Вы находитесь на странице: 1из 8

Materials Science and Engineering B 134 (2006) 1–8

Review

Electrical conductivity and dielectric properties of some


vanadium–strontium–iron unconventional oxide glasses
F. Abdel-Wahab a , M.S. Aziz b , A.G. Mostafa c , E.M. Ahmed b,∗
aPhysics Department, Faculty of Science in Aswan, South Valley University, Aswan, Egypt
b Physics Department, Faculty of Science in Damyetta, Mansoura University, Damyetta, Egypt
c Physics Department, Faculty of Science, AL-Azhar University, Nasr City, Cairo, Egypt

Received 4 February 2006; received in revised form 2 July 2006; accepted 23 July 2006

Abstract
An unconventional oxide glass system of the composition (80 − x)%V2 O5 ·x%SrO·20%FeO (0 ≤ x ≤ 40) was prepared by the press-quenching
technique. The electrical conductivity (σ dc and σ ac ) was measured in the temperature range from 308 to 588 K. σ dc was found to follow a mixed
conduction of both small polaron hopping (SPH) and variable range hopping (VRH) mechanisms up to 418 K. Above this temperature σ dc follows
SPH conduction mechanism only. σ ac was determined at four frequencies (102 , 103 , 104 , 105 Hz). It was found that the correlated barrier hopping
(CBH) was the applicable mechanism, and the exponent factor (s) varied from 0.1 to 0.83. The theoretical fitting between the proposed models
and the experimental data showed good agreement. σ dc and σ ac were found to decrease with the increase of SrO content, and both the dielectric
constant ε and loss ε were found to increase with temperature and decrease with the frequency.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Unconventional oxide glasses; Iron–vanadate glasses; Electrical and dielectric properties

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2. Experimental details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1. Sample preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2. Electrical conductivity measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3. Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3.1. dc Conductivity and activation energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3.2. Relation between the activation energy and the mean V-ion spacing (R) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.3. Oxygen anions density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.4. ac Conductivity and dielectric properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
4. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1. Introduction their important application [1–6]. Such glasses exhibit also many
interesting physical properties such as switching phenomenon,
Oxide glasses containing transition metal ions (TMI) are of electrochromic properties, etc. [7–12].
special interest because of their semiconducting behavior and In glasses containing vanadium both V4+ and V5+ ions are
presented, and the electrical conduction results by the hopping
of the unpaired 3d1 electron between V4+ and V5+ states [5,6].
∗ Corresponding author. Tel.: +20 57 403980; fax: +20 57 403868.
The unpaired electrons induce polarization around vanadium
E-mail addresses: elagwany75@mans.edu.eg, ions and hence, form polarons. Also in iron containing glass
elagwany75@yahoo.com (E.M. Ahmed). both Fe+2 and Fe+3 ions are present and the electrical conduction

0921-5107/$ – see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.mseb.2006.07.025
2 F. Abdel-Wahab et al. / Materials Science and Engineering B 134 (2006) 1–8

