Вы находитесь на странице: 1из 10

Materials Chemistry and Physics 89 (2005) 428–437

Structure and ionic transport studies of sodium borophosphate


glassy system
P.S. Anantha, K. Hariharan∗
Department of Physics, Solid State Ionics Laboratory, Indian Institute of Technology Madras, Chennai 600036, India

Received 12 July 2004; received in revised form 15 September 2004; accepted 30 September 2004

Abstract

Sodium borophosphate glasses of composition (mol%) 50Na2 O–50[xB2 O3 –(1−x)P2 O5 ], 0 ≤ x ≤ 0.8 have been prepared by melt quenching
method and characterized through XRD, DSC, FTIR and impedance spectroscopy techniques. The glass transition temperature increases with
the substitution of B2 O3 due to the cross-linking of the network and the FTIR study shows the presence of different structural units in the
network. The ionic conductivity study as a function of composition of B2 O3 shows increment in conductivity with two conductivity maxima at
10 and 30 mol% of B2 O3 and conductivity variations with temperature follow an Arrhenius type behaviour. Transport numbers evaluated for
ions and electrons show that Na+ ions are the mobile species in the investigated systems. The frequency dependence of the electric conductivity
follows a simple power law feature. The analysis of various electrical parameters as a function of temperature in different complex planes
shows that the charge transport occurs by the hopping mechanism.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Glasses; Glass transitions; Electrical conductivity

1. Introduction addition of another glass former oxide such as B2 O3 [3–6].


In general, the competitive network formation by two net-
Fast ion conducting glasses have been extensively studied work formers is expected to enhance the cationic diffusion
due to their potential applications in solid-state ionic devices. that shows promising characteristics such as high ionic con-
For practical use, these glasses must have high ionic conduc- ductivity along with good thermal stability [7–8]. In this con-
tivity as well as thermal and electrochemical stability. Sodium text, the present work is directed towards the preparation and
ion conducting borate glasses [1] have been investigated as electrical characterization of sodium borophosphate glasses
they were considered as potential candidates for solid-state of the composition (mol%) 50Na2 O–50[xB2 O3 –(1−x)P2 O5 ]
batteries. Sodium phosphate glasses [2] are reported to have in which P2 O5 of sodium phosphate system is progressively
high thermal expansion coefficient and electrical conductivity replaced by B2 O3 .
with low glass transition temperature as compared to sodium
borate glass. However, their poor chemical durability to wa-
ter limits their practical use. Sodium metaphosphate glass 2. Experimental
is composed of phosphate chains and cross-linking these
chains, either by decreasing the modifier to phosphorus ratio Sodium borophosphate glasses of composition (mol%)
or by adding another glass former oxide significantly changes 50Na2 O–50[xB2 O3 –(1−x)P2 O5 ], 0 ≤ x ≤ 0.8, were prepared
the glass properties. Earlier studies indicate that the chemi- using analytical grade Na2 CO3 , (NH4 )H2 PO4 and B2 O3
cal durability of phosphate glasses can be improved by the reagents as starting materials. The raw materials were
weighed to the desired composition and the mixture was
∗ Corresponding author. Tel.: +91 44 22578667; fax: +91 44 22578651. heated in a platinum crucible first at 600 ◦ C for 1 h to re-
E-mail address: haran@iitm.ac.in (K. Hariharan). move the volatile products and then melted at 900–1050 ◦ C

0254-0584/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2004.09.029
P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437 429

