ArticlePDF Available

Anisotropy of the Electron and Hole Drift Mobility in KNbO3 and BaTiO3

Authors:

Abstract and Figures

We determine the anisotropy of the mobility of photoexcited charge carriers in single crystals of KNbO3 and BaTiO3 by means of a purely optical method. In orthorhombic KNbO3 the mobility anisotropy is measured for both electrons and holes giving mua/muc = 1.05+/-0.06, mub/muc = 2.9+/-0.3 for electrons and mua/muc = 1.15+/-0.09, mub/muc = 1.9+/-0.2 for holes. In BaTiO3 we find for holes mua/muc = 19.6+/-0.6. Our experiment also demonstrates that the acoustic phonon contribution to the dielectric permittivity is substantially different for spatially modulated electric fields as compared to the homogeneous case.
Content may be subject to copyright.
VOLUME 78, NUMBER 1 PHYSICAL REVIEW LETTERS 6JANUARY 1997
Anisotropy of the Electron and Hole Drift Mobility in KNbO
3
and BaTiO
3
Pietro Bernasconi, Ivan Biaggio, Marko Zgonik, and Peter Günter
Nonlinear Optics Laboratory, Institute of Quantum Electronics, Swiss Federal Institute of Technology,
ETH Hönggerberg, CH-8093 Zürich, Switzerland
(Received 6 September 1996)
We determine the anisotropy of the mobility of photoexcited charge carriers in single crystals of
KNbO
3
and BaTiO
3
by means of a purely optical method. In orthorhombic KNbO
3
the mobility
anisotropy is measured for both electrons and holes giving m
a
m
c
1.05 6 0.06, m
b
m
c
2.9 6 0.3
for holes and m
a
m
c
1.15 6 0.09, m
b
m
c
1.9 6 0.2 for electrons. In BaTiO
3
we find for holes
m
a
m
c
19.6 6 0.6. Our experiment also demonstrates that the acoustic phonon contribution to the
dielectric permittivity is substantially different for spatially modulated electric fields as compared to the
homogeneous case. [S0031-9007(96)02051-0]
PACS numbers: 71.38.+i, 72.20.Fr, 72.20.Jv, 77.84.Dy
The drift mobility of photoexcited holes in tetragonal
barium titanate (BaTiO
3
) shows an anisotropy of about
a factor of 20 between the two different crystallographic
directions. A recent work [1] has pointed out that this
anisotropy has the same temperature dependence as the
anisotropy of the dielectric permittivity, a fact which lead
to speculation upon a possible linear relation between the
drift mobilities and the dielectric permittivities. Several
hypotheses have been examined trying to explain these
phenomena but none was completely satisfactory. The de-
scription of the charge transport by small-polaron hopping
was not completely successful [1], either, although it was
demonstrated that in BaTiO
3
this process is dominant [2].
We now present new measurements of the anisotropy of
the photoexcited charge carrier mobility in potassium nio-
bate (KNbO
3
), another anisotropic polar material of the
same family and with characteristics similar to those of
BaTiO
3
.
These two materials belong to the oxygen-octahedra
ferroelectrics with perovskite structure and are particularly
interesting for electro-optical, nonlinear optical as well as
for photorefractive applications [35]. Although at dif-
ferent temperatures, both crystals have the same sequence
of structural phase transitions, from the cubic to the tetra-
gonal, orthorhombic, and finally to the low temperature
rhombohedric phase. In the lower symmetry phases, both
crystals are ferroelectric and show a spontaneous polar-
ization P
S
associated with a weak lattice distortion. At
room temperature KNbO
3
is orthorhombic (point group
mm2) with the crystallographic b axis in the pseudocu-
bic [010] direction, while the a and c axes lie along the
pseudocubic [101] and [
101] directions. At room tem-
perature, BaTiO
3
has a tetragonal structure (point group
4 mm) with the axes parallel to the cubic ones. In both
materials the c axis is oriented parallel to P
S
.