occurs due to the hopping of the unpaired 3d5 electron between frequencies [102 , 103 , 104 , 105 Hz], and all measurements were
Fe2+ and Fe3+ sites forming another polaron. performed in the temperature range from 308 to 588 K.
The dominant charge carrier in such glasses may be elec-
trons, ions, polarons and/or protons [13,14]. All these types 3. Results and discussion
produce the same behavior when dc conductivity measurement
is considered. In addition, the effect of pilling-up of ions by 3.1. dc Conductivity and activation energy
electrodes (polarization) makes it impossible to differentiate
between various charge carriers [15,16]. Therefore, ac conduc- Fig. 1 shows the relation between σ dc and [103 /T] for all
tivity measurements are usually used to determine the type of samples. It appeared that these glasses have dc conductivities
the dominant carriers and to avoid the effect of electrode polar- in the range from 10−10 to 10−3 S cm−1 . The general behav-
ization. ior of the conductivity is the presence of two regions in [σ dc
It is suggested that electrical transport in TM doped oxide versus 103 /T] relation, one at relatively low temperature, while
glasses may occur by small polaron hopping between two differ- the other appears at high temperature. The change in these two
ent valency states of the TM ions [1,17,18], or by multiphonon regions is linear and these two straight lines intersected at def-
hopping of localized electrons with a weak electron–phonon inite temperature (Tx ) differ from one sample to another. It is
interaction [19]. There exists also a controversy over the nature supposed that at such definite temperature there may be a change
of hopping mechanism in different temperatures and composi- in the role played by strontium. According to Hirashima,
tion regions [20,21]. In addition some iron vanadium strontium
Tx = θD /2,
oxide glasses exhibit an interesting switching properties [22],
and hence additional work in this field have to be done. where (θ D ) is the characteristic Debye temperature that defined
However, this article aims to investigate the electrical con- by [23],
ductivity (dc and ac) and the dielectric relaxation properties of
V2 O5 ·SrO·FeO glass system as well as to clarify the type of hυ0 = kθD
charge carriers and the dominant conduction mechanism. where h is Planck’s constant, υ0 the optical phonon frequency
and k is Boltzmann constant. The obtained θ D and υ0 values
2. Experimental details are listed in Table 1. The characteristic Debye temperature of
the Sr-free sample was about 667.6 K. The introduction of only
2.1. Sample preparation 5 mol% SrO increases such temperature to 689.7 K. As SrO was
gradually increased from 5 up to 15 mol%, Debye temperature
A glass system of the composition: (80 − x) V2 O5 ·(x) SrO·20
show some fluctuations between the value 671.3 for sample
FeO, was prepared on the bases of the percentage molecular
weights, within the glass formation limits 0 ≤ x ≤ 40. Aldrich
and BDH companies supply all the used chemicals with purity
not less than 99.98%. Vanadium pentoxide and strontium oxide
were added as such while iron oxide was added as anhydrous
ferrous oxalate in order to achieve chemically reducing medium.
The finely mixed batches were melted in porcelain crucibles in
an electric muffle furnace for 90 min, at temperatures ranging
between 850 and 950 ◦ C depending on the batch composition.
The melts were stirred several times during melting, and they
were then poured between two pre-cooled stainless steel plates
in air. The obtained solid samples were directly transferred to the
annealing furnace at 200 ◦ C, which was then left to cooled very
slowly with a decreasing temperature rate of about 0.35 K/min
in order to obtain strain free samples.

2.2. Electrical conductivity measurement

The annealed glasses were then polished to obtain disk


shape samples of constant diameter (≈0.8 cm) and thickness
(≈0.1 cm). The used electrodes were obtained by evaporating
50 nm thickness aluminium metal on both sides of the samples.
The I–V characteristics were carried out to check the ohmic
behavior of these samples up to 100 dc volts. The dc conductivity
measurements were carried out using 617 Keithly electrometer,
while the ac measurements were carried out using a computer Fig. 1. The Ln (σ dc ) vs. [1000/T] for (80 − x)%V2 O5 − x%SrO − 20%FeO glass
controlled Stanford LCR bridge model SR 720 at four fixed system.
F. Abdel-Wahab et al. / Materials Science and Engineering B 134 (2006) 1–8 3

containing 10 mol% SrO, and the value 676.5 K for the sample

3.190614
3.291952
3.383693
3.468815
3.574794
3.706739
3.854209
4.008922
4.203523
Density
(g/cm3 )
that contains 15 mol% SrO. Finally, Debye temperature show
a gradual decrease as the SrO content was increased up to the
sample that contains 40 mol% SrO. The high difference in θ D

(1022 cm−3 )
between the Sr free sample and those containing SrO may be due

4.601148
NTM-ions

4.76069
4.69955
4.71331
4.61678
4.55404

4.56427
to the opening of the glass network structure produced by the

4.5352
4.6609
large Sr2+ volume, which increases the thermal vibration of the
network. This in turn act directly to increase the optical phonon
frequency and the vibrational amplitude of the glass network.
(1022 cm−3 )

