Anda di halaman 1dari 21

LOW TEMPERATURE FERROELECTRICS: DIELECTRIC NONLINEARITY,

LOSSES, AND PARAMETRIC INTERACTIONS AT ULTRAHIGH FREQUENCIES

I. V. Ivanov, I. M. Buzin,* UDC 537.226.4


G. V. Belokopytov, V. M. Sychev,
and V. F. Chuprakov

CONTENTS
Introduction ............................................................... 684
I. Nonlinear Dielectric Permeability. ..................................... 685
1.1 Determination of the Characteristics of Nonlinear Dielectrics .... 685
1.2. Dielectric Permeability at Low Temperatures and Hysteresis
Phenomena ................................................ ...... 688
1.3. Reversible Characteristics e(E) and ~ Y n ( E ) ...................... 689
;.4. Structural Phase Transitio~ in SrTiO3 and Temperature Dependence
of Bhk I .......................................................... 690
2. Dielectric Losses ...................................................... 692
2.1. Method for Measuring Dielectric Losses ........................... 693
2.2. Dielectric Losses in Strontium Titanate (for T ~ 78~ ........... 693
2.3. Dielectric Losses in Potassium Tantalate ......................... 694
2.4. Dielectric Losses in SrTiO~ in the Temperature Range 4.2-78~ .... 695
2.5. Nonlinearity in Dielectric Losses ................................ 696
3. Parametric Interactions in Distributed Oscillatory Systems Based on
Ferroelectric Crystals ........................................... 698
3.1. Distributed Parametric Systems Based on Ferroelectrics ........... 698
3.2. Theoretical Analysis of Parametric Interactions in Ferroelectric
Resonators ....................................................... 699
3.3. Experimental Investigation of Parametric Effects in Ferroelectric
Resonators ....................................................... 700
Conclusions ................................................................ 701
Literature Cited ........................................................... 702

INTRODUCTION
Substances that have the characteristic features of ferroelectrics, but remain in the
nonpolar phase down to absolute zero, are often called virtual ferroelectrics [I]. Typical
virtual ferroelectrics are strontium titanate and potassium tantalate, which have a structure
that is isomorphic to the nonpolar phase of the classical ferroelectric barium titanate.
The phonon spectra of SrTi03 and KTaO3 exhibit a soft polar mode, whose frequency for the
zero wave vector decreases considerably with deep cooling. Thus, at T = 4.2~ the frequency
of the soft mode in strontium titanate is 15 cm -~, while in potassium tantalate it is 20 cm -I
[2, 3]. A distinguishing feature of these crystals is the fact that condensation of the polar
mode and the ferroelectric phase transition are inhibited by zero-point quantum oscillations
of the crystal lattice. Due to this, the region near the phase transition, where the dielec-
tric permeability is high, while the crystal lattice is very sensitive to an electric field,
mechanical pressure, the presence of impurities and defects, is not several degrees wide, as
in usual ferroelectrics, but extends by tens of degrees into the region of cryogenic temper-
atures. All this not only makes virtual ferroelectrics interesting and convenient objects
for studying the physical properties of nonlinear dielectrics near phase transitions, but it
also makes it possible to create active uhf elements based on them [4]. In oscillatory
systems, constructed with the use of these crystals, it is possible to realize efficient
nonlinear interactions between electromagnetic oscillations and on this basis to create new

*Deceased.

M. V. Lomonosov State University, Moscow. Translated from Izvestiya Vysshikh Uchebnykh


Zavedenii, Fizika, No. 8, pp. 6-27, August, 1981.

684 0038-5697/81/2408-0684507.50 9 1982 Plenum Publishing Corporation


Itilii f "1"-"~--7~ ~"R]

o
C::gJ ulu:"
2,J
2 - 1

1
!
7
~

i I Iliil I I I I I liq I I iiiii I i I ililJ


o
4 5 0 1 0 20 4 8 ~ 4 0 4 /dl# 20 406880
T,~ T,~

Fig. I. The temperature dependence of the dielec-


tric permeability of SrTi03 (a) and KTaO3 (b) for
different bias fields.

practically promising devices: parametric amplifiers, mixers, and so on. It is important,


in so doing, that the fundamental (phonon)'losses in virtual ferroelectrics be small, since
at low temperatures the equilibrium population of excited vibrational levels of the lattice
decreases [I, 5, 6]. Operation of ferroelectric parametric systems at low temperatures is
also favorable for attaining low noise levels.
The considerations listed above determine the continuing interest in virtual ferroelec-
trics. A complete review of the comprehensive studies carried out on this subject in recent
years would be too long. In the present work, we concentrate on a more restricted theme: a
discussion of the results of studies on dielectric nonlinearity and losses in virtual ferro-
electrics at uhf frequencies, as well as parametric interactions in distributed structures,
containing virtual ferroelectrics. In accordance with the present situation, our primary
attention is directed toward experimental results, while the theoretical problems are exam-
ined primarily in terms of their relation to the experimental data.

I. NONLINEAR DIELECTRIC PERMEABILITY

1.1. Determination of the Characteristics of Nonlinear Dielectrics


The study of the dielectric properties of ferroelectrics is very closely related to
determining the vibrational properties of their crystal lattice. In accordance with the
Lyddane--Sachs--Teller equation [I, 7], the dielectric permeability of ferroelectrics E(m = O)
is determined by the characteristic frequencies of transverse mti and longitudinal ~li opti-
cal type phonons:

~(0) = ~ t ~ 1 7 6l I ..-7, (1)

where n is the number of atoms in a unit cell. The dielectric nonlinearity of ferroelectrics
is an indication of the anharmonicity of their lattice vibrations. The microscopic theory
does not yet make it possible to determine the vibrational and dielectric characteristics
starting only from the composition and structure of the ferroelectric. For this reason, the
main source of information on the dielectric permeability and dielectric nonlinearity is
experiments.
The method and technique of experimental investigation of the nonlinear dielectric per-
meability of low-temperature ferroelectrics were developed in [8-12]. They permit simulta-
neous measurement of the dielectric permeability of a ferroelectric specimen at a frequency
of 500 MHz and efficient generation of the second harmonic (I000 MHz) at different values of
the temperature and constant bias electric field (Eo). Here, the coefficient a~Yn(Eo), de-
fined by the relation

E (2~) = - - 1 ~dyn(Eo)[E(~)]~, (2)


2

685
dyn
or , c m ' k V ":t ~ idyn, c m ' k V ' l ~\
t I ! "------1---"'----- I I ~ [

~d ~ ~ k"2~K
~ *re6
Iv4

a b

4~
t4 ~.~ 4ge

0 0
4 # 12 l~ 20
/ 2 '' E,, tiV-em- ZS, kV" em -1

Fig. 2. The first coefficient of dynamic nonlinearity of


SrTiOs (a) and KTa03 (b) as a function of the bias field for
different temperatures.

where E(~) and E(2~) are the amplitudes of the measuring field and the field of the second
harmonic, appears as a characteristic of the dielectric nonlinearity. This is equivalent
to the idea that the dielectric permeability under the action of a high frequency field E(~)
varies as follows:
(E0, E (~)) = ~ (E0)(1 -- ~Yn(E0).E (~) +... ). (3)
A small variation in the bias field also leads to a change in permeability:
St
s (E0 + AE0) = ~ (Eo)(1 - - o, (Eo) Aeo + . . . ), (4)
but due to possible dispersion, the coefficients of dynamic and static nonlinearity (a~Y n a n d
~st) are not equal to one another.
i
Thus, for example, in barium titanate a~Y n ~ 2a~ t [ 1 0 ] .
The experimental details are described in [8-12], as well as in [4]. For this reason,
we will now proceed directly to presenting and discussing the results of the measurements.
The dielectric properties of single-crystal strontium titanate at ultrahigh frequencies
were investigated in [II, 13-16], while those of potassium tafitalate were investigated in
[14, 17-19]. The dielectric permeability and the coefficient of dynamic nonlinearity of
these crystals were measured in the temperature range from 4.2 to 300~ in bias fields from
0 to 25 kV/cm. The characteristic function e(T, Eo) is shown in Fig. I, while ~(T, Eo) is
shown in Fig. 2. The measurements were carried out on specimens of strontium titanate for
three crystallographic orientations with a measuring field E(m) and a bias field Eo, directed
in the specimen either along the crystallographic axis [I00] or [110] or [III] (see Fig. 3).
Only the orientation [I00] has been studied in potassium tantalate.
The values of the dynamic nonlinearity obtained in the experiment are very high. At the
temperature of liquid helium, the maximum values of ~dyn(Eo) are 1.2-1.6 cm/kV for SrTiOs and
i
0.12-0.20 cm/kV for KTa03.
Let us discuss the extent to which the information presented on the dielectric proper-
ties of SrTiOa and KTa03 is sufficient for a physical characterization of their nonlinearity
and for calculating the nonlinear interactions of electromagnetic waves in structures con-
taining ferroelectrics. In formulating the corresponding electrodynamic problems, all infor-
mation concerning the dielectric nonlinear properties of the media is concentrated in the
material equations, where the properties of the nonlinear material are characterized by a
number of coefficients of nonlinear permeability (or nonlinear dielectric susceptibility)
[20, 21]. Each of these coefficients is a tensor and depends on the frequency of the inter-
acting electric fields and on temperature. However, since KTa03 and SrTi03 (above II0~
have cubic symmetry, such a description turns out to be not too complicated. Indeed, in this
case, the relation between the field and polarization can be represented in the form:
~A
e(t) = ~ z (~)( ~ ) P ( ~ ) e~ + z(~)(~,, ~ , ~ ) P ( % ) P ( ~ ) P(%) ~(~'+~~ + .... (5)

