Sunteți pe pagina 1din 7

Journal of Non-Crystalline Solids 355 (2009) 348354

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


journal homepage: www.elsevier.com/locate/jnoncrysol

FTIR and ultrasonic investigations on modied bismuth borate glasses


H. Doweidar a, Yasser B. Saddeek b,*
a
b

Glass Research Group, Physics Department, Faculty of Science, Mansoura University, Mansoura 35516, Egypt
Physics Department, Faculty of Science, Al-Azhar University, Assiut 71524, Egypt

a r t i c l e

i n f o

Article history:
Received 6 June 2008
Received in revised form 11 December 2008
Available online 21 January 2009
PACS:
43.35.Ae
61.43.Fs
62.20.Dc
62.80.f
78.30.Ly
S10.15

a b s t r a c t
Studies on xRO  30Bi2O3  (70x)B2O3 glasses have been carried out (0 6 x 6 30 mol%, R = Zn, Ba). Elastic
properties and Debye temperature have been investigated using sound velocity measurements at 4 MHz.
The ultrasonic parameters along with the IR spectroscopic studies have been employed to explore the
role of divalent cations in the structure of the studied glasses. Analysis of infrared spectra indicates that
RO is preferentially incorporated into the borate network, forming BO4 units. It is assumed that Bi2O3
enters the structure in the form of BiO6 only. The change of density and molar volume with RO content
reveals that BO4 units linked to R2+ cations are denser than those linked to positive sites in the Bi2O3 network. Predicted values of four co-ordinated boron put forward questions about the reliability of assignment of structural units that Bi2O3 may form.
2008 Elsevier B.V. All rights reserved.

Keywords:
Acoustic properties
FTIR measurements
Infrared properties
Oxide glasses
Borates

1. Introduction
Boron atoms in borate crystals and glasses usually co-ordinate
with either three or four oxygen atoms forming BO3 or BO4 structural units. These two fundamental units can arbitrarily be combined to form either super-structural units or different BxOy
groups like boroxol rings, pentaborate, tetraborate and diborate
groups, etc. The fraction of these structural units depends on the
nature and concentration of the added modiers [15].
In terms of the crystalline conguration, Bi2O3 is an oxide having a high valence cation of low eld strength and high polarizability. So, its glassy phase cannot conventionally be compared with
pure B2O3 glass. Bi2O3 like ZnO can occupy both network-forming
and network modifying positions in the borate network glasses
and, as a result, the physical properties of such glasses exhibit discontinuous changes when the structural role of the cation changes
[612]. Recent IR, UVVis and thermal studies on bismuthatebased glasses revealed the presence of distorted BiO6 polyhedra
as the main structural units in such glasses [1315].

* Corresponding author. Tel.: +20 103620536.


E-mail address: ysaddeek@gmail.com (Y.B. Saddeek).
0022-3093/$ - see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.jnoncrysol.2008.12.008

On these bases, glasses containing Bi2O3 and ZnO have attracted


a considerable attention due to their wide applications in the eld
of glassceramics, thermal and mechanical sensors, reecting windows, or may be used as layers for optical and opto-electronic devices, etc. These glasses have a long infrared (IR) cut-off, and high
third-order non-linear optical susceptibility which makes them
ideal candidates for applications as infrared transmission components, ultra fast optical switches, and photonic devices [1625].
The aim of the present work is to investigate the inuence of
divalent oxides such as ZnO, and BaO on the structure and properties of the thermally stable 0.3Bi2O30.7B2O3 [4] glasses by means
of IR, density and ultrasonic techniques.
2. Experimental procedures
Glass samples of the formula xRO  30Bi2O3  (70x)B2O3,
0 6 x 6 30 mol%, (R = Zn or Ba), have been prepared by the melt
quenching technique. Required quantities of Analar grade ZnO,
BaO, Bi2O3, and H3BO3 were mixed together by grinding the mixture repeatedly to obtain a ne powder. The mixtures were melted
in porcelain crucibles in an electrically heated furnace under ordinary atmospheric conditions at a temperature of about 1273 K for