depending upon the composition for about 2 h with frequent Table 1


stirring to ensure the homogeneity. The melt was then rapidly Glass transition temperature, density and molar volume of (mol%)
50Na2 O–50[xB2 O3 –(1 − x)P2 O5 ] glasses
quenched on to a polished copper mould in air at 120 ◦ C to
avoid shattering of the quenched samples due to thermal stress Composition Tg (◦ C) Density (g/cc) Molar volume (cc)
x (±1.0) (±0.003)
and the samples were annealed well below the glass transi-
tion temperature for about 6 h. Weight loss measurements on 0 284 2.505 40.70 ± 0.05
0.2 332 2.516 37.65 ± 0.05
melting show that the loss was found to be less than 2% for 0.34 359 2.539 35.40 ± 0.04
all compositions. The sodium content was checked through 0.5 412 2.550 32.89 ± 0.04
flame photometry and the error was found to be around 2%. 0.6 422 2.548 31.50 ± 0.04
All the samples were analysed by X-ray diffraction with Cu 0.8 444 2.524 28.94 ± 0.03
K␣ radiation. The glass transition temperature, Tg , was de-
termined using Differential Scanning Calorimetry from room initially up to x = 0.5 after which it decreases; whereas, the
temperature to 500 ◦ C with a heating rate of 10 ◦ C per minute. molar volume decreases almost linearly with composition.
Density measurements were carried out by the fluid displace- The increase in density indicates that the macromolecular
ment method in xylene at room temperature using a balance of structure becomes increasingly more rigid with the substi-
10−4 g sensitivity. Infrared absorption spectra were recorded tution of B2 O3 . DSC curves of selected compositions are
on a FT-IR spectrometer in the 1500–400 cm−1 range using shown in Fig. 1. The glass transition temperature, Tg , in-
pellets made with KBr and finely ground glass powder in a creases non-linearly with the substitution of B2 O3 (Table 1).
20:1 ratio by weight. In an oxide network, Tg is known to depend upon the packing
For electrical conductivity measurements, 1–2 mm thick in the network and hence the rapid increase of glass transi-
bulk glass samples were polished and the flat surfaces were tion temperature in the borophosphate glasses can be related
coated with silver paint to form the electrodes. The sam- to the cross-linking of the network thus providing better ther-
ples were then placed under vacuum in a conductivity cell mal stability for the mixed former glass. Thus, the variation
between two silver discs and the temperature of the sample in Tg with composition can be attributed to the tightness of
was monitored using a chromel–alumel thermocouple posi- packing in the network.
tioned very close to the sample with an accuracy of ±1 ◦ C.
The conductivity of the sample was measured at various 3.2. FTIR studies
temperatures by impedance spectroscopy using a HP 4192A
impedance analyser in the frequency range 5 Hz–13 MHz. The FTIR spectra of the glass samples corresponding to
The electronic contribution to the measured total conductiv- the different compositions are shown in Fig. 2. The posi-
ity has been carried out by dc polarization technique using a
cell of the configuration Ag|glass electrolyte|Ag. A constant
voltage was applied to polarize the cell using millivolt source
and an electrometer was employed to measure the resulting
current as a function of time. A galvanic cell of configuration
Na|glass electrolyte|NaCoO2 was constructed to study the
nature of ion transport. The cathode material, NaCoO2 was
prepared by solid-state reaction route from stoichiometric
amount of Na2 CO3 and Co3 O4 [9]. The material was charac-
terized through XRD and the resulting open circuit voltage of
the galvanic cell was measured using a high input impedance
electrometer.

3. Results and discussion

3.1. XRD, DSC and density studies

All glasses of the composition (mol%) 50Na2 O–


50[xB2 O3 –(1−x)P2 O5 ] were transparent and colourless. The
amorphous nature of the prepared glasses was confirmed us-
ing XRD and the glass-forming region (0 ≤ x ≤ 0.8) has been
identified. Also, the hygroscopic nature of the sodium phos-
phate glass has been considerably reduced on the addition of
B2 O3 in the network. The density and molar volume values Fig. 1. DSC spectra of glasses in the system of composition (mol%)
of the glasses are tabulated in Table 1. The density increases 50Na2 O–50[xB2 O3 –(1 − x)P2 O5 ].
430 P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437

