You are on page 1of 5

Chemical Physics Letters 424 (2006) 295–299

www.elsevier.com/locate/cplett

Relaxation of silver ions in superionic borate glasses


S. Bhattacharya, A. Ghosh *

Department of Solid State Physics, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700 032, West Bengal, India

Received 11 April 2006; in final form 20 April 2006


Available online 29 April 2006

Abstract

We have investigated the conductivity spectra at different temperatures for the AgI doped silver borate superionic glasses and esti-
mated the concentration of mobile Ag+ ions as a function of temperature and composition. It is observed that the concentration of
mobile Ag+ ions is independent of temperature and only a fraction of the total Ag+ ions in these glasses participate in the diffusion pro-
cess. The increase in conductivity due to the insertion of AgI can be attributed to the increase in hopping rate of Ag+ ions which is sup-
ported by FT-IR results.
Ó 2006 Elsevier B.V. All rights reserved.

1. Introduction dispersed in the host Ag2O–2B2O3 network [6] and the sec-
ond proposes that AgI forms small clusters or microdo-
The understanding of ion diffusion process in superionic mains of an ‘amorphous-like’ AgI with tetrahedral
glasses is of both practical and academic interest [1–4]. coordination to that of crystalline superionic a-AgI [7].
These materials are currently under investigation for tech- Raman spectroscopy result [8] suggests existence of such
nical application as solid electrolytes in electrochemical a-AgI domains, especially at higher AgI concentrations
devices such as batteries, sensors, electrochromic displays, although (EXAFS) [9], 109Ag NMR [10] and X-ray diffrac-
etc. due to their high ionic conductivity, high stability tion [11] support absence of an ordered AgI phase resem-
and large available composition ranges. It is also interest- bling a-AgI. In the microdomain model, the Ag+ cations
ing in the academic level to understand the nature of ion exist in two coordination environments, one related to
diffusion in these materials. the iodide anion and the other with oxygen ions (bridging
A particularly interesting class of fast-ion conductors is or non-bridging oxygens) [6]. In the latter model, the Ag+
the AgI-doped silver borate glasses which can accommo- cations share a common coordination with iodide anions
date a high content of AgI in a disordered phase without and oxygen atoms [5]. Recently X-ray and neutron-diffrac-
any evidence of crystallization [1,2]. The best conducting tion data of silver borate glasses have been used [12] for
glasses may reach conductivity up to 102 S/cm at room structural modelling for fast ion conducting glass systems
temperature [2]. The role of Ag2O seems to be quite differ- using the reverse Monte Carlo method [13] which predicts
ent from other modifying cations, because the local coordi- that most of the Ag+ ions are coordinated to both I ions
nation is characterized by an unusually low coordination and BO4 units in the network [12]. It is concluded that,
number and a short distance, as shown by EXAFS mea- while the short-range order of the borate network is unaf-
surements [5]. Much attention has been focused on how fected by AgI doping, by increasing the amount of AgI
AgI is introduced into the glassy borate network. Mainly content improves the medium-range order of the glass by
two models have been proposed: first suggests that AgI is inducing ordering between the neighboring boron–oxygen
coordinated with the BO4 units and thus is homogenously chain segments. It is therefore proposed that the improve-
ment of medium-range order in glass can be connected with
*
Corresponding author. Fax: +91 33 2473 2805. the network expansion and the creation of new pathways
E-mail address: sspag@iacs.res.in (A. Ghosh). for ion transport [13].

0009-2614/$ - see front matter Ó 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.cplett.2006.04.077
296 S. Bhattacharya, A. Ghosh / Chemical Physics Letters 424 (2006) 295–299

