for the TAM composite, we have nearly measured through the relaxation. For
the Cabot BT-8 composite, the relaxation appears very
broad, with the loss peak shifted to a higher frequency of
3.16 GHz. From the measured relaxation frequency and Eq.

FIG. 10. Comparison of dielectric spectra measured from the various composites ~error bars not shown for clarity!

~5!, d was determined to be 0.23 mm. The calculated domain


width is nearly the same as the mean particle size, thus supporting the existence of predominately single domain particles.
The BT-16 composite material yielded the lowest e r
measured at 10 MHz. The e r of the BT-16 composite exhibited a very gradual decrease with increasing frequency, up to
;2.5 GHz; above this frequency, the onset of resonant behavior is apparent. This resonance may be attributed to the
existence of single domain particles. If this is the case, the
resonant frequency could then be related to the width of the
resonating particle. Unfortunately, using the techniques outlined in this article, higher frequency measurements of the
necessary accuracy could not be carried out. It appears the
resonant frequency of the BT-16 composite sample is higher
than that of the FGBT ceramic. This is expected because of
the smaller particle size of the hydrothermal BT-16 particles,
as well as to the fact that in the case of the ceramic, it is
likely that the grain resonances are highly coupled, leading
to lower resonance frequencies. It is clear, however, that the
reduction in particle size and subsequent modulation of the
domain structure has affected the relaxation phenomenon.
C. Domain size correlation

In those cases where it was possible to measure through


the relaxation frequency, for both ceramic and composite
samples, a domain width was calculated. Although the grains
in the ceramic are subject to different stress fields, it is nevertheless, of interest to compare the size dependent data. A
summary of the calculated domain widths is presented in
Table VII. Table VII shows that with decreasing grain or
TABLE VII. Summary of calculated domain widths.

Sample

Grain/particle
size
~mm!

fr

Calculated domain
width
~mm!

CGBT
14.4
771 MHz
0.98
TAMC
1.33
1.50 GHz
0.48
FIG. 9. Spectrum of dielectric properties measured from TAM ~1.3 mm!
BT-8C
0.19
3.16 GHz
0.23
composites using LCR, lumped impedance, and cavity perturbation
techniques.
[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

J. Appl. Phys., Vol. 83, No. 6, 15 March 1998

particle size, the domain size decreases as well. The quasifree ferroelectric BaTiO3 particles approach a single domain state at a crystallite size of ;0.2 m m.
V. CONCLUSIONS

Several measurement techniques were employed to measure the dielectric properties of ceramic BaTiO3 of different
grain sizes, and BaTiO3 powder-polymer matrix composites
with different particle sizes, through microwave frequencies.
Dielectric measurements revealed the presence of sizable relaxation phenomena in ceramics of average grain sizes of 14
and 2 mm as well as in composites fabricated with powders
having average particle sizes of 1 and 0.2 mm. The ceramic
with an average grain size of 0.26 mm and the composite
containing powders with an average particle size of 66 nm
exhibited resonant behavior. The observed dispersive behavior was interpreted within the context of existing models,
which confirmed the existence of single domain particles at
;0.2 m m. The shift in relaxation frequency was attributed to
the modulated domain structure, and demonstrates the ability
to tune the high frequency properties of ferroelectrics
through control of the microstructure.
ACKNOWLEDGMENTS

The authors graciously thank Paul Rehrig for his help in


preparation of the fine grain ceramics. This work was supported by the NSF Contract No. DMR 9223847, Size Effects
in Ferroics.
D. C. Dube and S.-J., Jang, Proc. Symp. Ceram. Dielectr. 8, 315 ~1989!.
D. Pozar, Microwave Engineering ~AddisonWesley, Massachusetts,
1990!.
3
M. Lanagan, Ph.D. thesis, The Pennsylvania State University, 1987.
4
J. Obhi and Anil Patel, Integr. Ferroelectr. 5, 155 ~1994!.
5
J. F. Scott, D. Galt, J. C. Price, J. Beall, R. Ono, C. A. Paz de Araujo, and
L. D. McMillan, Integr. Ferroelectr. 6, 189 ~1995!
6
R. E. Jones, Jr., P. D. Maniar, A. C. Cambell, R. Moazzami, and C. J.
Mogab, Integr. Ferroelectr. 5, 235 ~1994!.
7
F. J. Elmer and S.-J. Jang, Army Science Conference Proceedings, 1988
~unpublished!, p. 283.
1
2