686
dyn cm. RV"1

0 4Y 42 ~g 44. 4g
kV 9em-1
Fig. 3. ~dyn as a function of the
bias field for different orientations
of a single crystal of SrTi03 at T =
4.2~

and, in addition, ~(i) becomes a scalar, while ~(3) has only two independent components.
This means, in particular, that in order to determine ~(3), it is enough to carry out mea-
surements on specimens for two orientations.
There is also another circumstance that can be used to simplify (5). It is well estab-
lished that displacive type ferroelectrics in the nonpolar phase do not exhibit dispersion
in a frequency range down to millimeter waves, with the exception of frequencies in the
vicinity of mechanical resonances of the specimen mm [22, 75]. Thus, if the region near
acoustic resonances is neglected and it is assumed that the frequency of the uhf electric
field acting on the crystal is much greater than mm, then in order to describe the electric
nonlinearity, it is sufficient to have only four values of ~(3): X(3)(0, 0, 0); X(3)(0, 0,
~); X(3)(0, ~, ~); and ;(3)(~, ~, ~). It is not difficult to show that two of these, namely,
X(3)(0, 0, m) and X(S)(0, m, ~), are determined from measurements of the reversible charac-
teristics e(Eo) and the coefficient of dynamic nonlinearity, respectively, and, in addition,
it is these values of ;(3) that must be ~Lown in analyzing three-frequency parametric inter-
actions in nonlinear dielectric resonators [23].
For displacive type ferroelectrics, the low-frequency as well as the uhf dielectric
properties can be described using thermodynamics, where the starting point is the represen-
tation of the energy of the crystal as a series in powers of the polarization. Thus, in
crystals with cubic symmetry, it is assumed [24] that

r = r a(p~ + P] + P;) + (1/2)~1 (P~ + P ~ + P~) +


D i D ~ + p l 2p2~ ) H- ....
+ ~2(P~P32 ~, --1--3 (6)
where P1, P i , and P3 a r e t h e components o f t h e p o l a r i z a t i o n . C a l c u l a t i o n s c a r r i e d o u t [11,
16, 23, and 24] p e r m i t e s t a b l i s h i n g a r e l a t i o n s h i p between t he p a r a m e t e r s o f t he t hermo -
dynamic p o t e n t i a l , t he c o e f f i c i e n t s ~ ( 3 ) , and t h e e x p e r i m e n t a l l y d e t e r m i n e d q u a n t i t i e s c (E o ,
T) and aqYn(Eo, T). For t h e c r y s t a l l o g r a p h i c d i r e c t i o n s [100], [110], and [ l l l ] , s u c c e s s i v e
d i f f e r e n t i a t i o n of (6) g i v e s :
E = 2~P + 2BhutP3; (7)
X = 2~ + 6BnKtP~; (8)
~dyn= 12X-2(P) BhxtP, (9)
where X is the inverse dielectric susceptibility (X = 4w/E), while Bhk ~ are linear combina-
tions of ~ and ~2:
B,00 = ~,, 2B,,0 = ~, + ~, 3B~,~ = ~, § (10)

687
TABLE I. Values of the Constants
eL, Tc, and C for Single Crystals
of SrTiO3 and KTa03

Crystal ~g ITc, ~K[fC'10-4,1


*K References

SrTiO3 39 38 8,7 I13, 14, 16]


KTaQ 4,5 4 6,4 [14, 17, 19]

Thus, where thermodynamic theory is valid, dielectric measurements permit finding the values
of the expansion coefficients in (6). In analyzing the dielectric properties of virtual
ferroelectrics, it was established that for these substances most of the qualitative char-
acteristics following from thermodynamics are valid, in particular, the appearance of a
maximum in the function e(T) with application of a bias field and increase in the dynamic
nonlinearity with decreasing temperature with simultaneous shift in the maximum of a~Yn(Eo)
toward weaker fields. Further, a comparis ~ of the dynamic and static nonlinearities shows
that in both SrTiOa and KTaO3 they coincid~ within the limits of error. This also agrees
with thermodynamic theory, since a calculation [23, 25] with the use of well-known data on
elastic and electromechanica! parameters of strontium titanate [47, 48] and potassium tanta-
late [49] showed that the difference between ~st and ~dyn must be less than 4% for SrTiO3 and
i
20% for KTa03 ([I00] orientation).

1.2. Dielectric Permeability at Low Temperatures and


Hysteresis Phenomena
The function e(T) at E = 0 at high temperature satisfies the Curie--Weiss law:
= ~ L + C/(T---:~). (l I)
The values of eL, C, and Tc for SrTiO~ and KTa03 crystals, determined from the results in
[11, 13, 14, 16, 17, and 19], are presented in Table |. They agree well with data obtained
by other workers. Usually in ferroelectrics, the transition to the ferroelectric phase
occurs for T c r ~ T c . * However, in both SrTiO3 and KTaO3, without any external actions, the
ferroelectric phase transition is absent. Below II0~ for SrTiO3, and below 50~ for KTaO3,
the temperature behavior of the dielectric permeability deviates from (11) becoming increa-
singly flat and for T § 0, 3e/~T § 0, as should occur according to the third law of thermo-
dynamics. This behavior of e(T) is a result of quantum stabilization of the paraelectric
phase: the transition of ions at a location with a dipole moment differing from zero is
inhibited by unavoidable zero-point lattice vibrations. A number of works are concerned
with the theoretical description of the low-temperature behavior of the dielectric perme-
ability of virtual ferroelectrics [26-30]. A thorough comparison of the characteristics
[26, 28, 30] with experimental curves e(T) obtained for low-frequency measurements down to
0.03~ is carried out in [31]. At the same time, the function e(T, Eo), obtained in [29]
using a model of two anharmonically coupled oscillators (see Section 1.3), corresponds best
to the experimental data [13-19].
As far as the low-temperature behavior of the dielectric permeability is concerned,
[32, 33, 17], where a maximum was observed in the function e(T) for potassium tan-
talate, indicating a transition in this crystal into the ferroelectric phase below 10-12~
should also be recalled. However, in measurements described in [19, 34, 35], which made use
of more Derfect KTaO3 crystals, the maximum was absent. It appears that the maximum in e(T)
is associated with impurities. Thus, it is shown in [36-38] that small additions of Na or
Li are not distributed in the crystal lattice at locations of K atoms, but form noncentral
impurities, whose ordering induces a ferroelectric phase transition. On the other hand, pure
KTaO3 single crystals must remain in the paraphase down to absolute zero.
The function e(T) for single-crystal SrTiO3 does not have a maximum. At the same time,
the observation of dielectric hysteresis losses in [40, 41] suggested the possibility of a
transition below 20~ at least for part of the SrTi03 crystal, into the ferroelectric phase.
At the same time, the latest works [42, 43] have shown that hysteresis phenomena in SrTi03

*The case Tcr > Tc corresponds to a phase transition of the first kind.

688
a

-2 -I 0 I 2
E, kV" era'*
c(:yn, cm" kV "i 9
I

--LA--I
-z II1 i ,. 2
Jt , kv,em-'

F i g . 4. D e p e n d e n c e o f t h e d i e l e c t r i c permit-
6 i v i t y o f t h e dynamic n o n l i n e a r i t y o f SrTiOs
on t h e d i s t o r t i o n f i e l d for certain cycles of
alteration o f m a g n i t u d e and d i r e c t i o n o f t h e
distortion field.

and KTaOs do not result only from an impurity-induced phase transition. They can also be in-
duced by transport of electric charge in the bulk or near surface layers of the dielectric.
The curves shown in Fig. 4 were observed for crystals investigated in [16, 19]. They show
that after rer~ving the preapplied constant electric bias field the dielectric permeability
of the crystals does not return to its initial value. In this case, the final signal of the
second harmonic remains and disappears only when an electric field Eb with opposite polarity
is applied. After several cycles of changes in the magnitude and sign of the bias field, the
behavior of e(Eo) and ~ Y n ( E o ) e x h i b i t s a characteristic hysteresis form. However, in con-
trast to the coercive field in ferroelectrics, the magnitude of Eb increases with increasing
polarizing field and, in addition, Eb tends to increase and not decrease with increasing tem-
perature. Chemical etching of the specimens, as well as illumination of the specimens,
greatly influences the hysteresis effects. In particular, etching the surface of high qual-
ity SrTiO3 single crystals eliminates hysteresis. This leads to the conclusion that in
perfect SrTiO~ crystals hysteresis is mainly related to the state of the surface. At the
same time, in KTa03 and in SrTiO3 crystals with defects, volume charge, apparently localized
in trapping levels, plays an important role. The nature of these levels is not yet complete-
ly understood. They could be partly related to oxygen vacancies [43, 44].

1.3. Reversible Characteristics e(E) and ~ Y n ( E )


Analysis of the curves e(E, T) and ~ Y n ( E , T) shows that the experimental data can be
approximated by the function [29]
(E, T) = %0 {[(~2 _+ 7a)1,2 + ~lv/a_ff,[(~2 4-73) 1/2- - ~]213__7 } - - I , (12)
where ~ = E/En, q = ( 0 F / T o ) P ( T / 0 F ) -- 1, and t h e c o n s t a n t s r En, OF, and To = Tc depend on
t h e m a t e r i a l , w h i l e t h e f u n c t i o n ~, d e s c r i b i n g t h e t e m p e r a t u r e d e p e n d e n c e o f t h e d i e l e c t r i c
permeability, is calculated in [29]. The f u n c t i o n s ( E ) , g i v e n by ( 1 2 ) , i s e q u i v a l e n t t o ( 7 ) -
(9).