We determine the anisotropy of the mobility in the bulk
of the material from the photoconductivity measured by
a purely optical method. Using light energies smaller
than the material band gap, the photoconductivity is due
to charge photoexcitation from energy levels introduced
by impurities or intentional doping. In contrast to direct
photoconductivity measurements, our results are not influ-
enced by the quality of contacts between crystal and elec-
trodes or by parasitic buildup of screening charges close
to the electrodes themselves or by surface conductivity,
all of which are particularly inconvenient effects in highly
insulating materials like KNbO
3
and BaTiO
3
.
We measure the bulk photoconductivity by detecting
the decay of the space-charge electric fields generated
by the photoconductivity itself. Two interfering laser
beams first create a spatially sinusoidal charge carrier
excitation rate which leads to a charge displacement from
the brighter to the darker regions of the interference
pattern. At steady state and small modulation depth of the
interference pattern, this process converges to a sinusoidal
charge distribution inside the impurity centers [6,7] and
the resulting dc electric space-charge field modulates the
refractive index of the crystal via the linear electro-
optic (Pockels) effect. The amplitude of the generated
phase grating is monitored by the diffraction of a weak
probe beam incident at the Bragg angle. The diffracted
intensity is proportional to the square of the space-charge
field amplitude [6,8]. We study the decay of the phase
grating when it is erased by a strong homogeneous
illumination. By measuring the grating decay times
for various orientations of the grating wave vector, we
determine the anisotropy of the photoconductivity and
thus of the charge mobility.
The amplitude of the space-charge field grating decays
exponentially. The exponential time constant t is given
by [5]
tI
e
eff
e
0
emnI
µ
1 1 k
2
g
k
2
D
1 1 k
2
g
k
2
0
, (1)
where e
eff
and e
0
are the effective [9] and the vacuum
dielectric constants, e is the unit charge, n is the density
of free charge carriers with mobility m, and k
g
is the
modulus of the grating vector. The Debye screening
106 0031-90079678(1)106(4)$10.00 © 1996 The American Physical Society
VOLUME 78, NUMBER 1 PHYSICAL REVIEW LETTERS 6JANUARY 1997
wave vector k
0
and the inverse diffusion length k
D
are
two material constants which depend on the impurity
concentration [5]. The density of photocarriers nI is
responsible for the light intensity dependence of Eq. (1).
This formula is valid for sinusoidal modulations of the
space-charge distribution obtained by a low contrast
illuminating interference pattern.
We choose the magnitude of k
g
to be much smaller
than both k
0
and k
D
. In this limit the term between
parentheses on the right hand side of Eq. (1) can be safely
neglected. The decay time constant is then directly given
by the dielectric relaxation time e
eff
e
0
兲兾共emn. Provided
that the number density of photoexcited charges nI
stays constant, we can write the following expression for
angular dependence of the decay time constant:
tq t
q 0
e
eff
q
e
eff
c
m
c
mq
, (2)
where q is the angle between the grating wave vector
and the c axis of the crystal and m
c
mq describes
the angular dependence of the mobility. In Eq. (2) an
effective permittivity e
eff
must be used because of the
sinusoidal spatial modulation of the internal electric field.
In such cases e
eff
is neither the one corresponding to a
strain-free (clamped) crystal, nor the one corresponding
to a stress-free (unclamped) crystal. The calculation of
the effective dielectric constant for a particular direction
of the space-charge field has to take into account the
piezoelectric contributions generated by the mechanical
response of the crystal to the internal sinusoidal electric
space-charge field [9].
In general, the mobility can be described by a diagonal
second rank tensor [10]. Although the static electric
field E is always parallel to the grating wave vector,
the drift current is generally not. Only the component
of the current density parallel to the grating wave vector
is responsible for the decay of the space-charge grating.
By calculating this component for every direction of the
grating wave vector, we can compute the effective scalar
mobility as a function of q . The mobility ratio appearing
in Eq. (2) is then given by
mq
m
c
cos
2
q 1
m
a,b
m
c
sin
2
q , (3)
where the subscripts a , b, and c denote the axes along
which the mobility is considered.