However, despite many experimental and theoretical (Magna IR-750, Series II) at 25 °C and at relative humidity
efforts [1–3,12,13] the mechanism of Ag+ ion dynamics in of 50–60%. For FT-IR measurements, pellets of thickness
AgI doped borate glasses is partly understood yet due to 1 mm and diameter 13 mm were obtained by pressing a
the difficulty in separating the contribution of the ion con- mixture of 1 part of glass and 60 parts of KBr at a pressure
centration and the mobility from the measured conductiv- of 200 kg/cm2. The electrical measurements such as capac-
ity. There is no general agreement about the fraction of itance and conductance of the samples were carried out on
Ag+ ions participating in the dynamic process, which is gold coated samples of diameter 1 cm in the frequency
one of the most critical parameters to explore the transport range 10 Hz to 2 MHz using a RLC meter (QuadTech,
mechanism of the presently studied system. model 7600)and in the temperature range 93–393 K.
Conductivity spectroscopy is a well established method
for characterizing the hopping dynamics of the ions. A 3. Results and discussion
variety of different materials have been studied using this
technique. It is well documented in the literature [14,15] The frequency dependence of the conductivity at differ-
that in the usual experimental frequency window and tem- ent temperatures for a typical glass composition is shown
perature range, the overall frequency response of the real in Fig. 1a. At low frequencies the conductivity is indepen-
part of the conductivity can be described by dent of the frequency corresponding to the dc conductivity.
n
r0 ðxÞ ¼ rdc ½1 þ ðx=xc Þ ; 06n<1 ð1Þ The frequency dispersion starts at a higher frequency as the
temperature is increased. The dispersive portion is first
which is the sum of a constant dc conductivity, rdc and a analyzed by subtracting the dc conductivity from the mea-
fractional power law dependence with an exponent n. xc sured conductivity as shown in Fig. 1b. It is noted that in
is a characteristic crossover frequency from dc to dispersive
conductivity. Both the dc conductivity and the crossover
frequency, above which r 0 (x) µ xn, are thermally acti- -5
1x10
vated. However, there is another contribution to the dis-
persive conductivity, which consists of a nearly frequency
independent dielectric loss and corresponds to an almost -6
10
linear frequency dependent term of the form r 0 (x) = Ax
for the real part of the complex conductivity. At sufficiently
σ' (ω) (Ω-1 cm-1)

low temperatures or high frequencies, the Ax term domi- 10


-7
T=153K
nates over the power law dependence of the conductivity T=158K
T=163K
given by Eq. (1) [16]. T=173K
T=183K
In this Letter, we have investigated the conductivity 10
-8
T=193K
T=203K
spectra for AgI doped silver diborate glasses at different T=213K
temperatures and estimated Ag+ ions participating in the T=223K

-9 (a)
transport process. We have shown that the increase in the 10
conductivity due to the AgI doping is not due to the 3
10 10
5 7
10
increase in the number of mobile Ag+ ions, but is due to ω [rad s-1]
the increase in the mobility of Ag+ ions. The results are
clarified also by the analysis of the FT-IR absorption spec-
tra of the glasses.
-6
10
(σ' (ω)-σdc) (Ω-1 cm-1)

2. Experimental

Glass samples of compositions xAgI  (1  x)(Ag2O– -7


10 T=153K
2B2O3), where x = 0–0.5 were prepared in two steps. First, T=158K
T=163K
Ag2O–2B2O3 base glass was obtained by melting AgNO3 T=173K
T=183K
and H3BO3 at 750 °C for 6 h in a platinum crucible and -8 T=193K
10 T=203K
then quenching the melt in an aluminum mould. In the sec- T=213K
T=223K
ond step, appropriate amounts of AgI and the base glass
(b)
were mixed, melted in platinum crucible at 750 °C for -9
10
30 min and poured in an aluminum mould to get final glass 4
10
5
10 10
6
10
7

samples. Glass samples of thickness 0.1 cm were obtained ω [rad s ]


-1

for x = 0–0.5. Glass formation was confirmed from X-ray


diffraction. The density of the samples was measured using Fig. 1. (a) Conductivity spectra for the 0.2AgI  0.8(Ag2O–2B2O3) glass
composition at different temperatures shown in the inset. The solid lines
Archimedes’ principle with acetone as an immersion liquid. are best fits to Eq. (1). (b) Variation of the dispersive conductivity
The FT-IR spectra of the bulk glass samples in absorption (r 0 (x)  rdc) with frequency for the same glass composition and temper-
mode were recorded in a Nicolet FT-IR spectrophotometer atures as in (a).
S. Bhattacharya, A. Ghosh / Chemical Physics Letters 424 (2006) 295–299 297