524 cm−1 [12]. As phosphate is replaced by glass-forming


B2 O3 , new structural groups will result, in which the coor-
dination number of boron is 4. When P2 O5 is replaced with
B2 O3 (for x = 0.2), some new bands are observed around 700
and 776 cm−1 . The bands in the region of 700 cm−1 can be
assigned to the bond-bending vibration of B–O–B linkages
in the borate network. [13]. The small sharp band appear-
ing at 776 cm−1 is characteristic of the borate glass matrix
and can be attributed to different four-fold coordinated bo-
rate groups such as triborate, ditriborate and diborate groups
[14]. The bands appearing in the region 860–1100 cm−1 are
due to B–O bond stretching of tetrahedral BO4 units which
are overlapped with phosphate bands. These bands are as-
signed to the vibration of tetraborate (triborate and pentab-
orate) groups of BO4 units [14,15]. The bands appearing in
the region 1150–1450 cm−1 are attributed to the B–O bond
stretching of trigonal BO3 units [14,16]. The fact that the band
at 1250 cm−1 is still observed at high alkali oxide, where no
boroxol rings are present, indicates that tetraborate groups
should also contribute to the absorption. This group proba-
bly consists of PO4 and BO4 units to form BPO4 by sharing
the negative charge supplied by the oxygen of the modifier
[17]. The phosphate units, which progressively disappear as
Fig. 2. Infrared spectra of (mol%) 50Na2 O–50[xB2 O3 –(1 − x)P2 O5 ] B2 O3 content increases, bridge the borate groups. At high
glasses. B2 O3 content (x > 0.5), the bonds in the 800–1100 cm−1 re-
gion gets merged into a single band and finally for x = 0.8, a
broad band appears in the region 900–1100 cm−1 . This fea-
tions and assignments of the absorption bands of glasses ture indicates that the glasses with x > 0.5 are borate rich
are summarised in Table 2. Assignments of various peaks glasses. A detailed characterization could not be done be-
in the spectra have been made on the basis of literature cause it is expected that the presence of more than one glass
reports [10–16]. The sodium metaphosphate (x = 0) glass former greatly increases the number of structural units of
is characterized by a strong broad absorption band around a glass. However, the infrared studies of the investigated
1277 cm−1 which is attributed to P O asymmetric stretching systems suggest that the appearance of new (partially re-
vibrational mode. The weak band at 1155 cm−1 has been as- solved) bands are consistent with the presence of new borate
signed to the symmetric stretching modes (PO2 ) of metaphos- or borophosphate units and the phosphate units bridging the
phate chains [11]. The strong band at 1094 cm−1 is attributed borate groups.
to asymmetric stretching of PO3 − groups in PO4 3− units.
The weak band at 1019 cm−1 is assigned to the symmetric
stretching mode of (PO3 ) units. The band at 883 cm−1 is at- 3.3. Conductivity studies
tributed to the asymmetric stretching of P–O–P groups which
confirms the chain structure of metaphosphate glasses. The 3.3.1. Conductivity as a function of concentration
broad band at 769 cm−1 may be attributed to the symmet- To investigate the conduction characteristics, the
ric stretching vibration of P–O–P rings. The bending vibra- impedance and phase angle of the samples of different com-
tion (δ) of P–O bands are characterized by a broad band at positions have been measured as a function of frequency.

Table 2
Assignments of Infrared spectra peaks of (mol%) 50 Na2 O–50[xB2 O3 –(1 − x)P2 O5 ] glasses
Attribution x=0 x = 0.2 x = 0.34 x = 0.5 x = 0.6 x = 0.8 Reference
Bending vibration (δ) of P–O–P bands 524 524 540 545 541 520 12
νs of P–O–P bands 769 776 758 777 769 775 10–12
νa of P–O–P bands 883 904 914 852 835 852 10–12
νs of PO4 groups 1019 1015 1020 10–12
νa of PO4 groups 1094 1092 1097 10–12
νa of PO2 groups 1277 10–12
B–O–B bending vibration bands 720 702 688 702 13
B–O stretching bands of BO4 units 860–1100 860–1100 850–1100 860–1100 880–1100 14, 15
B–O stretching bands of BO3 units 1248–1391 1173–1384 1240–1385 1357–1380 1183–1385 14, 16
P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437 431