In contrast to previous studies [1,11] where only
two extremal points were measured, we determine the
full dependence of the decay times t of the space-
charge grating on the angle q between k
g
and the c
axis. We choose a very large grating period of 45 mm
(k
g
0.14 mm
21
) in order to neglect the terms in the
parentheses on the right hand side of Eq. (1). Taking into
account that k
0
, k
D
. 1.25 mm
21
, the error introduced is
less than 1%. The measurements have been performed
in many different KNbO
3
crystals including pure, iron
doped, p-type and n-type samples, and in a nominally
pure BaTiO
3
probe. The crystals, cut with the surfaces
normal to the crystallographic axes, are mounted so
that they can be rotated around the surface normal
bisecting the directions of the two beams which produce
the grating (I
w1
0.03 Wcm
2
and I
w2
0.21 Wcm
2
).
These two beams and the erasing beam (I 3.5 Wcm
2
)
are provided by an Ar
1
laser operating at 488 nm. All
polarizations are always adjusted perpendicular to the c
axis to reduce possible beam coupling or beam fanning
effects [12]. The monitor beam (He-Ne laser at 633 nm,
I 0.01 Wcm
2
) is polarized along the c axis. While
rotating the samples, we readjust all the polarizations by
means of l2 plates in order to ensure the same writing
and, in particular, identical erasing conditions. In this way
nI remains unchanged and independent of q .
After the grating has reached the steady state, the
grating decay is induced by blocking the two writing
beams and by simultaneously switching on the strong
erasing beam. All the decay curves measured for different
q follow a single exponential function over at least ten
decay time constants. We perform a least-squares fit of
the angular dependence of t using Eq. (2), in which the
angular dependence is governed by the single parameter
m
a,b
m
c
. The scaling parameter t0 does not influence
the form of the curves; e
eff
q 兲兾e
eff
q 0
is known [9,13].
Figure 1 shows the grating decay time vs grating
wave vector orientation in the BaTiO
3
crystal, while
Figs. 2 and 3 present the results for KNbO
3
. The solid
lines correspond to the least squares fit using Eq. (2).
The agreement is remarkable. The mobility anisotropies
obtained from the fits are summarized in Table I. The
measurement performed in the hole conducting BaTiO
3
crystal agrees with the value obtained in Refs. [1,11,14].
Experiments in electron conducting samples of BaTiO
3
could not be performed because no suitable crystal was
FIG. 1. Decay time of the photoinduced space-charge grating
in a nominally pure BaTiO
3
crystal as a function of the angle
between k
g
and the c axis of the crystal. The solid line
represents the fit with Eq. (2) using e
eff
q . The dashed curve
is the fit obtained when using the uncorrected clamped or
unclamped dielectric constants.
107
VOLUME 78, NUMBER 1 PHYSICAL REVIEW LETTERS 6JANUARY 1997
FIG. 2. Decay time of the photoinduced space-charge modu-
lation in an iron doped hole conducting KNbO
3
crystal as a
function of the angle between k
g
and the c axis of the crystal.
Angular dependence in the a-c plane (above) and in the b-c
plane (below). The solid line represents the fitted curve calcu-
lated from Eq. (2).
available. The only previous measurements in KNbO
3
have been published in Ref. [15], but they are affected by
a poor accuracy [16]. Since the absolute value of the drift
band mobility along the c axis for the electrons in KNbO
3
is m
c
0.5 6 0.1 cm
2
Vs
21
[17] we have m
a
0.6 6
0.2 cm
2
Vs
21
and m
b
1.0 6 0.3 cm
2
Vs
21
. These
values are similar to m
c
0.13 6 0.03 cm
2
Vs
21
and
m
a
1.2 6 0.3 cm
2
Vs
21
found in BaTiO
3
by Hall
experiments [18]. Unfortunately, no data are available on
the absolute hole band mobility for these two materials.
Our measurements lead to several observations. The
mobility ratios do not depend appreciably on the doping
concentration since a concentration of the trapping centers
of the order of 10
25
per unit cell can influence the charge
density nI but not, or only to a negligible extent, the
charge carrier mobility. In KNbO
3
, the mobility ratio
m
a
m
c
for both electrons and holes is very similar and
close to unity, while the ratio m
b
m
c
shows a larger
anisotropy. In BaTiO
3
, m
a
m
c
is even larger.