the frequency window the slopes give power law exponent x = 0.0
(0.65) which is much less than unity and independent of 7 x = 0.1
temperature. Thus, the nearly constant loss region is absent x = 0.2
x = 0.3
in the investigated frequency window and the temperature 6 x = 0.4

log10 [ωh (rads-1)]


range, and the data can be described by the power law x = 0.5
model (Eq. (1)). The conductivity spectra for all composi- 5
tions were then fitted to Eq. (1). We have replaced rdc in
Eq. (1) by the following expression for rdc given by the 4
Nernst–Einstein relation
rdc ¼ q2 d 2 nc xh =12pkT ; ð2Þ 3

where nc and d is the mobile ion concentration and the 2


jump distance, respectively. It has been pointed out [17] (a)
that short range displacements of the ions at higher fre- 3 4 5 6 7 8 9
quencies are thought to be coupled to the ionic environ- -1
1000 / T (K )
ment. Interionic interactions become important, and the
short range hopping is often viewed as the highly correlated
motion in which the ions perform several reiterated for- 10
ward–backward hops before completing any successful for-
ward displacement [17,18]. The power law dispersion (Eq. 9
(1)) is thus a consequence of this reiterated hopping and oc-

log10 [ωh (rads-1)]


curs down to the crossover frequency xc below which suc- 8
cessful hops can be completed [17]. Thus in the present
approach it can be tacitly assumed [17–19] that the cross-
over frequency, xc in Eq. (1) is close to the hopping fre- 7
quency of the mobile Ag+ ions, presented as xh in Eq.
(2). Recent NMR and conductivity experiments of some 6
crystals and AgI-doped fast ion conducting glasses [20–
(b)
22] have indicated that xc is slightly longer than xh, which 5
does not change the values of nc drastically, justifying our 0.0 0.1 0.2 0.3 0.4 0.5
assumption for the present AgI-doped glasses. Three x
parameters nc, xh and n were obtained at different temper-
Fig. 2. (a) Arrhenius plots of the hopping frequency of xAgI  (1  x)-
atures for all glass compositions. We have taken the jump (Ag2O–2B2O3) glasses, obtained from the best fits of conductivity spectra,
distance as the nearest Ag–Ag distance for these glasses ob- for different values of x, to Eq. (1). The solid lines are the least-square
tained from structural studies [12]. straight line fits of the data. (b) Compositional variation of the hopping
We note that the values of power law exponent n are frequency of mobile ions xh at room temperature (298 K).
almost independent to AgI content within experimental
errors. The temperature dependency of xh is shown in that a fraction of Ag+ ions participate in the dynamic pro-
Fig. 2a. Fig. 2b shows the variation of the hopping rate cess. In Fig. 4b, we have also included the mobile ion con-
xh of mobile Ag+ ions at room temperature as function centration obtained from NMR experiments [23]. We note
AgI content. It is observed that the hopping rate increases that the fraction of mobile ions obtained from the present
linearly with the increase of AgI content in the host glass analysis are comparable with those reported from NMR
network. From Fig. 2a it is noteworthy that xh shows experiments [23].
Arrhenius behavior with activation energy (Eh) very close The FT-IR absorbance spectra of AgI doped and
to the activation energy (Er) for the dc conductivity rdc undoped silver diborate glasses is shown in Fig. 5a in the
(Fig. 3). The Arrhenius temperature dependence of nc is frequency range 500–1550 cm1. Strong band envelopes
shown in Fig. 4a, which indicates that nc is almost indepen- centered at 1350 and 1000 cm1 are observed in all glass
dent of temperature. The variation of the estimated mobile compositions. These features are attributed to the asymmet-
ion concentration nc with AgI content in the compositions ric stretching of B–O bonds in borate triangles BO3 and
is shown in Fig. 4b. The variation of the total Ag+ ion con- BO2O (1350 cm1) and in tetrahedral units, BO 4
centration is also shown in the same figure. It is clear in the (1000 cm1), respectively [24,25]. The remaining band enve-
figure that the mobile Ag+ ion concentration obtained lopes centered at 700 cm1 arise from the deformation
from the analysis is much less than the total Ag+ ion con- modes of borate network structures [7]. All the spectra in
centration. Thus, a fraction of the Ag+ ions are mobile and Fig. 5a present quite similar band shapes, indicating that
are weakly dependent on composition. AgI addition does not cause the formation of new structural
It is worthy to mention here that the NMR experiments units in the borate network, although the AgI content are
of AgI doped borate glasses [23] have also demonstrated varied in a rather broad range, 0 6 x 6 0.5. However, it is
298 S. Bhattacharya, A. Ghosh / Chemical Physics Letters 424 (2006) 295–299