This was also noticed obviously as Sr2+ cations were just intro-
1.69033
1.62275
1.57794
1.49314
1.41674
1.35009

1.22865
1.13808
1.3193
NO-ions

duced into the glass network where υ0 showed jump increase


from 1.39 × 10−13 up to 1.58 × 10−13 s−1 [for Sr free sample
and that contains only 5 mol% SrO, respectively.
(1022 cm−3 )

The dc conduction mechanism can be recognized by the


2.17047
2.14857
2.07465
2.01447
1.97148
1.98724
1.91982
1.85832 investigation of the following three expressions:
2.2252
NV-ions

1- The thermal activation process, which is applied to treat data


above room temperature giving rise to single straight line.
This process follows an Arrhenius relation:
εp × 10−12

 
at 503 K

E
σ(T ) = σ0 exp − (1)
4.25
3.76
3.33
3.17
2.83
2.63
2.29
2.43
2.08

kT
where σ 0 is a pre-exponential value, E is the activation
energy and T is the absolute temperature. This case cannot
εp × 10−12

be applied on the data in this article.


at 313 K

2- The small polaron hopping (SPH) mechanism [24] which


4.90
4.14
3.60
3.37
3.17
2.97
2.57
2.70
2.22

is considered as a modification of the previous mechanism.


Such model obeys the following equation:
 
E
σ(T ) = (σ0 /T )exp − (2)
rp (nm)

kT
0.157
0.159
0.161
0.164
0.167
0.169
0.171
0.175
0.179

where,
υ0 NC(1 − C)e2 R2
σ0 =
(1013 s−1 )

exp(−2αR)
k
1.39
1.44
1.40
1.41
1.35
1.32
1.29
1.28
1.26

[in non adiabatic region]


υ0
Chemical compositions and physical properties of V2 O5 –SrO–FeO glasses

and
υ0 NC(1 − C)e2 R2
σ0 =
θ D (K)

667.6
689.7
671.3
676.5
648.2
632.8
620.1
613.4
604.9

k
[in adiabatic region that is exp(−2aR) = 1]
where N is the transition metal ion density, C the fraction
FeO

of the reduced transition metal ion, R the average spacing


20

between any two transition metal ions, e the electronic charge


and α is the tunneling factor. This model presents the con-
ductivity in the form of two intersected straight lines with
SrO

0
5
10
15
20
25
30
35
40

two different activation energies.


composition

3- The variable range hopping (VRH) model, in which σ (T) was


presented as a function of T−1/4 according to the following
V2 O5
Glass

equation [5,15]:
80
75
70
65
60
55
50
45
40

σ(T ) = σ0 exp[−(T0 /T )1/4 ] (3)


where T0 is a characteristic temperature. Eq. (3) becomes appli-
sympole
Table 1

cable only if the conductivity curve shows the nonlinear behavior


Glass

S1
S2
S3
S4
S5
S6
S7
S8
S9

below room temperature.