FIG. 3. Decay time of the photoinduced space-charge modu-
lation in an iron doped electron conducting KNbO
3
crystal as a
function of the angle between k
g
and the c axis of the crystal.
Angular dependence in the a-c plane (above) and in the b-c
plane (below). The solid line represents the fitted curve calcu-
lated from Eq. (2).
The low mobilities observed in both materials can be
understood on the basis of a small-polaron model [19].
In the one dimensional case, the small-polaron hopping
mobility is of the type [20]
m
ea
2
J
2
¯hk
B
T
p
4E
A
k
B
Tp
expE
A
k
B
T , (4)
where a is the interatomic distance, E
A
is the activation
energy, and J is the overlap integral of the electronic
wave functions of the adjacent sites. The anisotropy
of the overlap integral can be considered as the origin
of the mobility anisotropy [21]. For highly localized
wave functions, the overlap integral can be assumed to
vary exponentially as J ~ exp2ar
0
where r
0
is the
localization length. Thus using Eq. (4) we can write
m
a,b
m
c
µ
a
a,b
a
c
2
exp22a
a,b
2 a
c
兲兾r
0
. (5)
In KNbO
3
the lattice constants along the a, b, and
108
VOLUME 78, NUMBER 1 PHYSICAL REVIEW LETTERS 6JANUARY 1997
TABLE I. Photoconductivity type (P.C.), dopant concentra-
tion (Nom. pure: Nominally pure), and drift mobility ratios
in single crystals of KNbO
3
and BaTiO
3
.
Sample P.C. Doping
m
a
m
c
m
b
m
c
KNbO
3
p Fe: 150 ppm
a
1.05 6 0.06 2.9 6 0.3
p Nom. pure · · · 3.2 6 0.4
p Nom. pure · · · 2.9 6 0.3
n Fe: 20 ppm
a
1.15 6 0.09 1.9 6 0.2
BaTiO
3
p Nom. pure 19.6 6 0.6
n Nom. pure
b
9.2 6 5.1
c
a
Concentration in FeNb atomic ratio in the crystal.
b
Reduced in flowing hydrogen [19].
c
As determined from Hall measurements in Ref. [19].
c axes are a
a
5.6896 Å, a
b
3.9692 Å, and a
c
5.7256 Å which, combined with the mobility ratios, give
a polaron localization r
0
between 0.5 and 3.5 Å. These
values would be compatible to the small-polaron model.
However, this is not the case for BaTiO
3
where the
polaron localization was calculated to be 0.03 Å [1],
which appears to be unphysically small.
In BaTiO
3
another puzzling relation was noticed be-
tween mobility and dielectric permittivity andMahgerefteh
et al. advanced the hypothesis that the two quantities are
not independent. In their work they found m
c
m
a
Re
c
e
a
where R is a constant close to 2 and temperature
independent over the whole tetragonal phase. Our mea-
surements have been carried out at room temperature but
the results demonstrate that R has no universality pretence
since its values in KNbO
3
are close to 4 and 9 for m
c
m
a
and m
c
m
b
, respectively.
As already pointed out above, the solid curves in all the
figures have been calculated using the effective dielectric
permittivity which takes into account the piezoelectric as
well as the elastic contributions in the case of a sinusoidal
space-charge field. To show the importance of using the
correct expression for e
eff
, in Fig. 1 we also plot the curve
of Eq. (2) when e
eff
is calculated from either the clamped
or the unclamped values. An evident disagreement with
the measured data is obtained. Our measurement of the
dielectric relaxation time as a function of the direction of
the space-charge field is thus the first direct experimental
confirmation of the validity of the expressions for the
effective dielectric constant given in Ref. [9].
In conclusion, we have shown that optical methods
can provide a reliable tool for the characterization of
the charge drift mobility in photoconductive materials.