H. Doweidar, Y.B. Saddeek / Journal of Non-Crystalline Solids 355 (2009) 348354

3 h to homogenize the melt. The glass formed by quenching the


melt on a preheated stainless-steel mold was immediately transferred to another furnace where it was annealed at 623 K for 3 h.
The obtained glasses were lapped and two opposite sides were polished to be suitable for the use in the ultrasonic velocity measurements. Non-parallelism of the two opposite side faces was less
than 0.01 mm.
X-ray diffraction patterns were recorded to check the non-crystallinity of the glass samples using a Philips X-ray diffractometer
PW/1710 with Ni-ltered, Cu Ka radiation (k = 1.542 ) powered
at 40 kV and 30 mA. The patterns (not shown) revealed broad
humps characteristic of the amorphous materials and did not reveal discrete or any sharp peaks.
KBr pellet technique was used to obtain the infrared spectra of
the glasses, at room temperature. The spectra were recorded in the
wavenumber range of 4002000 cm1 with a resolution of 4 cm1
by an infrared spectrophotometer type JASCO, FT/IR 430 (Japan).
The infrared spectra were corrected for the dark current noises,
and normalized to eliminate the concentration effect of the powder
sample in the KBr disc. The obtained spectra were deconvoluted to
enable shedding further light on the structural changes of BO3 triangles and BO4 tetrahedra, as they are the basic units in these
glasses.
The density (D) of the glass samples was determined using
Archimedes technique by using toluene as an immersion uid. At
least, three samples of each glass were used to determine the density. A random error in the density values was found as 1%. The
molar volume (Vm), and the mean atomic volume (Va), have been
determined as M/D, and Vm/n, respectively, where M is the molar
weight of the glass and n is the number of atoms in the unit formula.
The ultrasonic velocities, longitudinal (mL) and shear (mT), at
room temperature (300 K) were obtained using the pulse-echo
method. In this method, x-cut and y-cut transducers (KARL DEUTSCH), operated at a fundamental frequency 4 MHz, and a digital
ultrasonic aw detector (KARL DEUTSCH Echograph model 1085)
were used. The uncertainty in the measurement of the ultrasonic
velocity is 10 m/s. The two velocities besides the density were utilized in determining two independent second-order elastic constants (SOECs), namely, the longitudinal (C11) and shear (C44)
moduli. C11 and C44 describe the elastic strain produced by a small
stress in isotropic amorphous solids like glasses. For pure longitudinal waves C11 = D v 2L , and for pure transverse waves C44 = D v 2T .
The elastic bulk modulus (Ke), Youngs modulus (Y), Debye temperature (hD) and the micro-hardness (H) can be determined from C11
and C44 [2]. The uncertainty in the elastic moduli is 0.15 GPa.

349

Fig. 1. Infrared spectra of the investigated xBaO  30Bi2O3  (70x)B2O3 glasses.


Numbers at the plots represent BaO content in mol%.

3. Results
Figs. 1 and 2 show the effect of composition on the infrared
spectra of the studied glasses. In general, there are three principle
bands at around 700, 1000 and 1350 cm1 with a shoulder at
1240 cm1 for all glasses. Shoulders or weak bands appear around
In
the
spectra
of
445,
485,
520
and
565 cm1.
xZnO  30Bi2O3  (70x)B2O3 glasses (Fig. 2), the broad band around
1000 cm1 seems to be convoluted of two bands around 900 and
1050 cm1. The spectra in Figs. 1 and 2 show mostly no change
with composition in the center of the band around 1350 cm1. In
the case of xBaO  30Bi2O3  (70x)B2O3 glasses (Fig. 1), the center
of the bands around 1000 cm1 shifts to lower wavenumbers, as
the BaO content increases. In addition to the above features a small
absorption band appears in the spectra of all glasses around
3465 cm1 (not shown in the gures). Such a band is attributed
to stretching vibration of OH ion [25]. The spectra do not show
any appreciable change in the peak position and the intensity with
the modier oxide content or its type.

Fig. 2. Infrared spectra of the investigated xZnO  30Bi2O3  (70x)B2O3 glasses.


Numbers at the plots represent ZnO content in mol%.

Table 1 summarizes determined and derived data for the studied glasses. The data reveal a linear increase in D for both glass series. It is noticed that the density of BaO-containing glasses is higher
than the corresponding glasses containing ZnO. Similarly, the

350

H. Doweidar, Y.B. Saddeek / Journal of Non-Crystalline Solids 355 (2009) 348354

Table 1
The density D, molar volume Vm, longitudinal vL and transverse vT ultrasonic velocity, Debye temperature hD and the estimated values of Youngs modulus Y, bulk modulus Ke and
the micro-hardness H of the glasses xRO  30Bi2O3  (70x)B2O3 (R = Zn, Ba).
RO (mol%)

D (g cm3)