4 F. Abdel-Wahab et al. / Materials Science and Engineering B 134 (2006) 1–8

ual increase of SrO at the expence of V2 O5 , makes the glass


structure becomes more disordered and appears in the form of
loosely backed structure [28]. On the other hand, it can be seen
from Fig. 1 that the onset of the lower activation energy region
(II) increases with the increase in the V2 O5 concentration, and
this in turn indicates that there is a sort of hopping of electrons
that moves between V-ions using lower activation energy. Also,
Hirashima et al. [29] suggested that in iron vanadate glass sys-
tems, the higher the amounts of iron content the lower is the
activation energy which means that there is another sort of hop-
ping takes place in which some electrons hop between V and Fe
ions in addition to that appeared between V-ions and between
Fe-ions separately.
Therefore, with the greater tendency to form polarons
between V+4 and V+5 as well as between Fe+2 and Fe+3 , it
is assumed that the conduction at high temperature range per-
formed via SPH model between different V and different Fe ions
in addition to that between V and Fe ions. Whereas the conduc-
tion at low temperature range performed by mixed conduction
mechanisms of both SPH and VRH models.
Fig. 3 shows the relation between Ln (σ dc ) and SrO con-
tent at three different temperatures, where it appeared that σ dc
decreases with the gradual increase of SrO content. It can be sup-
posed that the number electron hopping processes in the glass
network decreased due to the stopping of electrons by Sr ions that
act as obstacles. Also as strontium ions increased, it enters the
Fig. 2. The variation of the activation energy E with V2 O5 content at different interstitial vacancies to occupy the modifier positions replacing
temperatures. The solid lines are drawn as a guide to the eye. some iron cations that firstly act as modifiers, while vanadium
and iron occupy now the network former positions. Then the
Inspecting the above three models and considering the data
that presented in Table 1, it can be concluded that the SPH model
is the most dominant applicable mechanism for electronic trans-
port in the studied glasses.
Austin and Mott [6] assumed that a strong electron phonon
interaction exists at high temperature range. The activation
energy (E) is the energy resulted during polaron formation
with binding energy EH and the energy difference ED . Such
polaron may exist between the initial and final sites due to vari-
ations in the total arrangement of ions [24]. They also suggested
an expression for activation energy according to SPH model as:

E = EH + ED /2 for T > Tx

E = EH for T < Tx /2
The obtained values for the activation energy E in the two
temperature regions (low (II) and high (I) temperatures) are pre-
sented in Fig. 2, as a function of SrO content. It can be seen that in
the two temperatures E increases with the gradual increase of
SrO concentration. Also it is observed that activation energies at
high temperatures (I) are higher than the corresponding value at
low temperatures (II). Amano et al. [25] reported that the lower
activation energy at low temperatures may be due to the pres-
ence of another conduction mechanism with the (SPH) [mixed
conduction] that is both (SPH) and (VRH) could be applied
in this range of temperature, since the (VRH) model exhibits Fig. 3. Compositional dependence of dc conductivity at different temperatures.
lower activation energy than (SPH) model [25–28]. The grad- The solid lines are drawn as a guide to the eye.
F. Abdel-Wahab et al. / Materials Science and Engineering B 134 (2006) 1–8 5

conduction part due to the movement of some ions is stopped 0.3 eV [34]. This is due to the fact that the hopping energy EH
due to the blocking effect of Sr ions. Since Sr ions are of rela- depends on the dielectric constant (εp ) of a glass sample [35].
tively large ionic radius and high atomic weight in comparison The dielectric constant εp can be obtained from the equation.
with the usual alkali cations (Li, Na and K ions). So it is hard for  
Sr+ to move throughout the network and hence it adds no part ED e2 1 1
EH ≈ E = = − (6)
to the conductivity of these glasses. 2 4εp rp R
where
3.2. Relation between the activation energy and the mean
V-ion spacing (R) ε−1 −1 −1
p = ε∞ − εs