689
TABLE 2. The Values of the Constants Soo, el,
En, and E r for SrTi03 and KTa03 Single Crystals

Crystat ~.10-4 01-, K k v / e m kV~m Re,fences

SrTiO3 0,,23 166 !4 ][17]


0,23 160 14 0,07--0,15" |16]
KTaQ. L,6 117 0,.16 (k&l [14, 17I
1,5 68,5 0~,50 0~,:1 [19]
* [25]

The agreement between the data in [13, 14, |6, 19] and Eq. (]2) greatly improves (pri-
marily for KTaO3) if it is assumed that together with the electric field E, a local, randomly
distributed field Er also acts in the crystal, so that instead of ~2 in (12), (~2 + ~ ) ,
where ~p = Ep/En~ should be used [45]. In polycrystals, the field Er can be related to the
presence of charged defects [45, 46]. However, in strontium titanate and potassium tanta-
late single crystals, it apparently should be assumed to be a semiempirical characteristic
of the nonuniformities in the crystal. In particular, Er depends on the quality of the mate-
rial, which is especially noticeable in KTaO3 specimens.
The parameters Eoo, En, Ep, and eF, obtained by approximating the data in [13, 14, 17,
1 9 ] , are presented in Table 2. It is interesting that the KTa03 lattice turns out
to be more susceptible to the electric fields (since for it En is lower), but it still does
not make it especially sensitive to different types of inhomogeneities that lead to an in-
crease in Er.
The distortion of the reversible characteristics E(Eo) and ~ Y n ( E o ) due to the irrever-
sible phenomena noted above leads to large errors in determining the parameters in the thermo-
dynamic expansion (5). In particular, for this reason, the large spread in the values of BI
and B2, obtained by different methods for crystals with different quality [11, 50, 51] and
noted in [24], was unavoidable. It could have been expected that by avoiding hysteresis it
would be possible to increase the accuracy with which these constants are determined. How-
ever, measurements for high-quality SrTiO~ single crystals without hysteresis yielded results
that clearly contradicted the prerequisites of thermodynamic theory. Calculation of the
parameters Bhk I according to (7)-(9) showed that these coefficients depend on the bias field
(see Fig. 5a). We note that such a paradoxical result was obtained previously while inves-
tigating solid solutions as well (Ba, Sr) TiO3 [52]. For potassium tantalate, calculation
of the coefficients Bhk I according to the reversible characteristics adyn(Eo) and s(Eo) in-
volves large errors due to hysteresis. However, the same tendencies in the characteristic
Bhk I as in SrTiO3 are observed here as well (Fig. 5b).
It is characteristic that the variation in Bhk I is strongest in weak fields, while as
the field increases Bhk I approaches a constant asymptotically. This indicates that the ther-
modynamic theory cannot be improved by taking into account in (6) terms that contain higher
powers of the polarization (sixth and higher). The behavior of Bhk I also cannot be explained
as resulting from correlation effects. Correlations in the polarization must be significant
mainly in the vicinity of a phase transition, while the anomalous behavior of Bhk Z was ob-
served also in SrTi03 and in solid solutions (Ba, Sr)Ti03 many tens of degrees away from the
ferroelectric phase. Taking into account correlation effects in the microscopic theory also
does not lead to the characteristic behavior of Bhk I [53].
It is shown in [16] that the reason for the observed change in the coefficients of the
thermodynamic expansion with the field could be redistribution of the electric field in the
process of polarization of a nonuniform nonlinear dielectric. In this case, a formal calcu-
lation of the parameters Bhk % from the results of dielectric measurements in relatively weak
fields gives high values. In strontium titanate, such a nonuniformity can be the result of
a structural phase transition, which occurs at a temperature of about ll0~

1.4. Structural Phase Transition in SrTi03 and the


Temperature Dependence of Bhk l
As a result of a nonferroelectric phase transition, below ll0~ the SrTi03 lattice under-
goes a tetragonal distortion [54]. In this case, the crystal separates into twinned regions:

690
~x10, cgs lunits

4.2~

_ _ t I I _/~ r r J F

s kV" cm-~
4#1o,~ cgs ~ i ~
~. ~ K T~ Q3
u ....~ 4.2OK

Fig. 5. The field dependence of the coefficients


B1oo for a high-quality single crystal of SrTiOa
(after annealing and etching of the specimen) (a)
and for KTaO3 (b) at T = 4.2~

structural domains with dimensions of the order of 25-40 um [55, 56]. The dielectric char-
acteristics of single-domain strontium titanate are not well known [48, 57]. Thus, accor-
ding to [57], the maximum dielectric permeability of single-domain SrTiOa at 4.2~ equals
23-103 , which is known to be smaller than the values obtained with multidomained specimens
[]4, 15]. Nevertheless, it may be assumed as established that the anisotropy in c at helium
temperatures is very large (~li/~; ~0.25-0.4), which must lead to nonuniformities in the
multidomain specimen.
Strictly speaking, in the tetragonal phase, the expansion (6) is already inadequate for
representing the dielectric properties of strontium titanate and the coefficients Bhk ~ (10)
lose their physical meaning. Nevertheless, it is shown in [25] that they can be used as
average characteristics of multidomain material and, in addition, in the vicinity of ]|0~
the coefficients Bhk Z must undergo a jump. This jump was discovered previously in [39, I]]
(Fig. 6). In the remaining temperature intervals, the temperature dependence of the coeffi-
cients Bhk~(T ) for both strontium titanate and potassium tantalate is weak and does not show
any noticeable anomalies (Fig. 7).
In conclusion, it may be stated that the available information on the dielectric proper-
ties of the virtual ferroelectrics strontium titanate and potassium tantalate is sufficient
for radiophysical applications. From the point of view of the physics of ferroelectrics, it
shows that in general the dielectric permeability and dynamic nonlinearity follow the thermo-
dynamic theory and, in addition, down to the temperature of liquid helium, where the non-
linearity is largest, it is sufficient to take into account terms not exceeding the fourth
power of the polarization in the expansion of the energy in powers of polarization. The
deviations from thermodynamics for presently available quality of crystals are related n o t
to the fine microscopic effects (of the correlation type), but to circumstances that
were not previously considered in sufficient detail. These include: inhomogeneity of the
crystal, the presence of charges connected to impurities in the bulk and surface layers, and
so on. Such phenomena require further study in connection with the importance of improving
the quality of crystals and avoiding irreversible effects.
At the same time~ even for available crystals, it turns out to be possible to obtain
sufficiently precise and reproducible information on the anharmonicity of the lattice,
carrying out measurements in strong bias fields, where saturation of the induced polariza-
tion appears, while the inhomogeneity of the crystal and hysteresis phenomena show up weakly.
The values of the thermodynamic coefficients Bhk Z, as well as BI and ~2, for the crystals
studied in Ill, 13, 14-16] are presented in Table 3.

691
$~exIO~ cgs units
9 g,~,x..,o'~ cg~ units 8 F - ~ , , ~ , , -I

6 -----1 T'" , I "~

O 20 4o gOT, oi<80

~,oo x m ,H cg~i umt~

"t

i I 1, I 0 20 4Q ~o 8@
2 ,YO @0 l&f llO 12C T, ~ T OK -

Fig. 6 Fig. 7
Fig. 6. The temperat -I e dependence of the coeffi-
cients Bhk I in the vicinity of the phase transition
at T = II0~ for different orientations of a SrTiOs
single crystal [16].
Fig. 7. Temperature dependence of the coefficients
Bhk I for SrTi03 (a) and KTa03 (b).

2. DIELECTRIC LOSSES
Dielectric losses in ferroelectrics and, in particular, low-temperature ferroelectrics,
are distinguished by a number of specific features, related to the nature of ferroelectric-
ity itself and the presence of a soft mode. The anomalous low-frequency nature of the soft
o p t i c a l mode and its proximity to the acoustic branches of the phonon spectrum of ferroelec-
tric crystals determine fundamentally the higher level of dielectric losses in ferroelectrics
compared to linear dielectric ionic crystals [58]. Even with small anharmonicity of ferro-
electric crystals, having the usual order of magnitude for other ionic crystals, the proxim-
ity of the optical and acoustical branches of the phonon spectrum leads to intense energy
dissipation due to third, fourth, and higher order multiphonon processes. These processes
determine the lowest level of dielectric losses at ultrahigh frequencies, which is funda-
mental to ferroelectric crystals. Aside from these fundamental losses, numerous other pro-
cesses involving scattering of virtual phonons, excited in the lattice with the action of a
uhf external electric field on the crystal, are characteristic of ferroelectric crystals.
These processes include, primarily, scattering by inhomogeneities in a real crystal lattice,
which could be various types of dislocations, growth defects, impregnation of another phase,
block-separating and twin-separating boundaries, and so on. The possibility of additional
loss mechanisms, related to the electronic subsystem of ferroelectric crystals, also cannot
be excluded. These losses can be relaxational in nature with characteristic times that can-
not be estimated beforehand.
If we are talking about applications, then it should be noted that the efficiency of
nonlinear interactions of electromagnetic oscillations at ultrahigh frequencies in ferro-
electric crystals to a large extent depends on the possibility of decreasing losses to a
level that is close to the fundamental level and, in addition, the fundamental losses must
be quite small. We can hardly count on fundamental losses being small, if we keep in mind
ferroelectrics with high temperature transitions into the unpolarized paraelectric phase.

TABLE 3. Values of the Coefficients Bhk Z,


Bz, and B2 for SrTiOs near a Phase Transi-
tion at T = I I0~ [16]

cgs ' . . . .

I r
110--130 5.6 4,2 3,2 5,6+--0,2 2~25+--0,2
90 4~8 3,8 2,6 4,9+--0,4 1,9 +--0,4.