Vm (cm3 mol1)

vL (m/s)

vT (m/s)

hD (K)

Ke (GPa)

0 ZnO
5 ZnO
10 ZnO
15 ZnO
20 ZnO
25 ZnO
30 ZnO

4.99
5.09
5.17
5.24
5.33
5.41
5.50

37.76
37.17
36.67
36.29
35.82
35.36
34.92

3700
4046
4255
4300
4355
4391
4450

2054
2286
2391
2450
2500
2530
2580

292
323
336
341
345
349
353

39.6
47.0
53.3
54.0
55.7
58.2
60.1

52.9
66.1
73.8
77.9
82.1
86.7
91.3

3.1
4.1
4.5
5.0
5.4
5.7
6.2

5 BaO
10 BaO
15 BaO
20 BaO
25 BaO
30 BaO

5.32
5.55
5.69
5.96
6.16
6.24

36.24
35.48
35.33
34.47
34.03
34.26

3800
3900
3950
4138
4211
4290

2147
2192
2299
2346
2413
2450

325
339
343
350
352
356

39.7
42.1
43.3
52.3
56.6
60.4

67.6
76.1
81.1
90.3
95.9
102.6

5.3
6.4
7.1
7.1
7.4
7.9

ultrasonic velocity vL increases with increasing BaO or ZnO content.


The increase is however non-linear in the case of ZnO-containing
glasses. As a property that depends on the glass density, vL is greater for glasses containing ZnO than for BaO-containing glasses.
4. Discussion
4.1. Infrared spectral studies
It is known that pure B2O3 glass consists mainly of boroxol ring
B3O9/2 with a three co-ordinated BO3 unit. Addition of Bi2O3 will
change the borate structure by creating BO4 units at the expense
of BO3 structural units. Stone et al. [4] and Cheng et al. [19] studied
the binary glass system Bi2O3B2O3. The infrared spectra show that
by increasing the Bi2O3 content a broad absorption band arises in
the region 8501100 cm1. Bands in this region are not observed
in the infrared spectra of vitreous B2O3. Stone et al. [4] and Cheng
et al. [19] attributed infrared absorption between 900 and
950 cm1 to stretching vibrations of BO4 units that are connected
by the bismuth cations. The crystalline phase of the
70Bi2O3  30B2O3 (mol%) glass consists of 3Bi2O3  5B2O3. This phase
contains super-structural units in the form of isolated dipentaborate groups (B5O7
11 ) [4,19]. The dipentaborate groups are formed
due to acquisition of BO4 tetrahedron resulting in a highly crosslinked network.
Absorption of infrared radiation below 610 cm1 is attributed
either to cationic vibrations in the network [26] or to various
modes of BiO vibration in BiO6 [2729]. Absorption bands around
840 cm1 are related to stretching vibrations of BiO in BiO3 units
[6,28,30]. The latter could be observed only at sufciently high
concentration of Bi2O3 (x 6 0.07) in 99.5% [xB2O3(1x)Bi2O3]
0.5%CuO glasses [31]. BiO3 units could not be detected in
xB2O3(100x)Bi2O3 glasses, where 40 6 x 6 70 mol% [31].
Vibrations of borate units appear as absorption peaks in the
infrared spectra between 700 and 1550 cm1 [26,32,33]. Assignment of the vibration modes is summarized in Table 2. Due to
the absence of absorption bands at 840 cm1 in the infrared spectra (Figs. 1 and 2), therefore, only BiO6 units will be expected to
inuence the borate network of the studied glasses.

Y (GPa)

H (GPa)

The observed broad peak is the result of overlapping of individual bands with each other. Each individual band has its characteristic parameters such as its center (C), which is related to some
type of vibrations of a specic structural group, and its relative area
(A), which is proportional to the concentration of this structural
group. A deconvolution process, as described elsewhere [34],
should be performed to get such parameters. Because we are seeking to follow the change in borate matrix due to additions of RO,
the deconvolution process would be conned in the region above

Table 3
Deconvolution parameters of the infrared spectra of xRO  30Bi2O3  (70x)B2O3
glasses (R = Zn, Ba). C is the component band center (cm1) and A is the relative
area (%) of the component band.
0 ZnO
5 ZnO
10 ZnO
15 ZnO
20 ZnO
25 ZnO
30 ZnO
5 BaO
10 BaO
15 BaO
20 BaO
25 BaO
30 BaO