ε∞ and εs are the high frequency and static dielectrics, respec-


The activation energies (E) of conduction showed a
tively. The values of εp that obtained at two constant tempera-
strong dependence on the mean V-ion spacing (R) for a
tures (313 and 508 K) are listed in Table 1. From this table, it
ternary glass containing fixed concentration of the third compo-
appeared that εp decreases with the decrease of V2 O5 content
nent. The glass systems (xV2 O5 − 10SnO − (90 − x)TeO2 ) and
as well as with the increase in temperature.
(xV2 O5 − 20ZnO − (80 − x)TeO2 ) have been studied by Mori
The (E − R) relation for SPH conduction can be expressed
et al. [30,31]. The drown results showed that the electrical con-
according to Eq. (7) [32]:
duction is due mostly to SPH mechanism between the various
valance states of V-ions [9,30,31]. E = Eo + a(R − Ro ) (7)
In order to confirm the [E − R] relation in the present study,
the V-ion density (NV-ions ), was calculated using the following where a is constant, and it was found to be 1.7 and 1.4 ev nm−1
equation [32]. for the glass systems (V2 O5 − MnO − (80 − x)TeO2 ) [24] and
(xV2 O5 − 20Sb − (80 − x)TeO2 ) [23], respectively. However, a
NV−ions = 2[(dWtV2 O5 /MWV2 O5 )NA ] (4) was calculated where it was found to be 2.6 eV nm−1 for R
between 0.38 and 0.45 nm. Using the mean spacing between
where d is the density, WtV2 O5 the weight percentage of V2 O5 ,
V-ions (R) calculated from Eq. (5), the polaron radius rp is cal-
MWV2 O5 the molecular weight of V2 O5 and NA is Avogadro’s
culated according to the following Eq. [36]:
number. The obtained values are listed in Table 1. The mean
spacing between any two V-ions (R) was also calculated [32], 1  π 1/3
rp = R (8)
from the relationship 2 6
R = (1/NV−ions )1/3 (5) where rp is obtained to be in between 0.157 and 0.179 nm. The
calculated values of NV-ions , R and rp are exhibited in Table 1, and
Fig. 4 shows the relation between E and R for the present the obtained values of NV-ions and R were found to be very close
glass system, where it appeared that the activation energy E to those calculated by Sen and Ghosh [4,17] for the ((100 − x)
increased with the increase of R. These results were found to be SrO − x V2 O5 ) glass system.
in agreement with those obtained previously for V2 O5 –Z–TeO2
glass systems (Z = SnO [30], ZnO [31] or Sb2 O3 [33]). All the 3.3. Oxygen anions density
previous results as well as the results obtained in this article con-
firmed the dependence of the activation energy on the V–O–V Hirashima et al. [23] reported that the distance between V-
site distance. For V2 O5 containing glass, E is higher than ions of the linear bond, V–O–V in the system PbO–P2 O5 –V2 O5 ,
was estimated to be about 0.39 nm using the ionic radii of
O2− (=0.14 nm) and V5+ (=0.054 nm) [37] in crystalline oxide
form. The O-ion density can be calculated using the following
equation [32]:
NO-ion = 5[(d WtV2 O5 /MWV2 O5 )NA ]
+ [(d WtSrO /MWSrO )NA ]
+ [(d WtFeO /MWFeO )NA ] (9)
where Wt is the weight fraction and MW is the molar weight
of an oxide in a given sample. Fig. 5 represents the effect of
Sr ion concentration on the O-ion density, where it was found
that NO-ion decreases with the gradual increase of Sr cation con-
centration. This indicated that there is a gradual change in the
structure of the glass network as SrO was gradually increased,
due to the gradual decrease in the O-ion density. Thus, these
Fig. 4. The effect of mean vanadium ion spacing (R) on the activation energy results main that polarons become able to hop easily between
E for conduction. V-ions as SrO was decreased. This because Sr ions may act
6 F. Abdel-Wahab et al. / Materials Science and Engineering B 134 (2006) 1–8

dependence but it is almost frequency independent. The temper-


ature and frequency dependences of the ac conductivity for all
other glass samples show similar behavior.
The experimental values of the frequency exponenet factor (s)
as a function of temperature are presented in Fig. 7 for the glass
samples containing 5, 30 and 40 mol% SrO as representative
graphs. It is observed from this figure that s decreases with the
gradual increase of temperature.
It is known that the frequency-dependent ac conductivity
σ(ω) in all amorphous solids and glasses [24] follows the expo-
nential power low:

σ(ω) = Aωs (10)

where A is a pre-exponential constant and s is the frequency


exponent factor. The experimental value of the exponent s can
Fig. 5. Effect of the SrO concentration on O-ion. be obtained as.
d(ln σ)
s= (11)
as bridges between some vanadium ions in the glass network d(ln ω)
and hence increases the polaron hopping distances. Also, the
increase of Sr ions (higher ionic radius) at the expense of V The behavior of the exponent factor s as a function of tem-
ions (lower ionic radius) as well as the gradual decrease of perature can be used to determine the origin of the conduc-
oxygen act to compact gradually the glass network structure tion mechanism. Several models [38–41] based on the quantum
which in turn act to decrease the electronic transport in these mechanical tunneling and the classical hopping of charge carri-
glasses. ers over barrier has been proposed. A model based on quantum
mechanical tunneling of electrons through a barrier predicts that
the exponent s is a temperature-independent. But the experimen-
3.4. ac Conductivity and dielectric properties tally obtained data showed temperature dependent behavior for
all glasses. Thus, this model is not applicable for the present
Fig. 6 shows the temperature dependence ac conductivity for glass system. Another model [39,40] based on the classical hop-
the glass sample of the composition [50% V2 O5 ·30% SrO·20% ping of charge carriers over barrier predicts a decrease in the
FeO] at four different frequencies (open shapes). The change value of s with the increase of temperature. Such model was
in the dc conductivity is also included in the same figure (solid found in complete agreement with the obtained data. Accord-
circles) for comparison. It is clear from this figure that at low tem- ingly, it represents the most applicable model for the conduction
peratures the ac conductivity is considerably higher than the dc mechanism in the studied glasses. This model is the corre-
conductivity and both dc and ac changes shows extremely weak lated barrier hopping (CBH) that proposed by Elliott [38]. It
temperature dependence and strong frequency dependence. At has been applied firstly to investigate the transport properties
high temperatures the ac conductivity shows strong temperature in some chalcogenide semiconducting glasses and then it used
successfully to describe the conduction mechanism in many

Fig. 6. Frequency dependent conductivity as a function of the inverse tempera-


ture. Experimental data are presented as dots and the fitting using (CBH) model
are shown as solid lines; dc conductivity is shown for comparison. Fig. 7. Dependence of the exponent factor s on the temperature.
F. Abdel-Wahab et al. / Materials Science and Engineering B 134 (2006) 1–8 7

oxide glasses. Such model is based on the hopping of bipo-


laron over barrier (i.e. two electrons or polarons hop between
charged defects D+ and D− ) and it has been proposed to explain
the frequency dependence conductivity. The ac conductivity due
to CBH of an electron in the narrow band limit is given by Eq.
(12) [39]:
π3 2
σ(ω) = N εε0 ω R6ω (12)
12
where N is the density of pair sites, ε the dielectric con-
stant, ε0 the free-space dielectric permittivity and Rω is
the hopping length at frequency ω, which can be obtained
by
e2
Rω = (13)
πεε0 [Wm − kT (1/ωτ0 )]
Fig. 8. Variation of dielectric constant ε with the temperature.
e is the electron charge, k the Boltzmann constant, τ 0 the
characteristic relaxation time and Wm is the dc activation energy.
Finally, the frequency exponent s was expressed according independent of temperature below 400 K for the dielectric con-
the equation [40]: stant and below 370 K for the dielectric loss. They show weak
6kT frequency dispersion below these temperatures also, but they
s=1− (14) show a strong temperature dependence and frequency dispersion
Wm − [kT ln(1/ωτ0 )]
above these temperatures. The temperatures, at which the dielec-
That is according to this model the frequency exponent (s) tric constant and loss increased, appeared to increase gradually
appears to be a temperature dependent factor. The fitting using with the increase of frequency. The temperature dependence of
Eq. (12) for the ac conductivity and Eq. (14) for the exponent the dielectric constant and loss for all other glass compositions
factor s are shown as solid lines in Figs. 6 and 7, respectively for are qualitatively similar with slight differences in the tempera-
the selected samples. From Fig. 6, it is appeared that the fitting tures at which the frequency dispersion occurs.
of the ac conductivity (solid curves) for the selected sample The dielectric constant observed for the studied glass system
appears to be reasonable over the entire temperature range. A is relatively higher than those observed for some other con-
similar fitting is obtained for all other glass samples, indicating ventional glasses [42]. It is important to mention that, at low
that CBH model is the most applicable conduction mechanism frequencies and high temperatures, electrode polarization might
for all the studied glasses. The values of the parameters used in have significant contribution to the dielectric constant. However,
the calculations are given in Table 2. the dielectric constant is found to be independent on the thick-
Fig. 7 shows that the fitting of (s) for the selected samples ness of the samples and the electrode area at all frequency and
appeared to be also reasonable over the entire temperature range, temperature ranges. This indicated that the bulk effect and the
and also a similar fits are obtained for all other glass samples, electrode polarization have little contribution to the dielectric
which confirms the applicability of the CBH mechanism for the properties. This shows clearly the influence of the polarizability
investigated glass system.
Figs. 8 and 9 show the dielectric constant (ε ) and the
dielectric loss (ε ), respectively, for the sample of the com-
position [60V2 O5 ·20SrO·20FeO] as a function of tempera-
ture. It appeared that approximately both ε and ε are almost