692
tan ;'~zoJ '
J 1

tan !~ 1,.]4
fd
;: ) I
8 '~-1

4
ii~ I , I ~ r ,~ i. I l ~ I I I I ! IQ zo JQ #~
8o zo@ 1#0 7,~'0 s 2~"o Z#d ~ 0 T, ~ f, GHz
Fig. 8 Fig. 9

Fig. 8. The temperature dependence of tan ~ for single crystal


SrTi03: I) [61]; 2) [59]; 3) [64].
Fig. 9. The frequency dependence of tan ~ for single crystal
SrTiOa at T = 140~ I) [64]; 2) [61].

Only low-temperature virtual ferroelectrics, to all appearances, can have a low level of fun-
damental dielectric losses, since at a temperature close to absolute zero the equilibrium
population of phonon levels is small.

2.1. Method for Measuring Dielectric Losses


In studying dielectric losses in ferroelectrics at ultrahigh frequencies, the dielectric
resonator method gives the most reliable result [4, 22, 59-61]. If the dimensions of the
ferroelectric crystal are comparable to the length of the electromagnetic wave in the dielec-
tric, then this specimen becomes a resonator, in which each type of oscillation is charac-
terized by its own Q factor, Qo. At the same time, in view of the high dielectric perme-
ability almost all of the electromagnetic energy is concentrated in the dielectric. Corre-
spondingly, losses in the surrounding space (in the walls of the waveguide, to radiation) are
negligibly small [60-63]. For this reason, if the dielectric resonator is uniform, then in-
dependent of its shape and type of excited oscillations the tangent of the dielectric losses
of the material is equal to inverse the Q factor:
tan ~ = Qs (13)
Dielectric losses were studied experimentally in several ranges, covering the frequency
interval 0.5-16 GHz. In the long-wavelength region, the dielectric resonators were placed
in a measuring cell as inhomogeneities or an end-load of a coaxial transmission line. In the
centimeter wavelength range, a reflecting measuring scheme was used, while the dielectric
resonators were placed in a rectangular waveguide near the shorting piston. The technique
and experimental details are described in ~etail in [4, 61, 63].

2.2. Dielectric Losses in Strontium Titanate (T > 78~


A large number of works are concerned with studying the dielectric losses in single-
crystal strontium titanate at ultrahigh frequencies for T > 78~ [59, 61, 65]. Figure 8 shows
the experimental curves tan ~(T) for SrTiO3 from [61], as well as from [59, 64]. It should
be noted that in all cases tan ~ < I0 -3 and the behavior of the experimental curves obtained
by different workers is similar. Figure 9 shows the frequency dependence of tan ~ for single-
crystal SrTiOs, measured at a temperature corresponding to a minimum value of tan 6 (T = 140~
see Fig. 8). It is shown in [59] that in this range of temperatures the linearity of tan ~(f)
remains up to a frequency of 72 GHz.
A number of works are concerned with the theory of dielectric losses in ferroelectrics
at relatively high temperatures. Analysis of energy dissipation processes for uhf oscilla-
tions at frequency m in a single crystal having a soft mode mc (~c(T) >>~) leads to the
equation:

tan 6 = ~f/~c,* (14)


where F is the damping constant for the soft mode. A calculation in the region kT > hm~),

*According to estimates, for SrTi03 (hmc/k) ~ 23~ while for KTaO3 (hue/k) e 36~ [I].

693
tan ~ z
tan ;W s
f#8 , , J , i ,

#o

40

2~ r

/0
#
$

4 o,5
g 9 i i I i I , ~ p J I L~ I
4 a" 0 10 Z',*K a 20 .~a /o ~ *K
Fig. 10. Fig. 1I.
Fig. I0. The temperature dependence of tan ~(T) for a
single crystal of KTaOs near the liquid helium tempera-
ture: I) f = 13.6 GHz; 2) f = 5.5 GHz.
Fig. II. The temperature dependence of tan ~ for a
single crystal KTaO3: I) specimen before annealing; 2)
annealing for 40 hours, T = 900~

where ~ ~ (T -- Tc) [5], showed that the function F(T) has the form
F -= ~* + ~*T + X *T 2, (15)
where the first term describes scattering of critical phonons by lattice defects, while the
subsequent terms reflect damping due to three and four phonon processes. Comparison of (14)
and (15) with the experimental results (Figs. 8 and 9) shows that above 78~ dielectric
losses in strontium titanate indeed stem from damping of the soft mode.
A more rigorous microscopic theory [I, 66] has permitted taking into account more exactly
the factors determining the damping of the soft mode and estimating the quantities 8" and y*.
For SrTiO~, it was found that ~* = 3.7-109 deg-l,sec -I, T* = 2 .1"102~ deg -~'sec-1, which
agrees satisfactorily with experiment.
At the same time, the problem of low temperature losses, which is of greatest interest,
has still not been studied theoretically. Measurements carried out recently [67, 70, 19]
indicate the fact that losses in SrTiOs and KTaOs in the temperature range 4.2-78~ behave
very differently.

2.3. Dielectric Losses in Potassium Tantalate


Since in perfect single crystals of potassium tantalate there are no phase transitions
down to liquid helium temperatures, it may be expected that the characteristics of its di-
electric losses at ultrahigh frequencies are most representative for virtual ferroelectrics.
However, up to the present time, there has been no information on dielectric losses in KTaO~.
The only information is a private communication cited in [71], according to which at T = 300~
at a frequency of |0 GHz, tan ~ = 0.9.10 -s.
Since the level of losses can be expected to be small only in a system with few lattice
defects, tan 6 for KTaO3 specimens was measured only for crystals that were carefully pol-
ished and annealed for 40 hours at T = 900~ Dielectric resonators made of annealed potassium
tantalate permitted observing extremely low losses (Fig. I0) [19, 69]. At T = 4.2~ and a
frequency of ]3.6 GHz, tan ~ = 3.6-I0, 5 was obtained, which is the smallest value of tan
ever recorded for ferroelectrics in the uhf range. The behavior of the curve tan ~(T) indi-
cates the fact that for T < 4.2OK dielectric losses must be even smaller.
At room temperature, the maximum value of the Q factor of a KTaOs dielectric resonator

694
TABLE 4. Change in the Q Factor of SrTiOs Dielectric
Resonators after Several 8-Hour Annealing Cycles at
T = 900~ and after Chemical Etching
Number of Q~"10~s
dietec~ic before ]af~rfirst ] after third after fourth ] after
resonator working anneali~ annealing annealing etchi~
25 0,3,t 0(,5,9 0~60 0,67 0,88
34 0~50 0~87 0~95 1,21 2,0
3'2 0,80 1,,1 1,8 2,2 3~0

at a frequency of 12.8 GHz was Qo = 1300 (tan ~ ~ 7"10-4). This quantity coincides with
estimates in [l], tan g = (I-I0).I0 -4, at a frequency of ]0 GHz, indicating the fact that at
least at high temperatures dielectric losses in potassium tantalate are related to scatter-
ing of virtual uhf phonons by thermal vibrations of the crystal lattice. There are no esti-
mates of the possible magnitudes of tan d at 4.2~ in the literature.
A study of the temperature dependence of tan ~(T) has shown that for temperatures T =
30-40~ a clear maximum is observed in the dielectric losses. The maximum in tan ~ also
remains after the crystals are annealed, although it decreases greatly (Fig. II) [70]. It is
well known that in the temperature interval 30-40~ in KTa03 damping of longitudinal ultra-
sonic waves increases [72, 73]. Both effects have, evidently, a common cause: anomalous
proximity of the soft transverse optical and longitudinal acoustical branches of the vibra-
tional spectrum at low temperatures [72-74, 3, 106]. Interconversion of optical and acous-
tical phonons is what leads to the appearance of peaks in the temperature curves of dielec-
tric and ultrasonic losses. The maximum in the curve tan ~(T) is one more indication of the
fact that at T ~ 2 0 - 3 0 ~ the basic mechanism for dielectric losses is associated with the
lattice.

As a result of annealing the crystals, the level of dielectric losses in the entire
temperature range from 4.2 to 300~ decreases. In all probability, this is related to the
decrease in mechanical stresses, as well as to the number of oxygen vacancies and other
defects. The effect of annealing is especially noticeable in the range T < 30~ It is not
yet sufficiently clear as to how much the measured losses in this temperature interval ex-
ceed the level corresponding to the lattice mechanism. Probably, relaxation of volume charge
makes a noticeable contribution to losses at helium temperatures. The relaxational nature of
the losses is suggested, in particular, by the fact that the level of losses in KTaO3, mea-
sured at a frequency of 5.5 GHz at T = 4.2~ turned out to be greater than at the frequency
13.6 GHz (Fig. 10). In this connection, investigations of the frequency dependence of tan
at helium temperatures have special importance.