four oxygens in a covalently bonded network used as branch-


ing units between metaphosphate chains, with the forma-
tion of more cross-linked networks such as borophosphate
groups. The borophosphate ‘BPO4 ’ unit is made by corner-
sharing PO4 and BO4 tetrahedra [17], which increases the
reticulation of the network. The formation of BPO4 net-
works in the present system makes the glass heterogeneous
and could produce a weak binding area around the struc-
ture of BPO4 favouring the easy migration of Na+ ions [5].
Thus, the first conductivity maxima (x = 0.2) may be due to
the formation of BPO4 groups. At high borate concentra-
tion, the mixed former glass can be considered as borate
like glass and the observed second conductivity maximum
around x = 0.6 has been attributed to the strong interaction
of Na+ ions with BO4 groups. The observed non-linear con-
ductivity behaviour in the present system as a function of
B2 O3 /B2 O3 + P2 O5 ratio depends on the bonding between
Na+ ions and PO4 groups, Na+ ions and some type of BPO4
groups and Na+ ions and BO4 groups. Also an analysis of the
literature [4] indicates that mixtures of trivalent and pentava-
lent glass formers may give glass networks, which enhance
the conductivity. If more than one glass former is present,
the weak electrolyte model [21] predicts that the conduc-
Fig. 3. Variation of conductivity of (mol%) 50Na2 O–50[xB2 O3 –(1 − tivity goes through a maximum when another progressively
x)P2 O5 ] glasses as a function of B2 O3 /B2 O3 + P2 O5 . Inset: complex substitutes an anion of the one former. Similar observations
impedance plots in the system 50Na2 O–50[xB2 O3 –(1 − x)P2 O5 ] at 200
◦ C. have been reported by other workers in several mixed for-
mer systems such as (P2 O5 + V2 O5 ) [7] and (P2 O5 + TeO2 )
[22].
These values were used to obtain the real (Z ) and imagi-
nary (Z ) parts of the complex impedance and were plot- 3.3.2. Conductivity as a function of temperature
ted on a complex plane. The above plots were analysed In general the cationic conductivity is given by
using ‘EQUIVCRT’ non-linear least squares fitting pro-
gram [18] to obtain the bulk resistance. Knowing the bulk σ = neµ (1)
resistance and cell constant, conductivity has been calcu-
lated for various compositions. Fig. 3 shows the composi- where n is the number of mobile ions per unit volume and µ
tional dependence of electrical conductivity at 473 K for the is the mobility. Both are temperature dependent. Hence, the
glassy system. The inset of the figure shows the complex temperature dependence of number of mobile ions per unit
impedance plots from which the conductivity has been eval- volume in a mixed former sodium borophosphate glass can
uated. As P2 O5 is progressively replaced by B2 O3 , conduc- be explained through thermodynamical activity of sodium
tivity in general increases showing two maxima at 10 and oxide as follows. In sodium borophosphate glass, the net-
30 mol% of B2 O3 (B2 O3 /B2 O3 + P2 O5 ratio = 0.2 and 0.6, work modifier Na2 O, in the presence of network former mix-
respectively). Similar conductivity variation has been ob- ture (P2 O5 + B2 O3 ), dissociates according to the dissociation
served in other mixed former glasses [4–7]. The two con- equilibrium,
ductivity maxima have been qualitatively explained through Na2 O ⇔ Na+ + (ONa)− (2)
the thermodynamic activity enhancement of the dissociat-
ing salt [19,20]. As per weak electrolyte model [21], the In this glass, sodium cationic concentration (N) is high
enhancement of conductivity is due to Na2 O activity vari- and it has been evaluated using density values. It is
ations. The observed conductivity variation can be related found to vary between 1.48 × 1022 atoms cm−3 (x = 0) and
to the formation of different structural units of investigated 2.08 × 1022 atoms cm−3 (x = 0.8). The sodium ions can be
system. The structure of pure vitreous B2 O3 consists of a considered to be close enough for a cation to leave its normal
random network of boroxol rings and BO3 triangles con- site to occupy an already occupied neighbouring site to form
nected by B–O–B linkages through bridging oxygen atoms an interstitial pair. This facilitates a resident sodium cation
[13]. When this type of vitreous B2 O3 is added to metaphos- (dissociated as per Eq. (2)) to leave its normal position in
phate glass, B2 O3 is expected to induce the shortening of order to accommodate the incoming cation. Hence, the coop-
the metaphosphate chain length. The borate groups mainly erative motion of this pair of ions in the glass network [21]
in the BO4 tetrahedra form, in which B is coordinated to and thereby contributing to the cationic conductivity as given
432 P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437

by [8]
   
Ne2 l2 ν0 GF Gm
σ= exp − exp − (3)
6kT 2kT kT
where l is the jump distance between two adjacent sites,
ν0 the attempt frequency, GF = Hf − T Sf the free en-
thalpy associated with the formation of an interstitial pair,
Gm = Hm − T Sm the free enthalpy associated with the
migration of cations, Hm is the height of the potential barrier
between two sites and other terms have their usual meaning.
The above equation (Eq. (3)) can be simplified as
 