The measurement of the decay time of a photoinduced
space-charge fields as a function of the crystal orientation
allows the precise determination of the charge carrier drift
mobility ratios.
The localization length of the quasifree charge carri-
ers is calculated in KNbO
3
according to the small-polaron
model. Its value is consistent with the theory, while that
obtained for BaTiO
3
is too small. A further comparison of
the results in BaTiO
3
and in electron and hole conducting
KNbO
3
shows that the value which relates the mobility
ratio with the corresponding dielectric permittivity ratio is
a material parameter. The explanation of the temperature
independence of this parameter noticed in BaTiO
3
, and
the complete description of the charge transport mecha-
nisms in anisotropic polar materials such as KNbO
3
and
especially BaTiO
3
, are still open questions.
We are grateful to Hermann Wüest and Michael Ewart
for crystal growth and post-growth treatment and to
Jaroslav Hajfler for expert preparation of KNbO
3
samples.
The authors also thank Germano Montemezzani for help-
ful comments and enlightening suggestions.
[1] D. Mahgerefteh, D. Kirillov, R.S. Cudney, G.D. Bacher,
R.M. Pierce, and J. Feinberg, Phys. Rev. B 53, 7094
(1996).
[2] H. Ihrig and D. Hennings, Phys. Rev. B 17, 4593 (1978),
and references therein.
[3] M.E. Lines and A.M. Glass, Principles and Applications
of Ferroelectrics and Related Materials (Clarendon, Ox-
ford, 1977).
[4] P. Günter, Phys. Rep. 93, 199 (1982).
[5] Photorefractive Materials and Their Applications I: Fun-
damental Phenomena, edited by P. Günter and J. P. Huig-
nard, Topics in Applied Physics Vol. 61 (Springer-Verlag,
Berlin, 1988).
[6] N.V. Kukhtarev, V.B. Markov, S. G. Odulov, M. S.
Soskin, and V. L. Vinetskii, Ferroelectrics 22, 949 (1979).
[7] G.C. Valley and M.B. Klein, Opt. Eng. 22, 704 (1983).
[8] H. Kogelnik, Bell Syst. Tech. J. 48, 2909 (1969).
[9] M. Zgonik, P. Bernasconi, M. Duelli, R. Schlesser,
P. Günter, M. H. Garrett, D. Rytz, Y. Zhu, and X. Wu,
Phys. Rev. B 50, 5941 (1994).
[10] J.F. Nye, Physical Properties of Crystals (Clarendon,
Oxford, 1957).
[11] C.P. Tzou, T.Y. Chang, and R.W. Hellwarth, Proc. SPIE
Int. Soc. Opt. Eng. 613, 58 (1986).
[12] G. Montemezzani, A.A. Zozulya, L. Czaia, D Z. Ander-
son, M. Zgonik, and P. Günter, Phys. Rev. A 52, 1791
(1995).
[13] M. Zgonik, R. Schlesser, I. Biaggio, E. Voit, J. Tscherry,
and P. Günter, J. Appl. Phys. 74, 1287 (1993).
[14] S. Sochava, K. Buse, and E. Krätzig, Opt. Commun. 98,
265 (1993).
[15] S. Sochava, K. Buse, and E. Krätzig, Opt. Commun. 105,
315 (1994).
[16] K. Buse (private communication).
[17] M. Ewart, I. Biaggio, M. Zgonik, and P. Günter, Phys.
Rev. B 49, 5263 (1994).
[18] C.N. Berglund and W.S. Bear, Phys. Rev. 157, 358
(1967).
[19] D. Emin, Phys. Today 35, 34 (1982).
[20] D. Emin and A.M. Kirman, Phys. Rev. B 34, 7278 (1986).
[21] Y.G. Girshberg, E. V. Bursian, and Y. A. Grushevsky,
Ferroelectrics 6, 53 (1973).
109
... The ratio of the mobility along a-axis to that along c-axis is 9.2 for eand 19.6 for h + [74]. First, the ratio of surf of P+ surface (e -) to a/c state was much larger than this anisotropy. ...