2__L4. D ie!ectric Losses in SrTiO3 in the Temperature


Interval 4.2-78~
Dielectric losses in strontium titanate were investigated in the temperature interval
4.2-78~ in [43, 59, 67]. It was established that in contrast to KTa03 dielectric losses in
SrTiO3 increase with increasing temperature. However, the quantitative results of the exper-.
iments differed greatly. This, apparently, is related to the fact that the level of dielec-
tric losses in strontium titanate at helium temperatures depends to a large extent on how
the crystals were worked. This dependence is illustrated in Table 4 [67], from which it is
evident that the Q factors of dielectric resonators, cut out of different bulbs, differ greatly
from one another. Annealing and etching the resonators leads to a noticeable increase in
their Q factor. As a result of etching in a boiling 70-80% solution of orthophosphoric acid,
the etched layer is removed from the surface of the crystals by polishing. The large effect
of specimen surfaces with defects is also indicated by the fact that the Q factor of the
resonators is greater at the higher order oscillations than at the fundamental oscillation.
A value of Qo = 4.7"]03 (tan ~ 2 . | ' ] 0 -4) was attained for oscillations corresponding to
f = 906 MHz with the fundamental oscillation frequency of f = 548 MHz.
Figure 12 shows the temperature dependence of Qo for dielectric resonators made of
SrTiOs at different stages of working, measured in the ] GHz and 5 GHz ranges [67]. When
the maximum Q factor of the resonators is attained, Qo remains practically unchanged with

695
tan ~,/O ~
yooo
i i i 9

r * *
28

zaoa f8
[Jd~
I2
., I
1200
7f##J 8 A--!
/
/ i ~
o-J
4 /

Z#O~i ~ I *K f ! l I 1 I
I ; i I I I 0
4 72 7~ 20 2r 2~
2 4 ~"
GHz
Fig. 12 Fig. 13
Fig. 12. The temperature dependence of the character-
istic Q factor of SrTiO3 dielectric resonators: I)
before annealing, f = 5.5 GHz; 2) after annealing, f =
5.5 GHz; 3) after anne ing and etching, f = I GHz.
Fig. 13. The frequency dependence of tan ~ for stron-
tium titanate at T = 4.2~ l) according to [59]; 2)
before annealing; 3) after annealing and etching the
specimens.

temperature in the range T = 4.2-|0~ while with further increase in temperature, it in-
creases rapidly. For annealed specimens, the general nature of the behavior of tan ~(T) is
similar [59]. In unannealed specimens, the general level of dielectric losses is much
greater and, in addition, in some cases near T = 12-15~ as in [43], a maximum is observed
in tan 6. Near this temperature, for reduced SrTi03 crystals, in [76], a large increase in
Ithe damping of ultrasound was also observed, which strongly decreases as a result of anneal-
ing. It is well known that annealing ferroelectric materials of the oxygen octahedral type
in an oxygen atmosphere or in air leads to a decrease in the number of oxygen vacancies [22].
In strontium titanate, as in potassium tantalate, oxygen vacancies play the role of charged
phonon scattering centers.
Figure ]3 shows the values of tan 6 for SrTiO3, measured at different frequencies at
T = 4.2~ The minimum values of tan 6 at frequencies of I and 5.5 GHz were also obtained
after the specimens being studied were subjected to prolonged annealing and chemical etching.
The results of the measurements on the best specimens fall along the line tan d ~ ~. Never-
theless, there is reason to believe that the losses observed up to now up in SrTi03 greatly
exceed the level determined by the fundamental lattice mechanism. A similar assumption fol-
lows from a comparison of the level of losses in strontium titanate and potassium tantalate
at helium temperatures. The much larger losses in SrTi03 could be a result of scattering of
critical phonons by inhomogeneities, related to the structure of multidomain strontium tita-
hate below 110~ Moreover, losses of both strontium titanate and potassium tantalate
exhibit a nonlinear dependence on the amplitude of the uhf field with relatively low field
intensity.

2.5. NonlinearitY of Dielectric Losses


Nonlinear effects are observed at very low values of the uhf power ( p ~ ! MW), when no
changes in the characteristic frequency of the resonator due to heating of the dielectric
should occur [67, 68].
Experiments have shown that in strontium titanate losses increase with increasing uhf
field amplitudes exciting the oscillations. Typical dependences of the Q factor of the
resonators on the power P absorbed in them are shown in Fig. 14. In all cases, the results
of the measurements are approximated well by the equation
tan6= tan 6o -1- A V - P . (16)
The initial level of losses (tan do) strongly depends on the preceding heat treatment and
differs noticeably for separate specimens. At the same time, the quantity A varies from 5"
10 -3 to 2"10 -3 W-I/2. It should be noted that there is a tendency for the magnitude of A to
decrease after annealing.

696
0 ,,l I ! I " I

~w
Fig. 14. Qo of a SrTi03 dielectric
resonator as a function of the ab-
sorbed uhf power (T = 4.2~ f =
0.7 GHz): l) before annealing; 2)
after the first annealing cycle;
3) after the third annealing cycle;
4) after the fourth annealing
cycle; 5) after etching.

The dependence of the Q factor of dielectric resonators on the magnitude of the uhf
power dissipated in them was also studied at te~eratures above 4.2~ In the temperature
interval 4.2-I0~ (16) remains valid and, in addition, A varies little. With a further in-
crease in temperature (up to 26~ the nonlinear losses decease by more than an order of
magnitude.
The dependence of the dielectric losses on the uhf power was also observed in potassium
tantalate single crystals in an experiment at a frequency of 5.5 GHz [19]. However, these
measurements were carried out over a comparatively small range of uhf powers, so that it was
not possible to obtain any quantitative relations in this case. It is, however, r e m a r k ~ l e
that the measurements at frequencies 12-15 GHz [69-70], performed on KTa03 crystals of the
same quality as in [19], did not show any nonlinear losses up to fields of the order of 102
V/cm.
We note that the measurements discussed in Sections 2.3 and 2.4 were carried out at a
low power level, for which there is no nonlinearity in the losses. Two possible mechanisms
for losses in ferroelectrics in the nonpOlar phase are described in the literature. One of
these mechanisms is related to parametric regeneration of thermal acoustic phonons by the uhf
field [77, 78]. The other mechanism-[79], also parametric, consists of the fact that due to
electrostriction in the presence of static electric fields, acoustical oscillations are ex-
cited in the crystal at the same frequency as the uhf field. The excited oscillations pump
the parametric regeneration process of acoustical oscillations at other frequencies. However,
both of these mechanisms appear only at uhf field intensities of the order of 103-104 V/cm.
At the same time, in our e~eriments, the maximum values of the uhf electric field intensity
in dielectric resonators did not exceed 1-10 V/cm. In addition, a characteristic feature is
the behavior of the increment in tan ~(E) (16), which turned out to be proportional to the
first power of the electric field. At the same time, taking into account the central sym-
metry of SrTiO3 and KTa03 for nonlinear lattice absorption, it may be expected that A tan ~
E 2 .

In all probability, the observed characteristics of the dielectric absorption are re-
lated to the presence of oxygen vacancies and volume charge in these crystals. Vacancies in
strontium titanate are traps for charge carriers [44, 80]. The effective field Eeff in a
ferroelectric, acting on the trapped charge, is much greater than the microscopic value Eo.
In order to estimate the order of magnitude of Eeff, it is sufficient to use the Lorentz
equation for the local field:

Eef f = ~ + 4=~3. (] 7)
Choosing P = eE/4~, ~ = 2-104 , we find that the macroscopic field Eo = 1 V / c m corresponds to
Eeff = 104 V/cm. If instead of (17) we use the equations in [81] for the local fields at

697
the points of a Perovskite type lattice, then the values of Eeff will be even higher. A
field of such intensity can cause electrons to hop between traps. It is well known that in
semiconductors at helium temperatures in fields Eeff = 104 V/cm hopping conduction is non-
linear [82, 83]. It may be expected that the hopping mechanism primarily determines low-
temperature dielectric losses, observed in [67, 68].
Another manifestation of the nonlinear nature of losses in SrTi03 is the increase in
tan ~ with a constant bias field Eo applied to the resonator [43, 64, 84, 85]. In contrast
to the nonlinear effects described above, the characteristic tan ~(Eo) was observed not only
at helium temperatures, but also above 78~ [64]. In principle, the behavior of tan 6(Eo) in
displacive ferroelectrics can be a result of restructuring of the vibrational spectrum of the
crystal in a constant electric field [86]. However, the increments to tan 6 obtained in the
experiment are several orders of magnitudes greater than the theoretical estimates [86].
In order to create a constant bias field in a ferroelectrie resonator, metallic elec-
trodes are deposited on its surface. The Q factor of the resonator at helium temperatures,
after deposition of the electrodes, and even in the absence of a bias field, decreases con-
siderably. In addition, the sharp decrease in the Q factor is even observed When supercon-
ducting electrodes, which contribute negligible additional losses due to the skin effect, are
deposited [84, 87]. Somewhat better results were obtained with the use of electrodes, whose
thickness is less than the skin layer at the working frequency. When copper electrodes with
O
~|0 A are deposited, the Q factor of the dielectric resonator at a frequency of 670 GHz, ex-
cited at the lower type of oscillation H11o, decreases by not more than 40%. These facts show
once again the important role played by the state of the surface layers of the ferroelectric
in uhf dissipation processes.
Let us summarize the dielectric losses in virtual ferroelectrics. There are sufficient
reasons for asserting that above 78~ the main losses in SrTi03 and KTa03 single crystals
stem from damping of the soft mode due to the anharmonicity of the crystal lattice. The pic-
ture of the losses at lower temperatures is more complicated and has not been sufficiently
well studied as yet on the experimental and especially on the theoretical level. The prob-
lems concerning the influence of structural domains in SrTi03 on scattering of soft mode
phonons, the mechanism for dielectric losses in surface (near-electrode) layers, the relation
of nonlinearity in losses to relaxation of the usual charge, and the frequency dependence of
the losses are of great current interest. Investigation of these problems will permit de-
creasing further losses in low temperature ferroelectrics. However, even the levels of uhf
losses in SrTi03 single crystals and especially in KTa03 that have been attained are much
less than in high-temperature nonlinear materials, for example, BaTi03 (see Table 5, Section
3.2).