Ea
σT = σ0 exp − (4)
kT
with
 
Ne2 l2 ν0 1/2 Sf + Sm
σ0 = exp (5)
6k 2
and
Hf
Ea = + Hm (6)
2
In the present work, the conductivity of the glassy system
Fig. 4. Arrhenius plots for the glasses in the system (mol%)
has been found to obey the Arrhenius relation below the glass
50Na2 O–50[xB2 O3 –(1 − x)P2 O5 ].
transition temperature as given by Eq. (4). Fig. 4 shows the
variation of log σT versus 103 /T for these glasses and all the
curves has been fitted linearly with a correlation coefficient of above results justify the interstitial pair migration mecha-
r = 0.993 ± 0.004. The evaluated conduction parameters are nism of alkali ions and the migration processes are purely
summarised in Table 3. The values of the pre-exponential fac- enthalpic.
tor are almost constant with a mean value of log σ 0 = 5.6 ± 0.3
and hence the observed conductivity variations are closely re- 3.4. Transport number
lated to the activation energy Ea .
In our case, assuming a homogenous distribution of The contribution from the electrons to the total measured
sodium cations in the glass, the mean jump distance of two conductivity of the samples was investigated using dc polar-
adjacent alkali ion sites is related to alkali ion concentration ization method. For this method, two of the best conducting
as l = (1/N)1/3 . The ionic jump frequency υ0 , estimated from systems were polarized using two non-reversible [24] sil-
the far-infrared absorption [23] varies between 0.3 × 1013 ver electrodes. An electric potential of 200 mV was applied
and 1.5 × 1013 Hz in different oxide glasses. Taking the mean at 240 ◦ C on a cell arrangement (+) Ag|glass electrolyte|Ag
value of 1013 Hz for the jump frequency and neglecting the (−) and the resulting currents were measured as a function
entropy term, the value of σ 0 has been estimated for each of time. On application of a dc potential across the sample,
composition, which is in reasonable agreement with the ex- an initial current value of 80.4 ␮A, appeared which was pro-
perimental data as summarised in Table 3. The calculated portional to the applied field. It is known that with silver
value is slightly lower than the experimental one and the dif- reversible electrodes, there is a possibility of Ag+ ion migrat-
ferences can be attributed to the entropy contribution, which ing in the glass [25]. However, after three days, the migration
has not been taken into account in the calculation of σ 0 . The of ions due to the applied electric field is balanced by diffu-

Table 3
Conductivity data of (mol%) 50Na2 O–50[xB2 O3 –(1 − x)P2 O5 ] glasses
B2 O3 x Conductivity σ 200 ◦ C Activation energy Experimental σ 0 Calculated σ 0
(mol%) (S cm−1 ) (±0.5%) Ea (eV) (±0.01) (S cm−1 K) (±0.01) (S cm−1 K)
0 2.38 × 10−6 0.79 5.76 4.88
0.2 1.11 × 10−5 0.72 5.48 4.89
0.34 6.17 × 10−6 0.76 5.6 4.90
0.5 9.46 × 10−6 0.73 5.87 4.91
0.6 1.63 × 10−5 0.68 5.27 4.92
0.8 8.54 × 10−6 0.76 5.77 4.94
P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437 433

Fig. 5. Complex impedance plots for (mol%) 50Na2 O–10B2 O3 –40P2 O5


glass at various temperatures.

sion due to the concentration gradient and hence the cell gets Fig. 6. Plots of ac conductivity σ(ω) as a function of frequency for (mol%)
50Na2 O–10B2 O3 –40P2 O5 .
completely polarized with a steady-state current of 0.3 ␮A,
which is nearly three orders lower than the initial value. The
indicating an activated conduction mechanism and the bulk
ratio of final steady-state current to the initial value gives the
relaxation shifts to higher frequencies. The inclined small
electronic transference number and in the present case, it was
vertical spikes at the low frequency region of the impedance
found to be negligible.
plots can be attributed to the electrode polarization [27]. All
In order to study the mobile species responsible for the
the semi-circles are found to be depressed with their cen-
conduction features, a galvanic cell of configuration Na|glass
ters below the real axis and the amount of inclination in the
electrolyte|NaCoO2 was constructed and the open circuit
straight line is related to the width of the distribution of re-
voltage was measured using an electrometer. Sodium is dis-
laxation time. These results indicate the characteristics of
solved at the anode as Na+ ions, which are transported through
parallel combination of bulk resistance, capacitance and a
glassy electrolyte to the cathode where they are inserted into
constant phase element (CPE) of the sample to model the
the cathode material. The electrochemical reaction of the gal-
bulk response [28].
vanic cell as a whole is as follows:
The frequency dependent conductivity (σ(ω)) data ob-
charge tained at various temperatures are shown in Fig. 6. The con-
NaCoO2  Na1−x CoO2 + xNa+ (7)
discharge ductivity is found to be almost frequency independent in
low frequency region. The dc conductivity has been obtained
An open circuit voltage of 2.7 V at 300 K was obtained for
from the extrapolation of the frequency independent plateau
the above fabricated cell. The above value is comparable to
region. In the high frequency region, conductivity disper-
the theoretical voltage 2.8 V reported for the Na|NaCoO2
sion has been observed. According to the jump relaxation
couple [26]. This shows that the charge transport in the
model [29], the high frequency dispersion arises due to the
sodium borophosphate glasses is mainly due to Na+ ions.
high probability for the correlated forward–backward hop-
3.5. AC conductivity studies ping along with the relaxation of the dynamic cage potential.
The high frequency dispersion is predominant at lower tem-
Electrical impedance or admittance measurements are peratures and the dispersion shifts to higher frequency region
used to study ionic conduction in solids because they enable with the increase of temperature. The phenomena of the con-
the assessment of ionic conductivity and its dependence on ductivity dispersion in the high frequency region has been
a variety of materials characteristics. Fig. 5 shows the com- analysed using the Jonscher’s power law [30] feature
plex impedance plot of a typical composition of the glass σ(ω) = σ(0) + Aωn (8)
50Na2 O–10B2 O3 –40P2 O5 at various temperatures. As the
temperature increases, the radius of the semi-circular arc cor- where σ(0) is the frequency independent dc conductivity of
responding to the bulk resistance of the sample decreases the sample, A the weakly temperature dependent quantity
434 P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437