... First, the ratio of surf of P+ surface (e -) to a/c state was much larger than this anisotropy. Second, the T-dependence of surf of P+ and P+ surface was different from that of the anisotropy [74]. Third, inplane capacitance sometimes indicated that a-axis of a/c surface was aligned along the inplane conduction path, but the inplane conduction of the a/c state was much lower than that of P+ and P+ state (near 300 K). ...
Preprint
We find novel polar orders that yield electron (e-) and hole (h+) gas and depend on surface terminations, using density functional theory (DFT) that, unlike existing reports, relaxed all the ion positions of ATiO3 having spontaneous polarization Ps (A: alkali earth metal). By the experiments of atomic-oxygen cleaned surfaces of BaTiO3, we find both e- and h+ gas that are proven to originate from Ps and constrain electrostatic potential, which has been missing. These experiments that remarkably agree with the DFTs of defect free BaTiO3 reveal the properties of Ps-originated e-h+ and, for ferroelectric basics, an e-h+-posed intrinsic constraint on depolarization field arising from Ps in proper time ranges.
... It appears there is a direct correlation between E subs g and E KNO/subs g which explains why a lower interface band gap is obtained for AgCl. Most importantly, KNbO 3 is a p-type absorber [94], where minority carriers are electrons excited into the conduction band. Thus, electron injection from KNbO 3 thin-film conduction band to the substrate/electrode conduction band determines the efficiency of the solar cell. ...
Preprint
div> Conventional solar cell efficiency is usually limited by the Shockley-Queisser limit. This is not the case, however, for ferroelectric materials, which present a spontaneous electric polarization that is responsible for their bulk photovoltaic effect. Even so, most ferroelectric oxides exhibit large band gaps, reducing the amount of solar energy that can be harvested. In this work, a high-throughput approach to tune the electronic properties of thin-film ferroelectric oxides is presented. Materials databases were systematically used to find substrates for the epitaxial growth of KNbO3 thin-films, using topological and stability filters. Interface models were built and their electronic and optical properties were predicted. Strain and substrate-thin-film band interaction effects were examined in detail, in order to understand the interaction between both materials. We found substrates that significantly reduce the KNbO3 band gap, maintain KNbO3 polarization, and potentially present the right band alignment, favoring the electron injection in the substrate/electrode. This methodology can be easily applied to other ferroelectric oxides, optimizing their band gaps and accelerating the development of new ferroelectric-based solar cells. </div
Article
Full-text available
Here, the first experimental demonstration on the effect of incorporating new generation 2D material, MXene, on the thermoelectric performance of rare‐earth‐free oxide perovskite is reported. The charge localization phenomenon is predominant in the electron transport of doped SrTiO3 perovskites, which deters from achieving a higher thermoelectric power factor in these oxides. In this work, it is shown that incorporating Ti3C2Tx MXene in a matrix of SrTi0.85Nb0.15O3 (STN) facilitates the delocalization of electrons resulting in better than single‐crystal‐like electron mobility in polycrystalline composites. A 1851% increase in electrical conductivity and a 1000% enhancement in power factor are attained. Besides, anharmonicity caused by MXene in the STN matrix has led to enhanced Umklapp scattering giving rise to lower lattice thermal conductivity. Hence, 700% ZT enhancement is achieved in this composite. Further, a prototype of thermoelectric generator (TEG) using only n‐type STN + MXene is fabricated and a power output of 38 mW is obtained, which is higher than the reported values for oxide TEG.
Article
In recent years, forming high entropy oxides has emerged as one of the promising approaches to designing oxide thermoelectrics. Entropy engineering is an excellent strategy to improve thermoelectric performance by minimizing the thermal conductivity arising from enhanced multi-phonon scattering. In the present work, we have successfully synthesized a rare-earth-free single phase solid solution of novel high entropy niobate (Sr0.2Ba0.2Li0.2K0.2Na0.2)Nb2O6, with a tungsten bronze structure. This is the first report on the thermoelectric properties of high entropy tungsten bronze-type structures. We have obtained a maximum Seebeck coefficient of -370 μV K-1 at 1150 K, which is the highest among tungsten bronze-type oxide thermoelectrics. The minimum thermal conductivity of 0.8 W m-1 K-1 is obtained at 330 K, which is so far the lowest reported value among rare-earth-free high entropy oxide thermoelectrics. This synergistic combination of large Seebeck and record low thermal conductivity gives rise to a maximum ZT of 0.23 which is so far the highest among rare-earth free high entropy oxide-based thermoelectrics.