3. PARAMETRIC INTERACTIONS IN DISTRIBUTED OSCILLATORY SYSTEMS


BASED ON FERROELECTRIC CRYSTALS

3.1. Distributed Parametric Systems Based


on Ferroelectrics
The high dielectric nonlinearity together with the low level of dielectric losses in
low-temperature ferroelectrics permit observing large nonlinear effects in oscillatory uhf
systems. Since the dielectric permeability of strontium titanate and potassium tantalate
crystals at helium temperatures is very large, the length of an electromagnetic wave in
these media at frequencies in the centimeter range constitutes small fractions of a milli-
meter. In this connection, the most promising oscillatory structures for observing nonlinear
effects are distributed structures: dielectric resonators or extended traveling wave systems.
The use of nonlinear waveguides with a ferroelectric for parametric amplification was
proposed abroad back in the 1960s [88, 90]. The conditions for parametric amplification in
a waveguide filled with a nonlinear dielectric were analyzed theoretically in [91]. Travel-
ing wave systems have valuable potential qualities: wide amplification band and high sta-
bility. However, for normal operation of such amplifiers, it is necessary to provide for a
dispersion law in the control system that makes it possible for the waves at the pumping,
signal, and combination frequencies to interact efficiently and at the same time eliminates
parasitic interactions. In addition, there arises the problem of matching low-impedance
circuits, containing the ferroelectric, ~with standard uhf networks. Difficulties in solving

698
these problems have not permitted succeeding as yet in creating parametric uhf traveling
wave systems based on ferroelectrics.
Much better results have been obtained in experiments with ferroelectric resonators.
The idea of a resonating parametric amplifier based on a ferroelectric was proposed in [92,
93]. Low.dielectric losses in strontium titanate and potassium tantalate permit making di-
electric resonators out of these crystals with a high Q factor. Their resonant resistance,
in spite of the high dielectric permeability of the crystals, can be high enough that it
would not be difficult to solve the problem of matching and even coupling to external net-
works, greatly exceeding the critical value.
Due to the central symmetry of SrTi03 and KTaO~ crystals, three-frequency interaction
of electromagnetic oscillations in these materials is only possible in the presence of a
constant electric bias field Eo. As a result of parametric interaction of intense pumping
oscillations that modulate the dielectric permeability at frequency up and the weak signal
with frequency Us, the pump energy is converted into oscillations at the signal and combina-
tion frequencies. At the same time, the parametric interaction in the nonlinear resonator
will be efficient only if Up, ~s, and Ux = up -- u s coincide with the characteristic frequen-
cies of the resonator. In other words, the spectrum of the characteristic frequencies must
include the three frequencies u~, ui, and u~, which satisfy the following frequency matching
condition:
ol + m 2 = ms. (18)
C o m p a r a t i v e l y simple d e s i g n methods p e r m i t a t t a i n i n g (18) i n r e a l systems [94, 95, 87].

3.2. Theoretical Analysis of Parametric Interactions


in Ferroelectric Resonators
The characteristic features of the interaction of standing waves in a ferroelectric
resonator were clarified in analyzing the one-dimensional model [95, 96]. A general analysis
of three frequency parametric interaction, taking into account nonlinearity both in the di-
electric permeability and in the losses, as well as the temperature sensitivity of the reso-
nator characteristics, is carried out in [97, 98, 23]. The calculation showed that in the
case of a resonator, the following quantities are the analogs of the coefficient of modula-
tion of the reactive parameter in a lumped system m:
As(x) = [ zDs(x)
A A
(r) rot rot [~Dp(r) Dx(s) (r)] dV, (19)

where Ds(r), Dx(r), and Dp(r) are the normalized spatial distribution of the electric induc-
tion at characteristic frequencies w1, ~2, and u3, ~ = ~-1(Eo), while ~ = 3~ (3)r~0, u, u)"
P(Eo)/(4~)2. * It follows from (19) that the efficiency of parametric energy conversion in
the resonator is determined not only by the dynamic nonlinearity of the material, but also
by the spatial distribution of the working types of oscillations. Together with the require-
ment of frequency matching (18), for parametric interaction, it is important that the condi-
tion for spatial overlapping be satisfied:
As(x) ~ 0. (20)
Estimates show that the coefficients As(x) depend on the dielectric characteristics of the
material as follows:
A(s x) ~ ~,~(x)),~al2 V - m , (21 )
where V i s the volume of the r e s o n a t o r [98, 23].
The possibility of realizing a parametric amplifier in practice is determined primarily
by the pumping power required for parametric self-excitation of the system. According to
[98, 23] for optimal tuning of the resonator the threshold power equals:
~h = (1 + ~p)a mpmgm2(32~phsAxQ~ QiQs)_I, (22)
where Bp is the coupling coefficient for the resonator at the pumping frequency, while Q~,

*The analogy between m and As(x) is not complete. In particular, for nonlinear resonators,
it is impossible to introduce an equivalent volt-coulomb characteristic, as for lumped ele-
ments [23].

699
TABLE 5. Dielectric Nonlinearity and Losses in Ferroelectric
Active Materials
,, j :

Material ture *K
' I([--0
- , 5 GHz) . e *
e/kVll , = .----:=---~----
w(4y") References
BaTiO3 420 0,012~ 10 2,9.10s 0,03 1,0.10 -4 |10, 75]
BK-7 295 0,008 10 2,1.10 a 0,032 1,1.10 - ~ [100, 1Ol]
SrTiO~ 78 8.10 -5 20 1,4.103 0,02 3,4.10 - n I |61, 67, 16]
SrTiQ 4,,2 2.10 -4 0,2 2.104 1,6 2,2-10 -14 /
KTaOs 4,,2 3.10-~ "i" 1,0 4.10 s 0,16 1,6.10-14 [69, 70, 19]

*In the absence of Eo.


+f = 13 GHz.

Q2, and Q3 are the loaded Q factors of the corresponding types of oscillations. Thus, ex-
penditures of energy on parametric regeneration of the resonator can be decreased in three
ways: decreasing losses in the resonator, choosing an active material with a large dynamic
nonlinearity, and, finally, choosing the optimum configuration of the resonator, for which
the integrated coefficients for the inters ~ ion of the working types of oscillations are
maximum.
If it is assumed that the characteri~tlc Q factor of a nonlinear resonator can be re-
duced to a level determined only by volume losses in the dielectric (Q,Q2Q3 ~ tana~), then
it follows from (21) and (22) that the threshold for parametric excitation for resonators
with identical geometry will be determined only by the quality of the material: Pth ~ P,
where
p = [ ~r~(=~Y~2] -I tan ~ 6. (23)
Comparison of the characteristics of different nonlinear dielectrics (see Table 5) shows that
the advantage of resonators based on low-temperature ferroelectrics are indisputable.
Knowing the threshold for parametric self-excitation, as well as the parameters of the
cold resonator (Q factors and coupling coefficients for the working oscillations) permits
determining other characteristics of the resonating parametric amplifier as well [98]. Thus,
the gain of a tuned resonating amplifier in the reflecting regime equals
G ---- (I--7 --~s)=(1~ 7 + ~s) -=, (24)
where 8s is the coupling coefficient for a resonator at the signal frequency, while y is the
regeneration coefficient: Y = (P/Pth)(l + Bs), where P is the pump power.
One of the principle qualities of amplifiers based on low-temperature ferroelectrics is
the low level of their characteristic noise. Thus, the equivalent noise temperature of an
amplifier based on a resonator operating in the degenerate mode with G>> l is Tnois e = T/28s,
where T is the physical temperature of the crystal [99]. We note, however, that this equa-
tion is valid only in the case of equilibrium fluctuations. At the same time, it is well
known that for electric field intensities 103-104 V/cm, intense nonequilibrium fluctuations
can be excited in the ferroelectrics in active elements [102, 103, 79]. Parasitic parametric
interaction of the electromagnetic field with thermal acoustical phonons leads in this case
to a sharp increase in the noise temperature. Thus, the intensity of the pump field in a
resonating parametric amplifier has an upper bound of the order of 102 V/cm.
There is another circumstance that also imposes limitations on the allowable intensity
in the pump oscillations. Heating of the resonator due to absorption of pump energy leads
to detuning of the vibrational system and also to a decrease in the dielectric nonlinearity,
which is not favorable for parametric amplification.

3.3. Experimental Investigation of Parametric Effects in


Ferroelectric Resonators
The first experiments on observing parametric effects in nonlinear resonators were car-
ried out in microstrip structures based on VK-7 ferreceramics [96]. The experiment was car-
ried out at room temperature and, in this case, stabilization of the temperature of the
active element and compensation of heat-induced detuning caused significant difficulties.
The comparatively high level of losses did not permit realizing stable parametric amplifica-
tion in resonators based on the VK-7 ferroceramic.