C
A
C
A
C
A
C
A
C
A
C
A
C
A

685
1.95
681
2.1
684
1.6
688
2.2
691
1.7
692
1.9
697
2.1

883
9.4
892
18.8
904
20.9
886
14.7
861
3.2
886
16.5
875
13.8

1057
25.6
1071
22.9
1066
17.9
1050
22.8
986
31.9
1054
25.8
1039
31.2

1087
5.3

1225
3.5
1220
3.9
1231
2.7
1228
3.9
1220
7.6
1226
10.7
1207
5.9

1292
12.3
1283
10.3

1292
23.4

1426
46
1408
40.9
1356
52.9
1352
50.1
1348
40
1350
34.5
1399
17.8

1646
1.2
1498
1.1
1486
3.8
1484
6
1479
9.9
1468
10
1473
5.7

C
A
C
A
C
A
C
A
C
A
C
A

687
1.5
688
1.5
692
2.1
696
2.3
699
1.8
696
2.6

925
29.9
898
22.3
906
25.9
907
29.9
896
27.9
881
24

1034
10
1024
13.5
1023
20.9

1093
13.6
1075
22.6
1066
18.3
1094
3.3
1089
3.4
1094
1.9

1221
4.1
1228
5.2
1223
7
1220
16.1
1219
16.9
1201
7.4

1305
24.6
1326
27.5
1342
36.6
1357
30.6
1350
29.7
1302
27.8

1454
25.9
1468
20.8
1476
10.0
1475
7.8
1457
6.8
1405
11.5

1506
0.6

1467
3.9

Table 2
Assignment of absorption bands in the infrared spectra of the glasses xRO  30Bi2O3  (70x)B2O3 (R = Zn, Ba).
Wavenumber (cm1)

Assignment

400600
420
681699
861925, 9861094
12001500

BiOBi + BiO in BiO6 octahedral [2729]


ZnO tetrahedral bending vibrations of ZnO4 units [3739]
Bending vibration of BOB in BO3 triangles [26,32,33]
Stretching vibration of BO4 units in various structural groups [26,32,33]
BO stretching vibrations of trigonal BO3 units only [26,32,33]

H. Doweidar, Y.B. Saddeek / Journal of Non-Crystalline Solids 355 (2009) 348354

700 cm1. The deconvolution parameters of the bands for the


investigated glasses are given in Table 3.
Fig. 3 shows the deconvoluted spectrum of the sample having
15 mol% ZnO as an example. The deconvoluted peaks of the base
glass spectrum agree well with that reported by Cheng et al. [19]
and did not indicate the presence of the band at 840 cm1 which
is characteristic of BiO3 units [8]. Table 2 summarizes the major
observed absorption bands in the investigated glasses and their
vibrational types, respectively.
The parameters C and A of component bands (Table 3) can be
used to calculate the fraction N4 of BO4 units in the borate matrix.
N4 is dened as the ratio of [concentration BO4 units/concentration
of (BO3 + BO4) units]. The relative area of a component band corresponds to the relative concentration of the structural unit originating it [34]. Then, by assigning the component bands to their
structural units (Table 2) the N4 value can be calculated for each
composition. A tolerance in the N4 values is estimated as 5%.
Fig. 4 shows a common increase of N4 with increasing RO
content in xBaO  30Bi2O3  (70x)B2O3 and xZnO  30Bi2O3 
(70x)B2O3 glasses. The trend can be approximated to a linear increase in N4. The rate of increase is nearly the same for both types
of glass. A value of about 0.36 for N4 of the base glass, 30Bi2O3
(70x)B2O3, reveals that the B2O3 matrix is progressively modied
by Bi2O3. NMR investigations on Bi2O3B2O3 glasses [35,36] support the above conclusion. It is known that each B2O3 molecule

351

needs one oxygen atom to be converted into two BO4 tetrahedra.


This can be veried by sharing bridging oxygen atoms between
BO4 units and BiO6 units. Neutralization of the excess negative
charge on BO4 units implies the presence of positively charged
structural defects in the Bi2O3 network. BO4 units would also form
when introducing RO into Bi2O3B2O3 matrix. Each RO molecule
can convert two BO3 units into two BO4 units. It can be said that,
in the studied glasses, there is a competition between BiO6 units
and RO to form BO4 units. The quantity B4 (in mol%) of B2O3 in
the form of BO4 units can be calculated as