Conference Paper
The physical and optical properties of inorganic and organic materials with large photorefractive effect are often strongly anisotropic. Three important kind of anisotropies can be identified. First, the magnitude of the electro-optic effect and of the low frequency dielectric constant depend on the direction of the internal space-charge field and, specially in inorganic crystals, is strongly affected by mechanical coupling within the material. Second, charge transport, i.e. carrier mobilities, can differ significantly for different drift directions [1]. Finally, photoexcitation cross-sections can be anisotropic with respect to wave polarization. This affects the magnitude of the photorefractive space-charge fields and can give rise to a significant enhancement of two-wave mixing gain coefficients [2].
Article
Despite having large electron concentration, the electrical conductivity of Nb doped SrTiO3 severely suffers from poor carrier mobility. Even after taking several attempts by different research groups, it is still challenging to improve carrier mobility without affecting the Seebeck coefficient. In this report, 1-dimensional CNT has been introduced in SrTi0.85Nb0.15O3 (STN) matrix to fabricate STN + CNT nano composites as CNT is expected to have boosted the mobility by several orders. We have reported more than 1000 % increase in electrical conductivity after CNT addition. In contrast, Seebeck coefficient remains flat with CNT loading. This remarkable enhancement of electrical conductivity accompanied by essentially no change in thermopower leads to a 380% increase in the ZT parameter. Although CNT has been previously introduced in chalcogenides, binary oxides, and oxy-chalcogenides, such a remarkable surge in electrical transport with the incorporation of a small amount of CNT has never been reported. Moreover, to the best of our knowledge, this is the first report on thermoelectric properties of oxide perovskite composites with CNT. Our results suggest that forming oxide-CNT nano-composites can be a robust strategy to improve the thermoelectric performance in bulk materials for realistic applications.
Article
Full-text available
Measurements of the space-charge grating buildup in the dark are performed in potassium niobate using short pulse excitation at a wavelength of 532 nm. The experiments allow the separate determination of the lifetime of the photoexcited electrons and of their mobility in a number of KNbO3 crystals. The lifetime of (3.9±0.2) ns is shown to be an intrinsic property of KNbO3. Estimates of the electron mobility are refined to (0.5±0.1) cm2 V-1 s-1 in reduced KNbO3. From saturation effects we derive an electron donor photoexcitation cross section of (2.6±1.0)×10-17 cm2.
Conference Paper
We have measured the anisotropy of the mobility of photo-excited holes in nominally undoped single-crystal, poled, ferroelectric BaTi03 (p-BaTi03) at room temperature. In this crystal, a photorefractive grating results from the spatial rearrangement of trapped holes which have been excited by light-intensity beats and then drift and diffuse before recombining at similar empty traps. The resulting Coulomb field grating gives rise to a refractive index grating (the "photorefractive grating") via the electro-optic effect. In this uniaxial crystal the mobility of the holes is different parallel (µ//) or perpendicular (µT) to the crystal axis (c-axis). As is well known, a uniform light beam causes an established photorefractive grating to decay exponentially in time. We have measured this decay rate with the grating wavevector oriented at various angles to the c-axis and compared these rates with the prediction of the standard model which assumes simple drift and diffusion with direct recombination time t, 1 as extended by us to allow different drift and diffusion rates parallel and perpendicular to the c-axis. The data is satisfactorily fit by this model if we use µT/µ//= 18 ± 7 and µ //т = 5 x IO“10 cm2/Volt. We also require a ratio of dielectric constants £T/£//, = 25 ± 15, which is consistent with the ratio measurements by other methods in other samples. To fit the data, we also needed values for the density of excitable traps (- 2 x 1016 cm-3) and photoconductivity parallel to the c-axis, which we took from previously published measurements on the same crystal.2 Our measurements were made with 8 nsec pulses at A = 532 nm. We could see no effects of finite hole recombination time, which suggests that the recombination time must be less than 1 nsec.If this is true, and recombination is a direct process, then the mobility must be larger than 0.5 cm2/sec/Volt parallel to the c-axis.