700
The next natural step was the transition to investigation of the parametric interactions
in nonlinear resonators based on SrTiO3 [84, 87, 104, 14]. Together with a significant gain
in the required pump power, working in liquid nitrogen or liquid helium permits decreasing
the expected noise temperature of the parametric amplifier as well as increasing heat remov-
al and stability of the resonator temperature.
It was possible to observe experimentally parametric amplification and parametric
generation of uhf oscillations in strontium titanate dielectric resonators at T = 4.2~ [84,
I04]. The resonators were cut out of plates of single-crystal SrTi03. Their dimensions
were of the order of 0.5 x 1 x 3 mm. The bias voltage was applied to the metallic electrodes,
which were deposited on the surface of the resonator by burning them in or spray coating.
Pump frequency of the order of I~ GHz was chosen equal to the frequency of the second type of
the oscillation of the resonator f2. Regeneration effects were observed in the quasidegener-
ate reEime (fs = fx = f2/2). The system was tuned by introducing a small inhomogeneity (a
groove at the center of the resonator). Since almost all of the electromagnetic energy is
concentrated in the dielectric, the relative change in the characteristic frequencies of the
resonator as a function of the bias field occurred almost exactly according to the law f
~-i/2(Eo). For this reason, two-frequency matching (f2 = 2fi) was retained with electrical
frequency tuning in the range ]0-12%.
The threshold for parametric self-excitation in [84, 104] was about 2 W, which agrees
reasonably well with theoretical estimates [98, 23]: Pth = 0.6 W. Correspondingly, the
pump field intensity in the amplification regime did not exceed 40 V/cm.
At the same time, the level of losses fixed in the experiments [84, 87, 104] turned out
to be much greater than expected, starting from data on losses in dielectric resonators
without electrodes. A series of SrTi03 resonators with electrodes consisting of normal
(silver, copper) and superconducting (niobium, lead) materials was studied in [87]. The Q
factors of all resonators were in the range 30-260 at frequencies of 0.5-I GHz. Nonlinear
effects, described in Section 3.5, caused additional decrease in the Q factor.
Since the heat capacity and thermal conductivity of SrTi03 at helium temperatures are
not high [105, ;06], dissipation of pump power in the resonators with comparatively low Q
factors led to noticeable heating of the dielectric and detuning of the resonator. As a
result, parametric amplification and parametric generation were observed in [87, 84, 14] only
with pulsed pumping (pulse duration I-I0 ~sec, repetition frequency ]-3 kHz).
Thus, in creating continuously pumped parametric systems, the most urgent problem is
increasing the Q factors of nonlinear resonators. This can be successful if the mechanisms
for losses in the electrodes and the near-electrode layers are better understood and optimum
methods for depositing electrodeN on surface of a ferroelectric are developed. Other possi-
bilities for decreasing the noise level, such as using resonators with point electrodes [84]
or operating in a regime without bias, have not been exhausted. In the latter case, pump
oscillations modulate the dielectric permeability at twice the pump frequency.

CONCLUSIONS
The material presented above, concerning the properties and possible radiophysical
applications of low-temperature ferroelectrics, strontium titanate and potassium tantalate
crystals, can be summarized as follows.
The problem of the dielectric nonlinearity of these crystals can be assumed to be mainly
solved. The mechanism for the nonlinearity is known; successful approaches to the problem
exist: thermodynamic, microscopic, and model approaches; experimental results agree with
theory.
Much less success has been attained up to now in understanding the mechanism of dielec-
tric losses, in particular, in the uhf range, and especially with deep cooling of the non-
linear crystals. Information concerning losses in strontium titanate is relatively complete,
although in this case a number of fundamental problems remain. These include problems re-
lated to the effect of a structural phase transition at ]I0~ volume charge, and near-
electrode phenomena on the losses. Losses in potassium tantaiate have not been studied ade-
quately. It is necessary to study experimentally the frequency and temperature dependences
of losses in KTaO3 at least in the range 1-40 GHz with cooling to 4.2~ and lower. Compari-
son of the properties of KTa03 and SrTiO3 crystals can greatly improve our understanding of
the mechanisms for low-temperature dielectric losses in ferroelectrics.

701
The problem of parametric interactions in nonlinear systems based on low-temperature
ferroelectrics is solved simultaneously with the problem of losses because, as noted in
Section 3, excess losses in these crystals are the main obstacle to creating efficient para-
metric uhf systemsB operating in the continuous regime.

LITERATURE CITED
I. V. G. Vaks, Introduction to the Microscopic Theory of Ferroelectrics [in Russian],
Nauka, Moscow (1973).
2. Y. Yamada and G. Shirane, Phys. Rev., 177, 858 (1969).
3. J. Axe, J. Harada, and G. Shirane, Phys. Rev., BI, 1227 (1970).
4. O. G. Vendik (ed.), Ferroelectrics in uhf Technol---ogy [in Russian], Sovetskoe Radio,
Moscow (1979).
5. B. D. Silverman, Phys. Rev., 125, 1921 (1962).
6. V. Dvorak, Czech. J. Phys., BI7, 726 (1967).
7. R. H. Lyddane, R. G. Sachs, and E. Teller, Phys. Rev., 59, 673 (1941).
8. I. V. Ivanov and N. A. Morozov, Fiz. Tverd. Tela, 7, 3527 (1965).
9. I . V . Ivanov and N. A. Morozov, Fiz. Tverd. Tela, 8, 3218 (1966).
10. I . V . Ivanov and N. A. Morozov, J. Phys. 8oc. Jpn., 28, Suppl. 53 (1970),
II. I. M. Buzin, I. V. Ivanov, E. I. Rukin, and F. V. Chuprakov, Fiz. Tverd. Tela, 14, 2053
(1972).
12. I. M. Buzin, I. V. Ivanov, and V. F. Chuprakov, Vestn. Mosk. Gos. Univ. Ser. Fiz. Astron.,
21, No. 4, 54 (1980).
13. I~. M. Buzin, I. V. Ivanov, and G. V. Belokopytov, Fiz. Tverd. Tela, 18, 1407 (1976).
14. I. V. Ivanov, G. V. Belokopytov, I. M. Buzin, V. M. Sychev, and V. F. Chuprakov, Ferro-
eleetrics, 21, 405 (1978).
15. I. M. Buzin, I. V. Ivanov, N. N. Strepetova, and V. F. Chuprakov in: Abstracts of Reports
at the Ninth All-Union Conference on Ferroelectricity [in Russian], Rostov-on-Don (1979),
Pt. 2, p. 5.
16. G. V. Belokopytov, I. M. Buzin, I. V. Ivanov, N. N. Strepetova, and V. F. Chuprakov,
Fiz. Tverd. Tela (in press).
17. I. M. Buzin, V. F. Chuprakov, and L. N. Yudina, in: Ferro- and Piezoelectric Materials
and Their Application [in Russian], MDNTP, Moscow (1978), p. 119.
18. I. M. Buzin, I. V. Ivanov, N. N. Moiseev, and V. F. Chuprakov in: Abstracts of Ninth All-
Union Conference on Ferroelectricity [in Russian], Rostov-on-Don (1979), Pt. 2, p. 19.
19. I . M . Buzin, I. V. Ivanov, N. N. Moiseev, and V. F. Chuprakov, Fiz. Tverd. Tela, 22,
2057 (1980).
20. N. Bloembergen, Nonlinear Optics (Frontiers in Physics Series, No. 21), W. A. Benjamin,
Reading , Mass. (1965).
21. S. A. Akhmanov and R. V. Khokhlov, Problems in Nonlinear Optics [in Russian], VINITI,
Moscow (1964).
22. E. V. Bursian, Nonlinear Barium Titanate Crystal [in Russian], Nauka, Moscow (1974).
23. G. V. Belokopytov, Candidate's Dissertation [in Russian], Mosk. Gos. Univ., Moscow
(1978).
24. V. Ya. Fritsberg, A. P. Gaevskis, and G. Zh. Grinvald, Izv. Vyssh. Uchebn. Zaved. SSSR,
Fizika, No. I, 103 (1977).
25. G. V. Belokopytov, Dep. VINITI, 01.09.1980, No. 3951-80 Dep.
26. J. H. Barrett, Phys. Rev., 86, I18 (1952).
27. D. E. Khmel'nitskii and V. L. Shneerson, Fiz. Tverd. Tela, 13, 832 (1971).
28. A. S. Chaves, F. C. S. Barreto, and L. A. A. Ribeiro, Phys.-Rev. Lett., 37, 618 (1976).
29. O. G. Vendik, Fiz. Tverd. Tela, 14, 989 (1972).
30. R. Migoni, H. Bilz, and D. Bauerl-~, Phys. Rev. Lett., 37, I155 (1976).
31. K. A. Muller and H. Burkard, Phys. Rev., B19, 3593 (19~-9).
32. J. K. Hulm, B. T. Mattias, and E. A. Long, Phys. Rev., 79, 885 (1950).
33. D. G. Demurov and Yu. N. Venevtsev, Fiz. Tverd. Tela, 13, 669 (1971).
34. S. H. Wemple, Phys. Rev., 137, A1575 (1965).
35. W. R. Abel, Phys. Rev., B4, 2696 (1971).
36. T. G. Davis, Phys. Rev., B5, 2530 (1972).
37. B. E. Bugmeister and M. D. Glinchuk, Fiz. Tverd. Tela, 21, 1263 (1979).
38. I . N . Geifman, A. A. Sytikov, and B. K. Krulikovskii in: Abstracts of Ninth All-Union
Conference on Ferroelectricity [in Russian], Rostov-on-Don (1979), Pt. 2, p. 206.
39. I. M. Buzin, I. V. Ivanov, A. I. Korobov, and V. I. Levtsov, Izv. Akad. Nauk SSSR, Ser.
Fiz., 35, 1865 (1971).