Table 4
Fitting parameters of ac conductivity data to Eq. (8)
Temperature (K) n A (S cm−1 rad−n )
483 0.51 7.79 × 10−10
473 0.52 7.0 × 10−10
463 0.53 5.11 × 10−10
453 0.55 3.9 × 10−10
443 0.56 3.41 × 10−10
433 0.59 2.32 × 10−10
423 0.63 1.26 × 10−10

and n is the power law exponent which lies in the range


0 < n < 1.
In the present work, dc conductivity σ(0) has been ex-
tracted from the frequency independent plateau of log σ(ω)
versus log f plot. The values A and n have been varied simul-
taneously to get the best fits of the measured data to Eq. (8).
The parameters obtained by fitting are tabulated in Table 4.
The fitted curves represented by the solid lines in the figure
show reasonably good agreement between experimental and
calculated values.
Fig. 8. Plot of log ε vs. log f at different temperatures for the glass compo-
3.6. Dielectric analysis sition (mol%) 50Na2 O–10B2 O3 –40P2 O5 .

In order to understand the relevance of dielectric spectra frequencies, due to rapid reversal of field at the interface,
in the present work, the impedance data were converted into the contribution of charge carriers towards the dielectric con-
permittivity using the relationship stant decreases with increasing frequency. Finally, ε gets sat-
urated with a constant value of ε∞ (∼10). In ionic conducting
1
ε∗ = = ε − jε (9) materials, the conduction losses depend on dc conductivity
jωC0 Z∗ according to the relation
where C0 = Aε0 /d, ε0 is the permittivity of free space and ω σ(0)
is the angular frequency of measurement. Typical plots ex- εσ(0) = (10)
ωε0
hibiting the variation of log(ε ) and log(ε ) as a function of
frequency are shown in Figs. 7 and 8 at different tempera- In the present case, the conduction losses predominate
tures. The observed variation in ε at lower frequencies could at lower frequencies and hence the imaginary part of
be attributed to the formation of a space charge region at the permittivity ε shows 1/ω dependence on frequency at all
electrode and electrolyte interface. In log ε plot, at higher temperatures. Also ε is found to increase with increase in
temperature. However, the plot of log ε versus frequency
does not exhibit any peak and hence it is difficult to draw any
information regarding the dielectric loss peak frequency. But
for high conducting materials it is difficult to subtract the
dc conductivity effect from the total dielectric loss data with
sufficient accuracy. Therefore, the data were converted into
modulus formalism. It has an advantage that it suppresses
the information about electrode effects [27] and can be used
to study conductivity relaxation [31].
In the present work, the impedance data were converted
into electrical modulus using the relationship
1
M∗ = = jωC0 Z∗ = M  + jM  (11)
ε∗
Figs. 9 and 10 show the frequency dependence of real and
imaginary parts of the modulus spectra, respectively, at dif-
ferent temperatures. The real part of the electric modulus
M exhibits very small values at lower frequencies indicating
Fig. 7. Plot of log ε vs. log f at different temperatures for the glass compo- that electrode polarization makes negligible contribution to
sition (mol%) 50Na2 O–10B2 O3 –40P2 O5 . M* [27]. The observed dispersion in M at lower frequen-
P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437 435