B4 N4 B2 O3

Fig. 5 shows the dependence of B4 concentration on the RO content in the glasses investigated. As shown it rstly increases, indicating that the rst addition of RO contributes directly to the
formation of BO4 units. Afterwards B4 decreases steadily, in spite
of the increase in N4 (Fig. 4). The decrease in B4 can be due to
the decrease in the B2O3 when increasing the concentration of
RO. Fig. 5 leads to an important conclusion. The value of B4 is about
19 mol% for both types of glass at 30 mol% RO. This quantity of
B2O3 needs an equivalent concentration of either Bi2O3 or RO to
be converted from BO3 to BO4 units. Even if it is assumed that only
RO associates with B2O3, then there would be at least about
11 mol% RO to be introduced into the Bi2O3 matrix. The absence
of a peak at 610 cm1 that is related to BiO stretching vibrations
of non-bridging oxygen ions in BiO6 units [27] rules out the above
assumption. In certain cases ZnO can form ZnO4 tetrahedra that behave as network former units in the glass structure. An evidence
for formation of such units is the appearance of an absorption peak
around 420 cm1 in the infrared spectra [3739]. This peak is not
observed in Fig. 2 and formation of ZnO4 units in the studied
xZnO  30Bi2O3  (70x)B2O3 glasses is therefore excluded. Similarly, as far as the authors know, BaO enters the glass structure
only as a modier oxide. In the light of the N4 values of the studied
glasses, it can be said that the assignment of bands related to structural units of Bi2O3 and the structure of such glasses needs further
studies to be claried.
4.2. Density and molar volume

Fig. 3. Deconvolution of the infrared spectrum of the glass 15ZnO  30Bi2O3 


55B2O3.

The density is a powerful tool capable of exploring the changes


in the structure of glasses, and is affected by the structural softening/compactness, change in geometrical conguration, co-ordination number, cross-link density and dimension of interstitial
spaces of the glass. The density of the studied glasses (Fig. 6) in-

Fig. 4. Change of the fraction of four co-ordinated boron N4 with RO content in


xZnO  30Bi2O3  (70x)B2O3 and xBaO  30Bi2O3  (70x)B2O3 glasses. Lines are
tting plots of the data. Error limit for N4 values is estimated as 5%.

Fig. 5. Concentration of B2O3 converted into BO4 units as a function of RO content in


xZnO  30Bi2O3  (70x)B2O3 and xBaO  30Bi2O3  (70x)B2O3 glasses. Lines are
guide for the eyes. Uncertainty in the BO4 data taken as 3%.

352

H. Doweidar, Y.B. Saddeek / Journal of Non-Crystalline Solids 355 (2009) 348354

The linear change in Fig. 6 suggests that each component contributes to the density with a specic factor. In such a case, the
density can be given as

D fB C B fBi C Bi fRO C RO

where fB, fBi and fRO are, respectively, the factors with which B2O3,
Bi2O3 and RO contribute to the density and CB, CBi and CRO the molar
fraction of those oxides, respectively. Starting with fB = 1.823 g/cm3
(the density of vitreous B2O3 [42]) and solving Eq. (2) for different
values of D lead to fBi = 12.47 g/cm3, fBa = 6.25 g/cm3 and
fZnO = 3.45 g/cm3. The lled symbol in Fig. 6 represents calculated
densities obtained from Eq. (2) and the predicted factors. It appears
that the greatest contribution to density is from Bi2O3 units. In addition BO4 units linked to Ba2+ ions are considerably denser than
those linked to Zn2+ ions.
Fig. 6. Density D and molar volume Vm of xZnO  30Bi2O3  (70x)B2O3 and
xBaO  30Bi2O3  (70x)B2O3 glasses in dependence of the RO content. Lines are
tting plots of the data. A random error in the density values was found as 1%,
whereas the error limit in Vm values is estimated as 0.9%.