702
40. M. A. Saifi and L. E. Cross, Phys. Rev., B2, 677 (1970).
41. D. Itchner and H. Granicher, Helv. Phys. Acta, 37, 624 (]964).
42. U. T. Hochlir, Ferroelectrics, 7, 237 (1974).
43. Yu. A. Agafonov, O. G. Vendik, ~u. N. Gorin, A. S. Ruban, and V. V. Smirnyi, Izv. Akad.
Nauk SSSR, Set. Fiz., 39, 841 (1975).
44 A. N. Tsikin and N. A. S---hturbina, Fiz. Tverd. Tela, 12, 3353 (1970).
45 O. G. Vendik and A. B. Kozyrev, Fiz. Tverd. Tela, 17, 846 (1975).
46 O. G. Vendik, L. M. Platonova an, Fiz. Tverd. Tela, 13, 1617 (]971).
47 G. Schmidt and E. Hegenbarth, Phys. Stat. Sol., 3, 329 (1963).
48 H. Uwe and T. Sakudo, Phys. Rev., B]3, 271 (1976).
49 H. Uwe and T. Sakudo, J. Phys. Soc. Jpn., 38, 183 (1975).
50. G. Rupprecht, R. O. Bell, and B. D. Silverman, Phys. Rev., 123, 97 (]961).
51. P. A. Fleury and J. M. Worlock, Phys. Rev., 174, 613 (1968).
52. V. Ya. Fritsberg, G. Zh. Grinvald, and A. N. Gaevskis, Uch. Zap. Latv. Gos. Univ., 189,
47 (1974).
53 B. Ya. Balagurov and M. B. Geilikman, Fiz. Tverd. Tela, 14, 2362 (1972).
54 J. C. Slonczewski and H. Thomas, Phys. Rev., B;, 3599 (1970).
55 E. Sawaguchi, A~ Kikuchi, and Y. Kodera, J. Phys. Soc. Jpn., 18, 459 (1963).
56 F. W. Lytle, J. Appl. Phys., 35, 2212 (1964).
57 T. Sakudo and H. Unoki, Phys. Rev. Lett., 26, 851 (1971).
58 V. L. Gurevich, Fiz. Tverd. Tela, 21, 3453 (1979).
59 K. Bethe, Phillips Res. Rept. Suppl. No. 2.
60 R. O. Bell and G. Rupprecht, IRE Trans., MTT-9, 239 (1961).
61 I. M. Buzin, Vestn. Mosk. Gos. Univ. Ser. Fiz. Astron., 18, No. 6, 70 (1977).
62 A. Okaya and L. F. Barash, Proc. IRE, 50, 208] (1962).
63 I. M. Buzin and I. M. Angelov, Prib. Tekh. Eksp., No. 4, 114 (1974).
64 G. Rupprecht and R. O. Bell, Phys. Rev. 125, 1915 (1962).
65 G. Rupprecht and R. O. Bell, in: Physics of Failure in Electronics, M. Goldberg (ed.),
Baltimore (1963).
66. B. Ya. Balagurov, V. G. Vaks, and B. I. Shklovskii, Fiz. Tverd. Tela, 12, 89 (1970).
67. G. V. Belokopytov, I. V. Ivanov, N. N. Moiseev, A. V. Petrov, and V. M. Sychev, in:
Abstracts of Reports at the Ninth All-Union Conference on Ferroelectricity [in Russian],
Rostov-on-Don (1979), Pt. 2, p. 12.
68. V. M. Sychev, G. V. Belokopytov, and I. V. Ivanov, Vestn. Mosk. Gos. Univ. Ser. Fiz.
Astron., 21, No. 3, 78 (1980).
69. I. M. Buzin, I. V. Ivanov, V. A. Chistyaev, and V. F. Chuprakov, Pis'ma Zh. Tekh. Fiz.,
6, 457 (1980).
70. T. M. Buzin, I. V. Ivanov, and V. A. Chistyaev, Fiz. Tverd. Tela, 22, 2848 (1980).
7]. R. C. Miller and W. G. Spitzer, Phys. Rev., 129, 94 (1963).
72. H. H. Barrett, Phys. Rev., 178, 743 (1969).
73. G. Barrett, in: Physical Acoustics, Volume 6, W. Mason and R. Thurston (eds.), Academic,
New York (1970).
74. G. Shirane, R. Nathans, and V. Minkiewicz, Phys. Rev., ]57, 396 (1967).
75. Yu. M. Poplavko, Izv. Akad. Nauk SSSR, Ser. Fiz., 34, 2572 (]970).
76. C. K. Jones and J. K. Hulm, Phys. Lett., A26, 182 (1968).
77. O. G. Vendik, L. M. Platonova, and A. I. S0kolov, Izv. Akad. Nauk SSSR, Ser. Fiz., 33,
1167 ( 1 9 6 9 ) .
78. O. G. Vendik, L. M. Platonova, and A. I. Sokolov, Fiz. Tverd. Tela, II, 808 (1969).
79. A. M. Prudan, Candidate's Dissertation [in Russian], LETI, Leningrad--(1977).
80. O. I. Prokopalo, Izv. Akad. Nauk SSSR, Ser. Fiz. 39, 995 (1975).
81. J. C. Slater, Phys. Rev., 78, 748 (1950).
82. M. Morgan and P. A. Walley, Phil. Mag., 23, 661 (1971).
83. B. I. Shklovskii, Fiz. Tekh. Poluprovodn-~j 6, 2335 (1972).
84. I. V. Ivanov, G. V. Belokopytov~ and V. M. Sychev, in: Ferro- and Piezoelectric Mate-
rials and Their Application [in Russian], MI)NTP (1978), p. 82.
85. Yu. N. Gorin, Yu. D. Novinskii, and A. S. Ruban, Izv. LETI, 212, 85 (1977).
86. A. K. Tagantsev, Zh. Eksp. Teor. Fiz., 77, 1993 (1979).
87. I. V. Ivanov, G. V. Belokopytov, and V. M. Sychev, Vestn. Mosk. Gos. Univ., Sero Fiz.
Astron., 17, 753 (1976).
88. E. S. Cassedy, Proc. IRE, 47, 1374 (1959).
89. R. W. Landauer, IBM J. Res. Dev., 4, 391 (1960).

703
90. Y. Aoki, IRE Trans., MTT-8, 465 (1960).
91. Yu. K. Barsukov, Radiotekh. Elektron., 20, 1379 (1975); ibid., 1592 (1975).
92. T. R. Billetter, A. J. Giarola, and J. L. Bjorkstam, J. Appl. Phys., 35, 2159 (1964).
93. I. V. Ivanov, Vopr. Radioelektron., Ser. III (Radiodetali i Komponenty Apparatury),
3, 85 (1965).
94. T. V. Ivanov and I. M. Angelov, Inventor's Certificate No. 403,064, June 23, 1972,
Byull. Izobr., No. 42 (1973).
95. I. V. Ivanov, I. M. Angelov, and A. G. Laptev, Izv. Vyssh. Uchebn. Zaved., Radioelek-
tron., 17, No. I0, 28 (1973).
96. I. V. Ivanov, Vestn. Mosk. Gos. Univ., Ser. Fiz. Astron., 14, 501 (1973).
97. G. V. Belokopytov, Vestn. Mosk. Gos. Univ., Ser. Fiz. Astron., 18, No. 2, 61 (1977).
98. G. V. Belokopytov, Vestn. Mosk. Gos. Univ., Ser. Fiz. Astron., 18, No. 5, 103 (1977).
99. M. N. Malyshev, Candidate's Dissertation, LETI, Leningrad (1979).
100. T. N. Verbitskaya, I. V. Ivanov, and N. A. Morozov, Fiz. Tverd. Tela, 12, 1578 (1970).
I01. Yu. M. Poplavko, V. G. Tsykalov, and V. I. Molchanov, Fiz. Tverd. Tela, IO, 3425 (1968).
102. O. G. Vendik, A. A. Dakhnovich, A. S. Ruban, L. T. Ter-Martirosyan, and Yu. F. Yan-
chenko, Radiotekh. Elektron., 17, 1981 (1977).
103. L. T. Ter-Martirosyan, Radiotekh. Ele ron., 20, 2592 (1975).
104. I. V. Ivanov, G. V. Belokopytov, and V. M. Sych---ev, Pis'ma Zh. Tekh. Fiz., ~, I011 (1977).
105. Y. Suemune, J. Phys. Soc. Jpn., 20, I ~ (1965).
106. E. Steigmeier, Phys. Rev., j68, 523 (]968).

WAVES IN THIN SEMICONDUCTOR LAYERS WITH NEG~flVE


DIFFERENTIAL CONDUCTIVITY

A. A. Barybin and V. M. Prigorovskii UDC 621.375.029.64

CONTENTS
Introduction ............................................................... 704
I. Basic Equations and Boundary Conditions ................................ 706
2. Peculiarities of Wave Propagation in Semiconductor Structures with
Negative Differential Conductivity (NDC) ............................... 707
Thin Semiconductor Layers with Longitudinal Drift ...................... 708
Thin Semiconductor Layers with Transverse Drift ........................ 711
3. Experimental Results on Amplification Distribution in Semiconductors
with Negative Differential Conductivity ................................ 714
Traveling Wave Amplifiers with Longitudinal Drift ...................... 714
Traveling Wave Amplifiers with Transverse Drift ........................ 715
Conclusion ................................................................. 716
Literature Cited ........................................................... 716

INTRODUCTION
Electromagnetic wave propagation in unbounded plasma media has now been studied quite
fully [l-]O]. The possibility of creating solid-state uhf devices with distributed interac-
tion has caused interest in the study of wave processes in multilayer structures containing
thin semiconductor films with drifting charge carriers. These films (when connected to
auxilliary waveguide systems) play an active role, which is manifested in convective insta-
bility (spatial amplification) of transverse waves: retarded electromagnetic [I, II~14],
acoustic [I, 15-20], magnetostatic [I, 21-25], and magnetoelastic [26, 27], due to interac-
tion with natural charge carrier drift waves in the semiconductor.
Of special interest is wave propagation in semiconductor films with negative differen-
tial conductivity (NDC), which develops, for example, in intense electric fields because of
the intervalley transfer effect in n-GaAs type materials. We may distinguish in principle

V. I. Ul'yanov (Lenin) Leningrad Electrotechnical Institute. Translated from Izvestiya


Vysshikh Uchebnykh Zavedenii, Fizika, No. 8, pp. 28-41, August, 1981.

704 0038-5697/81/2408-0704507.50 9 1982 Plenum Publishing Corporation

Anda mungkin juga menyukai