Fig. 9. Variation of real part of modulus M with frequency for (mol%)


50Na2 O–10B2 O3 –40P2 O5 at various temperatures. Fig. 11. Plots of normalised (M /M max ) vs. log(f/fM ) for (mol%)
50Na2 O–10B2 O3 –40P2 O5 .
cies can be associated with the ease of migration of the con-
ducting ions or in other words the dispersion is mainly due ulus, M , exhibits a low value at lower frequencies which
to conductivity relaxation. As the frequency of the applied might be due to the large value of capacitance associated
field is increased, M reaches a constant value M∞ (=1/ε∞ ) at with the electrodes. The M plot shows a slightly asymmet-
higher frequencies. The imaginary part of the electric mod- ric peak at each temperature. As the temperature increases,
the peak position shifts to higher frequencies and the fre-
quency corresponding to maximum M is known as relax-
ation frequency fM . The peaks are asymmetric on both sides
of the maxima and the peak height decreases with temper-
ature. The frequency range on the lower side of the peak
frequency fM determines the range in which charge carriers
are mobile over long distances. At the frequency range above
fM , the carriers are spatially confined to their potential wells
[32]. Thus, the peak frequency fM is indicative of transition
from long range to short range mobility and is defined by
the condition 2πfM τ M = 1 [33], where τ M is the conductivity
relaxation time. The inset of Fig. 10 shows the temperature
dependence of relaxation frequency following an Arrhenius
relation fM = f0 exp(−EM /kT) where EM is the activation en-
ergy for the electrical relaxation. In the present work, an ac-
tivation energy of 0.73 ± 0.01 eV has been obtained. This
value is close to the value of activation energy from dc con-
ductivity indicating that the ionic migration is responsible
for the observed dc conductivity and modulus spectrum. The
similarity of dc activation energy and conductivity relaxation
activation energy is the evidence of ionic-hopping mecha-
nism for transport. Fig. 11 shows the normalised plots of
(M /M max ) versus log(f/fM ) at different temperatures which
Fig. 10. Variation of imaginary part of modulus M with frequency for are found to overlap on a single master curve indicating
(mol%) 50Na2 O–10B2 O3 –40P2 O5 at various temperatures. Inset: Arrhenius that the distribution of relaxation time is independent of
plot of the peak frequencies fM . temperature.
436 P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437

electrical conductivity shows two maxima with the substitu-


tion of B2 O3 . The observed variation in conductivity has been
explained through available structural reports and interstitial
pair migration mechanism. Transport number measurements
show that Na+ ions are the mobile species in the investigated
systems. The Jonscher’s universal power law is used to anal-
yse the ac conductivity data. The correlation between the
activation energies calculated through the analysis of dc con-
ductivity data and different complex plane data shows that
the transport in the investigated systems is through hopping
mechanism.

Acknowledgements

The financial support received from the Department of


Science and Technology (DST), India, in the form of a spon-
sored project is gratefully acknowledged.