Article
Scitation is the online home of leading journals and conference proceedings from AIP Publishing and AIP Member Societies
Article
Conductivity, Hall-effect, and Seebeck-coefficient measurements on single-domain crystals of n-type BaTiO3 are described. In the cubic phase above 126°C, the electron mobility is 0.5 cm2/V sec. When the material becomes tetragonal, the mobility perpendicular to the c axis increases rapidly to 0.54 cm2/V sec a few degrees below the transition temperature, then increases more slowly to 1.2 cm2/V sec as the temperature decreases to 26°C. The mobility parallel to the c axis decreases rapidly to 0.35 cm2/V sec a few degrees below the transition, then decreases slowly to 0.13 cm2/V sec at 26°C. From the Seebeck-coefficient measurements, the density-of-states effective mass of cubic BaTiO3 is (6.5±2)m0. The experimental data in the tetragonal state are interpreted in terms of a relatively simple distortion of the band structure applicable to cubic perovskite-type semiconductors.
Article
Non-steady-state photocurrent measurements are carried out with a photorefractive BaTiO3 crystal at the light wavelength 514 nm. We obtain the diffusion length for different crystal orientations, and derive the mobility ratio for charge carrier migration parallel and perpendicular to the c-axis. The frequency dependence of the photocurrent reveals two characteristic times.
Article
Splitting of the mobility tensor component at the transition of BaTiO3 crystal from the cubic state to the tetragonal state is explained within the limits of the small-radius polaron model by the change of the overlapping integral magnitude. A quantitative comparison is made between theory and experiment and for this purpose the data obtained by Berglund and Baer, and Murakami were reproduced. From the same point of view Wemple's data about the dependence of mobility on hydrostatic pressure and a great difference of mobilities in SrTiO3 are considered.
Article
We usually think of electrons in solids and liquids as behaving like free particles whose motion is impeded by occasional collisions. Evidence is now accumulating for charge carriers with qualitatively different behavior. These new charge carriers, in the simplest case, are extra electrons in condensed matter that become trapped in potential wells of their own creation, forming units known as small polarons. Excess charges trap themselves in solids or liquids by shifting the surrounding atoms, yielding entities whose behavior sheds light on phenomena ranging from charge solvation to photocarrier survival.
Article
Non-steady-state photocurrent measurements are carried out with a n-type photorefractive KNbO3:Fe crystal at the light wavelength 488 nm. Spatial frequency dependent measurements yield the effective trap density. From the dielectric relaxation time for different crystal orientations we deduce the mobility ratios for charge carrier migration parallel to the a-, b- and c-axis.
Article
The electrical transport properties of BaTiO3 are discussed against the background of recent defect-chemistry investigations and in the light of the small-polaron nature of the charge carriers. Seebeck-effect measurements and chemical investigations on highly doped BaTiO3:La ceramics are also reported. It was found for BaTiO3 that the dissociation energy required to set the conduction electrons free from their donor centers and also the kinetic-energy term of the small polarons contributing to the thermopower can be treated as negligible quantities.
Article
A coupled wave analysis is given of the Bragg diffraction of light by thick hologram gratings, which is analogous to Phariseau's treatment of acoustic gratings and to the dynamical theory of X-ray diffraction. The theory remains valid for large diffraction efficiencies where the incident wave is strongly depleted. It is applied to transmission holograms and to reflection holograms. Spatial modulations of both the refractive index and the absorption constant are allowed for. The effects of loss in the grating and of slanted fringes are also considered. Algebraic formulas and their numerical evaluations are given for the diffraction efficiencies and the angular and wavelength sensitivities of the various hologram types.