creases linearly with increasing RO content. The density value of


30Bi2O3  70B2O3 glass agrees well with that reported elsewhere
[40,41]. In spite of the higher electronegativity (1.65) and the smaller radius (1.33 ) of Zn, as compared with the parameters of Ba
(0.89 and 2.17 , respectively) it is noticed that the densities of
BaO-containing glasses are greater than those for corresponding
ZnO-containing glasses. Furthermore, the rate of density increase
is higher for the glasses containing BaO. It seems here that the heavier molecular mass of BaO predominates the effects of the mentioned parameters in determining the density change of the
studied glasses.
At the rst sight, the increase in the density can be related to
the replacement of B2O3 with oxides of greater molecular mass
such as ZnO or BaO. However it must be indicated that is not the
real reason in all cases. For example in Li2OB2O3 [42] or Li2O
SiO2 [43] glasses the density increases markedly with increasing
Li2O content at the expense of B2O3 or SiO2, which have greater
molecular mass. The reason of density increase in those cases is related to the type of structural unites that form when Li2O is incorporated into the glass structure. Depending on the composition of
Li2OB2O3 glasses, Li2O converts symmetric BO3 triangles into BO4
tetrahedra or converts the latter into asymmetric BO3 triangles.
Both the BO4 tetrahedra and asymmetric BO3 triangles are considerably denser than the symmetric BO3 triangles [42]. In Li2OSiO2
glasses, SiO4 tetrahedra having one, two, three or four non-bridging
oxygen ions are denser than those without non-bridging oxygen
ions [43].
The increase of density in Fig. 6 can be explained by considering
formation of BO4 units due to introducing RO in the borate matrix.
A compensation of the negative charge on the BO4 tetrahedra
would be veried from positively charged structural defects in
the Bi2O3 network or from one of the positive charges of R2+ cation
for each BO4 tetrahedron. As deduced from the analysis of infrared
spectra, formation of non-bridging oxygen ions in the Bi2O3 matrix
is not expected due to the absence of absorption bands at
610 cm1. It is then deduced that RO is preferentially incorporated
into the borate network. This means that in xRO  30Bi2O3 
(70x)B2O3 glasses BO4 units linked to R2+ cations would replace
those compensated with positive sites in the Bi2O3 network. The
increase in density in Fig. 6 reveals that the former type of BO4
units is denser than the latter. This trend indicates that the BO4
units linked to R2+ cations have considerably smaller volume than
those compensated with positive sites in the Bi2O3 network.
Formation of BO4 units linked to R2+ cations would then cause a
contraction in the molar volume, as shown in Fig. 6.

4.3. Ultrasonic and elastic behavior of glass


The variation of the values of longitudinal and shear ultrasonic
wave velocities with x are listed in Table 1. The ultrasonic velocity
vL and Debye temperature hD of the studied glasses increase as the
ZnO content or BaO content increases (Figs. 7 and 8, respectively).
vT is not shown as its variation shows a similar trend as that of vL.
Table 1 gives the estimated values of the elastic moduli; bulk modulus Ke, Youngs modulus Y and micro-hardness H. As appears from
Table 1 both Ke and Y as well as H increase with increasing the RO
content. The values of ultrasonic parameters of the base glass agree
with those reported elsewhere [3739].
The observed increase in the ultrasonic velocity can be explained by assuming that, Zn2+ or Ba2+ ions enter interstitially,
and as a result some type of modication of BOB, and BiOB
linkages, which already exist in the glass, into BOZn or BOBa
bonds may occur. Besides, the conversion of BO3 units into BO4
ones results in a decrease in the molar volume which contract
the glass network. Thus, an increase in the rigidity will contribute
to the increase in the ultrasonic velocity. This behavior indicates
that the replacement of B2O3 by ZnO or BaO improves the mechanical properties and the strength of the cross-links between chains
of the borate glasses.
Debye temperature hD plays an important role in solid materials
in the determination of elastic moduli and atomic vibrations. hD
represents the temperature at which all the low frequency lattice
vibrational modes are excited. It is known that Debye temperature
depends directly on the mean ultrasonic wave velocity [2], so

Fig. 7. Dependence of the longitudinal ultrasonic velocity vL, in


xZnO  30Bi2O3  (70x)B2O3 and xBaO  30Bi2O3  (70x)B2O3 glasses, on the RO
content. Lines are tting plots of the data. The uncertainty in the measurement of vL
is 10 m/s.

H. Doweidar, Y.B. Saddeek / Journal of Non-Crystalline Solids 355 (2009) 348354

353

attributed to the increase in the number of bonds per glass formula


unit and to the transformation of low co-ordinated structural units
BO3 into high co-ordinated structural BO4 units, which in turn
change the type of bonding of the investigated glasses. Such an increase in the bulk modulusvolume relationship was observed in
some other types of glasses [46,47]. The non-linear change of log
Ke reveals that the change in Va is related to change in the nature
of bonding and in the co-ordination of the constituent of the polyhedra in the structure.
5. Conclusion

Fig. 8. Composition dependence of the Debye temperature hD of the studied glasses.


The line is a tting plot of all presented data. The error limit in hD is estimated as
1%.

Debye temperature increases as the RO content increases (Table 1).