x=0.0
0 x=0.1
x=0.2
log10 [σdc T (Ω-1 cm-1 K)]

x=0.3 x=0.5
x=0.4
-2 x=0.5

arb. units
x=0.4

x=0.3
-4
x=0.2

x=0.1

-6 x=0.0
(a)

600 900 1200 1500


-8 Wavenumbers (cm ) -1
2 3 4 5 6 7 8 9
-1
1000 / T (K )
Fig. 3. Arrhenius plots of dc conductivity obtained from complex
impedance plots of xAgI  (1  x)(Ag2O–2B2O3) glasses, for different 1.2
values of x. The solid lines are the least-square straight line fits of the
data.

A4/A3
0.9

(b)
0.6
10
22 0.0 0.1 0.2 0.3 0.4 0.5
x
nc (cm -3)

x = 0.0
Fig. 5. (a) FT-IR absorption spectra for xAgI  (1  x)(Ag2O–2B2O3)
21 x = 0.1 glasses for different values of x. (b) Relative absorption A4/A3 as a
10 x = 0.2 function of AgI content in xAgI  (1  x)(Ag2O–2B2O3).
x = 0.3
x = 0.4
x = 0.5
(a)
observed that AgI addition induces a change in the relative
20
10
3 4 5 6 7 8 intensities of the envelopes at 1350 and 1000 cm1. The nar-
-1
1000/T ( K ) rowing of the 1350 cm1 envelope indicates the occurrence
of AgI-induced structural rearrangements in these glasses.
The integrated absorption of the characteristic band
envelopes between 780 and 1180 cm1 (A4) and between
1180 and 1570 cm1 (A3) was calculated to probe the rela-
10
22
tive population of borate tetrahedra (BO 4 ) and triangles
[Ag ] (cm- 3)

(BO3 and B2O), respectively, in the glasses. The A4/A3


ratio presented in Fig. 5b demonstrates a strong depen-
+

dence on the AgI doping into the host network. It is note-


worthy from Fig. 2b that hopping rate xh also varies in the
same manner with AgI content. Since addition of AgI to
(b) Ag2O–2B2O3 glass is not expected to change the total for-
21
10 mal negative charge on the borate network, the dependence
0.0 0.2 0.4 0.6 0.8 1.0
of the A4/A3 ratio on AgI content can be understood in
x term of changes in the population of borate species. This
Fig. 4. (a) Arrhenius plots of the mobile Ag+ ion concentration obtained can be expressed by the following chemical equilibrium
from the analysis of the conductivity spectra of xAgI  (1  x)(Ag2O– [26] that takes place in the melt and involves the isomeric
2B2O3) glasses, for different values of x. (b) Variation of Ag+ ion B2O and BO 4 species
concentration with composition for xAgI  (1  x)(Ag2O–2B2O3) glasses:
(h) total Ag+ ion concentration obtained from the composition and BO
4  B2 O

ð3Þ
density; (e) mobile Ag+ ion concentration nc obtained from the analysis
of the conductivity spectra; ($) mobile Ag+ ion concentration obtained The data presented in Fig. 5b indicate that when the AgI
from NMR experiments [23]. content is increased the above equilibrium shifts progres-
S. Bhattacharya, A. Ghosh / Chemical Physics Letters 424 (2006) 295–299 299