Table 2
Parameters calculated from fitting using (CBH) model for V2 O5 –SrO–FeO
glasses
Glass sympole τ 0 (sec) N (cm-3 ) ε

S1 1 × 10−13 7 × 1021 8
S2 9 × 10−13 4 × 1021 3.8
S3 1 × 10−10 1 × 1021 1.95
S4 1 × 10−13 9 × 1020 18.5
S5 1 × 10−13 5.5 × 1020 14
S6 2 × 10−13 2 × 1020 3
S7 1 × 10−13 8 × 1019 4.5
S8 1 × 10−13 5 × 1019 8.5
S9 1 × 10−13 1 × 1019 3.6
Fig. 9. Variation of dielectric constant ε with the temperature.
8 F. Abdel-Wahab et al. / Materials Science and Engineering B 134 (2006) 1–8

of strontium ions on the dielectric properties of the studied glass [11] A.I. Ivon, V.R. Kolbunov, I.M. Chernenko, J. Eur. Ceram. Soc. 23 (2003)
system. 2113.
[12] F.A. Cotton, G. Wilkinson, Advanced Inorganic Chemistry, Wiley Eastern
4. Conclusions Private Limited, New Delhi, 1972.
[13] Y. Abe, G. Li, M. Nogami, T. Kasuga, J. Electrochem. Soc. 143 (1996)
144.
The electrical transport properties of V2 O5 ·SrO·FeO uncon- [14] M. Nogami, Y. Abe, Phys. Rev. B 55 (1997) 12108.
ventional oxide glass system was thoroughly investigated in the [15] A.R. West, Solid State Chemistry and its Applications, Weily, New York,
temperature from 308 and 588 K and four different frequencies. 1995.
According to the obtained results it can be concluded that the [16] F.C. Garcia, I.C.S. Carvalho, E. Hering, W. Margulis, B. Lesche, Appl.
Phys. Lett. 72 (1998) 3252.
dc conductivity ranges from 10−10 to 10−3 S cm−1 in the whole [17] S. Sen, A. Ghosh, J. Phys.: Condens. Matter 11 (1999) 1529.
temperature range and it decrease with the increase of SrO con- [18] C.H. Chung, J.D. Mackenzie, L. Murawski, Rev. Chem. Miner. 16 (1979)
tent. The proposed conduction mechanism was the SPH model 308.
above 403 K while below 403 K mixed conduction of SPH and [19] K. Shimakawa, Philos. Mag. B 60 (1989) 377.
VRH models was proposed. Two different activation energies [20] S. Hazra, A. Ghosh, J. Chem. Phys. 103 (1995) 6270.
[21] L. Murawski, C.H. Chung, J.D. Mackenzie, J. Non-Cryst. Solids 32 (1979)
were found above and below 403 K, which confirm the presence 91.
of two different conduction mechanisms. The ac conductivity [22] M.S. Aziz, F. Abdel-Wahab, A.G. Mostafa, E.M. EL-Agwany, Mater.
was found to follow the CBH model in the whole temperature Chem. Phys. 91 (2005) 532–537.
range under consideration, and the exponent factor s was found [23] H. Hirashima, D. Arai, T. Yoshida, J. Am. Ceram. Soc. 68 (1995) 486.