The increase in hD can be attributed to strengthening of the structure as revealed by the decrease in the molar volume and the increase in the hardness.
The composition dependence of the bulk modulus can be explained by taking into account the relation between the bulk modulus and the mean atomic volume Va. The bulk modulus Ke and the
mean atomic volume Va are related to each other by the relation; Ke
Vm
a C [44] (where C is a constant). The value of m for oxide
glasses is 4 [45], and the variation of m in the bulk modulus
volume relationship is determined by the nature of the bonding,
and the co-ordination polyhedra. When the volume change occurs
without change in the nature of the bonding or change in the coordination polyhedra, the log Kelog Va plot in general will be linear. The bulk modulus of a covalent network is determined by the
bond density (number of bonds in a unit volume), and by the
stretching force constant. The force constant is related to the cation
eld strength of the modier, i.e., high eld strength cations polarize their environment strongly and enhance the iondipole interaction. Strengthen the cation eld strength will increase the packing
density due to the local contraction of the network around such a
cation together with the effects of increasing the bulk modulus.
The log Kelog Va plot for the investigated glasses (Fig. 9) shows
an increase in the bulk modulus with increasing the mean atomic
volume. The observed increase in Ke in spite of the increase in Va is

Fig. 9. Correlation between logKe and logVa for the studied glasses. The line is a
tting plot of all presented data.

Investigations on IR spectra and elastic properties of


xRO  30Bi2O3  (70x)B2O3 glasses (0 6 x 6 30 mol%, R = Zn or Ba)
have been carried out to elucidate the structural role of ZnO and
BaO. Analysis of IR spectra indicates that the Zn2+ and Ba2+ ions
are preferentially incorporated into the borate network as modiers forming BO4 units. The increase in density, and thus the decrease in molar volume, with increasing RO content indicate that
BO4 units linked to R2+ cations are denser than those linked to positive sites in the Bi2O3 network. The decrease in molar volume affects the ultrasonic properties of these glasses. The sound
velocities, Debye temperature, and the elastic properties show an
increasing trend with the increase in RO content. The N4 values obtained from analyzing the infrared spectra of the studied glasses
throw questions about the reliability of the present assignment
of the structural units that Bi2O3 forms in the glass structure.
Acknowledgments
One of the authors (Y.B. Saddeek) wishes to thank Al-Azhar University for the nancial support.
References
[1] J.E. Shelby, Introduction to Glass Science and Technology, The Royal Society of
Chemistry, UK, 1997.
[2] A. Varshneya, Fundamentals of Inorganic Glasses, Academic Press Inc., New
York, 1994.
[3] T. Yano, N. Kunimine, S. Shibata, M. Yamane, J. Non-Cryst. Solids 321 (2003)
137.
[4] C.E. Stone, A.C. Wright, R.N. Sinclair, S. Feller, M. Affatigato, D.L. Hogan, N.D.
Nelson, C. Vira, Y.B. Dimitriev, E.M. Gattef, D. Ehrt, Phys. Chem. Glasses 41 (6)
(2000) 409.
[5] C. Stehle, C. Vira, D. Hogan, S. Feller, M. Affatigato, Phys. Chem. Glasses 39 (2)
(1998) 83.
[6] A. Wells, Structural Inorganic Chemistry, fourth Ed., Clarendon Press, Oxford,
1975.
[7] W. Vogel, Chemistry of Glass, American Ceramic Society, Westerville, OH,
1985.
[8] L. Baia, R. Stefan, J. Popp, S. Simon, W. Kiefer, J. Non-Cryst. Solids 324 (2003)
109.
[9] S. Simon, M. Todea, J. Non-Cryst. Solids 352 (2006) 2947.
[10] V.C. Gowda, C.N. Reddy, K.C. Radha, R.V. Anavekar, J. Etourneau, K.J. Rao, J. NonCryst. Solids 353 (2007) 1150.
[11] K.J. Rao, Structural Chemistry of Glasses, Elsevier, North Holland, 2002.
[12] K. El-Egili, H. Doweidar, Phys. Chem. Glasses 39 (6) (1998) 332.
[13] J. Fu, H. Yatsuda, Phys. Chem. Glasses 36 (1995) 211.
[14] J. Fu, Phys. Chem. Glasses 37 (1996) 84.
[15] A. Pan, A. Ghosh, J. Mater. Res. 17 (2002) 1941.
[16] B.H. Venkataraman, K.B.R. Varma, Opt. Mater. 28 (2006) 1423.
[17] F. Borsa, D.R. Torgeson, S. Martin, H. Patel, Phys. Rev. B 46 (1992) 795.
[18] Y. Dimitriev, V. Mihailova, E. Gattef, Phys. Chem. Glasses 34 (1986) 114.
[19] Y. Cheng, H. Xiao, Wenming Guo, Weiming Guo, Thermochim. Acta 444 (2006)
173.
[20] H. Zheng, J.D. Mackenzie, Phys. Rev. B 38 (1988) 7166.
[21] H. Zheng, R. Xu, J.D. Mackenzie, J. Mater. Res. 4 (1989) 911.
[22] D.W. Hall, M.A. Newhouse, N.F. Borrelli, W.H. Dumbaugh, D. Weidman, Appl.
Phys. Lett. 54 (1989) 1293.
[23] M. Onishi, M. Kyoto, M. Watanabe, Jpn. J. Appl. Phys. 30 (1991) L988.
[24] W.H. Dumbaugh, Phys. Chem. Glasses 19 (1978) 121.
[25] Huaxin Li, Huixing Chen, Wei Lin, Lan Luo, J. Non-Cryst. Solids 352 (2006)
3069.
[26] E.I. Kamitsos, A.P. Patsis, M.A. Karakassides, G.D. Chryssikos, J. Non-Cryst.
Solids 126 (1990) 52.