References

[1] C.C. Hunter, M.D. Ingram, Solid State Ionics 14 (1984) 31.
[2] Y.M. Moustafa, K. El-Egili, J. Non-Cryst. Solids 240 (1998) 144.
Fig. 12. Variation of imaginary part of impedance Z with frequency for [3] A. Costantini, A. Buri, F. Branda, Solid State Ionics 67 (1994) 175.
(mol%) 50Na2 O–10B2 O3 –40P2 O5 glass at various temperatures. Inset: Ar- [4] G. Chiodelli, A. Magistris, M. Villa, Solid State Ionics 18–19 (1986)
rhenius plot of the peak frequencies fZ . 356.
[5] T. Tsuchiya, T. Moriya, J. Non-Cryst. Solids 38–39 (1980) 323.
[6] A. Magistris, G. Chiodelli, M. Duclot, Solid State Ionics 9–10 (1983)
611.
In the case of ideal solid electrolyte, the impedance and [7] R. Kaushik, K. Hariharan, Solid State Commun. 63 (1987) 925.
modulus spectra, i.e., the plots of Z and M versus frequency [8] D. Coppo, M.J. Duclot, J.L. Souquet, Solid State Ionics 90 (1996)
exhibit single Debye peaks, whose maximum occur at the 111.
same frequency. Fig. 12 shows the Z spectra as a function [9] Y. Ma, M.M. Deoff, S.J. Visco, L.C. De Jonghe, J. Electrochem.
of frequency at different temperatures. It is seen that Z peak Soc. 140 (1993) 2726.
[10] M. Harish Bhat, F.J. Berry, J.Z. Jiang, K.J. Rao, J. Non-Cryst. Solids
is broader than ideal Debye peak due to the existence of a 291 (2001) 93.
distribution of relaxation time. Also the peak frequency for [11] R.M. Almieda, J.D. Mackenize, J. Non-Cryst. Solids 40 (1980)
M and Z spectra at different temperatures do not occur at 535.
the same frequency indicating non-Debye type of relaxation [12] C. Dayanad, G. Bhikshamaiah, V.J. Tyagaraju, M. Salagram, A.S.R.
behaviour in the systems. The frequencies corresponding to Krishnamurthy, J. Mater. Sci. 31 (1996) 1945.
[13] J. Krogh-Moe, Phys. Chem. Glasses 6 (1965) 46.
peak maximum obeys the Arrhenius relation with an activa- [14] E.I. Kamitsos, M.A. Karakassides, G.D. Chryssikos, J. Phys. Chem.
tion energy of 0.76 ± 0.01 eV, which is shown as an inset in 90 (1986) 4528.
Fig. 12. Similar value of activation energy obtained from dc [15] C. Julien, M. Massot, M. Balkanski, A. Krol, W. Nazarewicz, Mater.
conductivity, modulus spectra and impedance spectra sug- Sci. Eng. B3 (1989) 307.
gests that the ion transport in the investigated glassy systems [16] E.I. Kamitsos, A.P. Patsis, M.A. Karakassides, G.D. Chryssikos, J.
Non-Cryst. Solids 126 (1990) 52.
is due to hopping mechanism. [17] M. Villa, M. Scagliotti, G. Chiodelli, J. Non-Cryst. Solids 94 (1987)
101.
[18] B.A. Boukamp, Equivalent Circuit, University of Twente, The
4. Conclusion Netherlands, Reports no. CT128/88/CT112/89, 1989.
[19] J.L. Souquet, Solid State Ionics 28–30 (1988) 693.
[20] A.C.M. Rodrigus, M.J. Duclot, Solid State Ionics 28–30 (1988)
The influence of partial replacement of phosphate by 729.
borate in 50Na2 O–50[xB2 O3 –(1−x)P2 O5 ] glasses has been [21] D. Ravaine, J.L. Souquet, Phys. Chem. Glasses 18 (1977) 27.
characterized through XRD, DSC, FTIR and ionic conduc- [22] D. Coppo, M.J. Duclot, J.L. Souquet, Solid State Ionics 90 (1996)
tivity studies. The FTIR study shows the existence of dif- 111.
[23] E.I. Kamitsos, M.A. Karakassides, G.D. Chryssikos, J. Phys. Chem.
ferent structural groups such as BPO4 and BO4 , which are
91 (1987) 5807.
favourable for ionic conduction. The observed variation of [24] A.V. Joshi, J.B. Wagner Jr., J. Phys. Chem. Solids 33 (1972) 205.
glass transition temperature with composition has been cor- [25] C. Thévenin-Annequin, M. Levy, T. Pagnier, Solid State Ionics 80
related with the structural modifications of the network. The (175) (1995) 175.
P.S. Anantha, K. Hariharan / Materials Chemistry and Physics 89 (2005) 428–437 437

[26] L.W. Shacklette, T.R. Jow, L. Townsend, J. Electrochem. Soc. 135 [31] J.M. Réau, X.Y. Jun, J. Senegas, Ch. Le Deit, M. Poulain, Solid
(1988) 2669. State Ionics 95 (1997) 191.
[27] L.M. Hodge, M.D. Ingram, A.R. West, J. Electroanal. Chem. 74 [32] J.M. Bobe, J.M. Réau, J. Senegas, M. Poulain, Solid State Ionics 82
(1976) 125. (1995) 39.
[28] D.P. Almond, A.R. West, Solid State Ionics 11 (1983) 57. [33] C.T. Moynihan, L.P. Boesch, N.L. Laberge, Phys. Chem. Glasses 14
[29] K. Funke, Solid State Ionics 94 (1997) 27. (1973) 122.
[30] A.K. Jonscher, Nature 267 (1977) 673.

Вам также может понравиться