McNeal, Jang, and Newnham

3297

R. Babbitt, T. E. Koscica, and W. C. Drach, Microw. J. 34, 63 ~1992!.


C. Kittel, Phys. Rev. 83, 458 ~1951!.
10
A. F. Devonshire, Philos. Mag. 42, 1065 ~1951!.
11
A. von Hippel, Z. Phys. 133, 1952.
12
Y. Xi, H. McKinstry, and L. E. Cross, J. Am. Ceram. Soc. 66, 637 ~1983!.
13
G. Arlt, U. Bottger, and S. Witte, Appl. Phys. Lett. 63, 602 ~1993!.
14
R. E. Newnham and Susan Troiler-McKinstry, Proc. Symp. Ceram. Dielectr. 8, 235 ~1989!.
15
H. Kniepkamp and W. Heywang, A. Angnew. Phys. 6, 385 ~1954!.
16
G. H. Jonker and W. Noorlander, Science of Ceramics ~Academic, New
York, 1962!.
17
N. C. Sharma and J. R. McCartney, J. Austr. Ceram. Soc. 10, 16 ~1974!.
18
R. J. Brandmayr, A. E. Brown, and A. M. Dunlap, U.S. Technical Report
No. E10M-2614, ~1965!.
19
K. Kinoshita and Akihiko Yamaji, J. Appl. Phys. 47, 371 ~1976!.
20
G. Arlt, D. Hennings, and G. de With, J. Appl. Phys. 58, 1619 ~1985!.
21
G. Arlt, J. Mater. Res. 25, 2655 ~1990!.
22
G. Arlt, Ferroelectrics 104, 217 ~1990!.
23
W. R. Buessem, L. E. Cross, and A. K. Goswami, J. Am. Ceram. Soc. 49,
33 ~1966!.
24
M. H. Frey, Z. Xu, P. Han, and D. A. Payne ~unpublished!.
25
J. G. Powles, Nature ~London! 162, 614 ~1948!.
26
L. Davis, Jr. and Lawrence G. Rubin, J. Appl. Phys. 24, 1194 ~1953!.
27
A. Lurio and E. Stern, J. Appl. Phys. 31, 1805 ~1960!.
28
Y. M. Poplavko, Sov. Phys. Solid State 6, 45 ~1964!.
29
Y. M. Poplavko, V. G. Tsykalov, and V. I. Molchanov, Sov. Phys. Solid
State 10, 2708 ~1969!.
30
G. Arlt, U. Bottger, and S. Witte, Ann. Phys. 3, 578 ~1994!.
31
T. Noma, S. Wada, M. Yano, and T. Suzuki, J. Appl. Phys. 80, 1 ~1996!.
32
S. Wada, T. Suzuki, and T. Noma, J. Ceram. Soc. Jpn. 104, 383 ~1996!.
33
M. Schoijet, Br. J. Appl. Phys. 15, 719 ~1964!.
34
A. K. Goswami, J. Appl. Phys. 40, 619 ~1969!.
35
M. N. Afsar, Dielectric Measurements of Common Polymers at Millimeter
Wavelength Range, IEEE MTT-S International Microwave Symposium,
June 1985, St. Louis, Missouri, p. 439442.
36
S. S. Stuchly, M. A. Rzepecka, and M. F. Iskander, IEEE Trans Instrum.
Meas. 23, 56 ~1974!.
37
B. W. Hakki and P. D. Coleman, IEEE Trans. Microwave Theory Tech. 8,
381 ~1960!.
38
Y. Kobayashi and Masayuki Katoh, IEEE Trans. Microwave Theory
Tech. 33, 586 ~1985!.
39
D. Kajfez and Pierre Guillon, Dielectric Resonators ~Artech House, Massachusetts, 1986!.
40
A. Parkash, J. K. Vaid, and Abhai Mansingh, IEEE Trans. Microwave
Theory Tech. 27, 791 ~1979!.
41
Landolt-Bornstein III edited by K.-H. Hellwege and A. M. Hellwege
~Springer, Berlin, 1981!, Vol. 16a.
8
9

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
130.64.11.153 On: Sat, 27 Sep 2014 23:06:23

Das könnte Ihnen auch gefallen