354

H. Doweidar, Y.B. Saddeek / Journal of Non-Crystalline Solids 355 (2009) 348354

[27] I. Ardelean, S. Cora, R.C. Lucacel, O. Hulpus, Solid State Sci. 7 (2005) 1438.
[28] Y. Dimitriev, M. Mihailova, in: Proceedings of the 16th International Congress
on Glass, vol. 3, Madrid, 1992, p. 293 (cited in [18]).
[29] M.E. Lines, A.E. Miller, K. Nassau, K.B. Lyons, J. Non-Cryst. Solids 89 (1987) 163.
[30] R. Iordanova, V. Dimitrov, Y. Dimitriev, D. Klissurski, J. Non-Cryst. Solids 180
(1994) 58.
[31] L. Baia, R. Stefan, W. Kiefer, J. Popp, S. Simon, J. Non-Cryst. Solids 303 (2002)
379.
[32] E.I. Kamitsos, M.A. Karakassides, G.D. Chryssikos, J. Phys. Chem. 91 (1987)
1073.
[33] E.I. Kamitsos, M.A. Karakassides, G.D. Chryssikos, Phys. Chem. Glasses 28
(1987) 203.
[34] Y. Moustafa, H. Doweidar, G. El-Damrawi, Phys. Chem. Glasses 35 (1994) 104.
[35] A. Bishay, C. Maghrabi, Phys. Chem. Glasses 10 (1969) 1.
[36] P.J. Bray, Annual Summary Report, 1964. Contract No. 562(26), 10 October
196310 October 1964 (Cited in [34]).

[37] L. Del Longo, M. Ferrari, E. Zanghellini, M. Bettinelli, J.A. Capobianco, M.


Montagna, F. Rossi, J. Non-Cryst. Solids 231 (1998) 178.
[38] S.G. Motke, S.P. Yawale, S.S. Yawale, Bull. Mater. Sci. 25 (1) (2002) 75.
[39] S. Bale, N.S. Rao, S. Rahman, Solid State Sci. (2007), doi:10.1016/
j.solidstatesciences.2007.09.017.
[40] S. Yawale, S. Pakade, C. Adgaonkar, Acoustica 81 (1995) 184.
[41] A. El-Adawy, Y. Moustafa, J. Phys. D: Appl. Phys. 32 (1999) 2791.
[42] H. Doweidar, J. Mater. Sci. 25 (1990) 253.
[43] H. Doweidar, Phys. Chem. Glasses 39 (5) (1998) 286.
[44] E. Gopal, T. Mukuntan, J. Philip, S. Sathish, Pramana J. Phys. 28 (5) (1987) 471.
[45] V. Rajendran, N. Palanivelu, B.K. Chaudhuri, K. Goswami, J. Non-Cryst. Solids
320 (2003) 195.
[46] R. El Mallawany, Tellurite Glasses Physical Properties and Data, CRC Press, New
York, 2002.
[47] S. Mahadevan, A. Giridhar, A.K. Singh, J. Non-Cryst. Solids 57 (1983) 423.

S-ar putea să vă placă și