sively to the left. Thus the Ag+ ions coordinate with both [3] J. Kincs, S.W. Martin, Phys. Rev. Lett. 76 (1996) 70.
I ion and BO 4 units with the insertion of AgI into the
[4] G. Carini, M. Cutroni, A. Fontana, G. Mariotto, F. Rocca, Phys.
Rev. B 29 (1984) 3567.
host glass network. The host glass network expands consid- [5] F. Rocca, G. Dalba, P. Fornasini, F. Monti, A.C. Wright, S.A. Feller,
erably to accumulate the dopant ions [12,13] and changes A.C. Hannon, (Eds.) Borate Glasses, Crystals and Melts, Soc. Glass
of the intermediate structure are indicated. This expansion Technol., Sheffield, 1997, p. 295.
enhances the hopping rate of mobile Ag+ ions as shown in [6] C. Chiodelli, A. Magistris, M. Villa, J.L. Bjorkstam, J. Non-Cryst.
Fig. 2b and hence increases the ionic conductivity. Solids 51 (1982) 143.
[7] J.J. Hudgens, S.W. Martin, Phys. Rev. B 53 (1996) 5348.
[8] A. Fontana, F. Rocca, M.P. Fontana, Phys. Rev. Lett. 58 (1987) 503.
4. Conclusions [9] G. Dalba, P. Fornasini, F. Rocca, J. Non-Cryst. Solids 123 (1990) 310.
[10] S.H. Chung, K.R. Jeffery, J.R. Stevens, L. Borjesson, Phys. Rev. B 41
We have studied the dynamics of Ag+ ions in superionic (1990) 6154.
silver borate glasses by conductivity spectroscopy and esti- [11] A. Musinu, G. Paschina, G. Piccaluga, M. Villa, J. Chem. Phys. 86
(1987) 5141.
mated the mobile Ag+ ions participating in the transport [12] J. Swenson, L. Borjesson, R.L. McGreevy, W.S. Howells, Phys. Rev.
process. We have observed that the mobile Ag+ ion con- B 55 (1997) 11236.
centration is less than the total Ag+ ions estimated from [13] D.A. Keen, R.L. McGreevy, Nature 344 (1990) 423.
glass compositions and density and is independent of AgI [14] A.K. Jonscher, Nature (London) 267 (1977) 673.
doping. Thus AgI insertion into the host network primarily [15] D.P. Almond, A.R. West, Nature (London) 306 (1983) 456.
[16] W.K. Lee, J.F. Liu, A.S. Nowick, Phys. Rev. Lett. 67 (1991) 1559.
enhances the mobility, i.e. the hopping rate of mobile Ag+ [17] D.L. Sidebottom, P.F. Green, R.K. Brow, Phys. Rev. Lett. 74 (1995)
ions by the improvement of the medium-range order of the 5068.
glass as observed from the analysis of FT-IR absorption [18] A. Bunde, P. Maass, J. Non-Cryst. Solids 131–133 (1991) 1022.
spectra. [19] A. Rivera, J. Santamaria, C. Leon, T. Blochowicz, C. Gainaru, E.A.
Roseler, Appl. Phys. Lett. 82 (2003) 2425.
[20] C. Leon, J. Santamaria, M.A. Paris, J. Sanz, J. Ibarra, L.M. Torres,
Acknowledgements Phys. Rev. B 56 (1997) 5302.
[21] K.L. Ngai, C. Leon, J. Non-Cryst. Solids 315 (2003) 124.
The work is supported partly by the Department of Sci- [22] N. Kuwatta, T. Saito, M. Tatsumisago, T. Minami, J. Kawamura, J.
ence and Technology (via Grant No. SP/S2/M43/99) and Non-Cryst. Solids 324 (2003) 79.
partly by Department of Atomic Energy (via Grant No. [23] P. Mustarelli, M.P. Infante, A. Magistris, C. Tomasi, L. Linati, Phys.
Chem. Glasses 44 (2003) 159.
2002/37/32/BRNS), Government of India. [24] E.I. Kamitsos, J.A. Kapoutsis, G.D. Chryssikos, J.M. Hutchinson,
A.J. Pappin, M.D. Ingram, J.A. Duffy, Phys. Chem. Glasses 36 (1995)
References 141.
[25] C.P. Varsamis, E.I. Kamitsos, G.D. Chryssikos, Solid State Ionics
[1] C.A. Angell, Annu. Rev. Phys. Chem. 43 (1992) 693. 136–137 (2000) 1031.
[2] T. Minami, J. Non-Cryst. Solids 56 (1983) 15. [26] E.I. Kamitsos, G.D. Chryssikos, J. Mol. Struct. 247 (1991) 1.

You might also like