to vary from 0.1 to 0.83. A good and reasonable fitting can be [24] N.F. Mott, E.A. Davis, Electronic Processes in Non-Crystalline Materials,
Clarendon, Oxford, 1979.
obtained between the proposed models and the experimentally [25] M. Amano, H. Sakata, K. Tanaka, T. Hirayama, J. Ceram. Soc. Jpn. 102
measured data. It was concluded also that both σ dc and σ ac were (1994) 424.
found to decrease with the increase of SrO content while both [26] V.K. Dahwan, A. Mansingh, M. Sayer, J. Non-Cryst. Solids 51 (1982) 87.
the dielectric constant ε and loss ε were found to increase with [27] A. Ghosh, B.K. Chaudhuri, J. Non-Cryst. Solids 83 (1986) 151.
temperature and decrease with the increase in the frequency. [28] H. Mori, H. Sakata, J. Ceram. Soc. Jpn. 102 (1994) 852.
[29] H. Hirashima, M. Mitsuhashi, T. Yoshida, J. Ceram. Soc. Jpn. 90 (1982)
411.
References [30] H. Mori, J. Igarashi, H. Sakata, Glastech. Ber. 68 (1995) 327.
[31] H. Mori, J. Igarashi, H. Sakata, J. Ceram. Soc. Jpn. 101 (1993) 1351.
[1] M. Sayer, A. Mansingh, Phys. Rev. B 6 (1972) 4629. [32] H. Mori, H. Matsuno, H. Sakata, J. Non-Cryst. Solids 276 (2000) 78.
[2] G.N. Greaves, J. Non-Cryst. Solids 11 (1973) 427. [33] H. Mori, T. Kitami, H. Sakata, J. Non-Cryst. Solids 168 (1994) 157.
[3] A. Ghosh, Phys. Rev. B 42 (1990) 5665. [34] D. Adler, Amorphous Semiconductor, Butterworths, 1972.
[4] S. Sen, A. Ghosh, J. Mater. Res. 15 (4) (2000) 995. [35] N.E. Mott, Adv. Phys. 16 (1967) 49.
[5] N.F. Mott, J. Non-Cryst. Solids 1 (1968) 1. [36] V.N. Bogomolov, E.K. Kudinev, U.N. Firsov, Sov. Phys. Solid State 9
[6] I.G. Austin, N.F. Mott, Adv. Phys. 18 (1969) 41. (1968) 2502.
[7] E.C. Zampronio, D.N. Greggio, H.P. Oliveira, J. Non-Cryst. Solids 332 [37] R.D. Shannon, C.T. Prewitt, Acta Crystallogr. B 25 (1969) 925.
(2003) 249–254. [38] S.R. Elliott, Philos. Mag. 36 (1977) 1291.
[8] A.K. Arof, B. Kamaluddin, S. Radhakrishna, J. Phys. III France 3 (1993) [39] S.R. Elliott, Adv. Phys. 36 (1987) 135.
1201. [40] A.R. Long, Adv. Phys. 31 (1982) 553.
[9] M.M. Ahmed, C.A. Hogarth, Phys. Stat. Sol. A 101 (1987) K49. [41] J.C. Dyre, Phys. Lett. A 108 (1987) 457;
[10] I.A. Gohar, Y.M. Moustafa, A.A. Megahed, E. Mansour, Phys. Chem. J.C. Dyre, J. Appl. Phys. 64 (1988) 2456.
Glasses 38 (1) (1997) 37. [42] S. Hazra, A. Ghosh, J. Appl. Phys. 84 (1998) 987.

Вам также может понравиться