18.03.2014 Views

Optical properties of the Na2O–B2O3–Bi2O3–MoO3 glasses

Optical properties of the Na2O–B2O3–Bi2O3–MoO3 glasses

Optical properties of the Na2O–B2O3–Bi2O3–MoO3 glasses

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Journal <strong>of</strong> Alloys and Compounds 494 (2010) 210–213<br />

Contents lists available at ScienceDirect<br />

Journal <strong>of</strong> Alloys and Compounds<br />

journal homepage: www.elsevier.com/locate/jallcom<br />

<strong>Optical</strong> <strong>properties</strong> <strong>of</strong> <strong>the</strong> Na 2 O–B 2 O 3 –Bi 2 O 3 –MoO 3 <strong>glasses</strong><br />

Yasser B. Saddeek a , K.A. Aly a,∗ , A. Dahshan b , I.M.El. Kashef c<br />

a Physics Department, Faculty <strong>of</strong> Science, Al-Azhar University, P.O. 71452, Assiut, Egypt<br />

b Department <strong>of</strong> Physics, Faculty <strong>of</strong> Science, Suez Canal University, Port Said, Egypt<br />

c Department <strong>of</strong> Physics, Faculty <strong>of</strong> Education, Suez Canal University, Al Arish, Egypt<br />

article<br />

info<br />

abstract<br />

Article history:<br />

Received 29 July 2009<br />

Received in revised form<br />

14 November 2009<br />

Accepted 17 November 2009<br />

Available online 24 November 2009<br />

PACS:<br />

61.43.Fs<br />

78.20.Ci,78.30.−j<br />

Keywords:<br />

Borate <strong>glasses</strong><br />

Bismuthate <strong>glasses</strong><br />

<strong>Optical</strong> <strong>properties</strong><br />

Glasses with compositions (100 − x)Na 2 B 4 O 7 –0.5Bi 2 O 3 –0.5MoO 3 , with 0 ≤ x ≤ 40 mol% have been prepared<br />

using <strong>the</strong> melt quenching technique. The optical transmittance and reflectance spectrum <strong>of</strong> <strong>the</strong><br />

<strong>glasses</strong> have been recorded in <strong>the</strong> wavelength range 300–1100 nm. The values <strong>of</strong> <strong>the</strong> optical band gap<br />

E opt<br />

g<br />

for indirect transition and refractive index have been determined for 0 ≤ x ≤ 40 mol%. The average<br />

electronic polarizability <strong>of</strong> <strong>the</strong> oxide ion ˛O2− and <strong>the</strong> optical basicity have been estimated from <strong>the</strong> calculated<br />

values <strong>of</strong> <strong>the</strong> refractive indices. The variations in <strong>the</strong> different physical parameters such as <strong>the</strong><br />

optical band gap, <strong>the</strong> refractive index, <strong>the</strong> average electronic polarizability <strong>of</strong> <strong>the</strong> oxide ion and <strong>the</strong> optical<br />

basicity with Bi 2 O 3 and MoO 3 content have been analyzed and discussed in terms <strong>of</strong> <strong>the</strong> changes in<br />

<strong>the</strong> glass structure. The results are interpreted in terms <strong>of</strong> <strong>the</strong> increase in <strong>the</strong> number <strong>of</strong> non-bridging<br />

oxygen atoms, substitution <strong>of</strong> longer bond-lengths <strong>of</strong> Bi–O, and Mo–O in place <strong>of</strong> shorter B–O bond and<br />

<strong>the</strong> change in Na + ion concentration.<br />

© 2009 Elsevier B.V. All rights reserved.<br />

1. Introduction<br />

The most common applications <strong>of</strong> <strong>glasses</strong> are based on <strong>the</strong>ir<br />

optical transparency in <strong>the</strong> visible region in a designed range <strong>of</strong><br />

wavelengths. Application <strong>of</strong> <strong>glasses</strong> for making optical instruments<br />

and lenses is based on <strong>properties</strong> like refractive index and its dispersion,<br />

which can be varied by varying chemical composition. One<br />

<strong>of</strong> <strong>the</strong> more important applications based on optical <strong>properties</strong> <strong>of</strong><br />

glass today is in <strong>the</strong> field <strong>of</strong> information technology through <strong>the</strong> use<br />

<strong>of</strong> glass fibers for transatlantic communication cables, telecoms and<br />

cable TV [1].<br />

Bi 2 O 3 exists in a monoclinic form, and has an irregular octahedral<br />

arrangement <strong>of</strong> six oxygen atoms located at distances from<br />

2.14 to 2.8 Å. Three <strong>of</strong> <strong>the</strong>se are appreciably closer to (2.14–2.29 Å)<br />

than <strong>the</strong> o<strong>the</strong>r three (2.48–2.8 Å). Thus, pure Bi 2 O 3 glass due to its<br />

low field strength and its high polarizability cannot be obtained<br />

compared with pure B 2 O 3 glass. However, in <strong>the</strong> presence <strong>of</strong> very<br />

small additions <strong>of</strong> conventional glass formers, it may build a glass<br />

network <strong>of</strong> BiO 3 pyramids. With larger additions (e.g., 9–75 mol%<br />

B 2 O 3 ), widespread areas <strong>of</strong> glass formation have been found [2,3].<br />

Also, <strong>the</strong> bismuth borate glass systems have nowadays-wide applications<br />

for optoelectronic materials such as laser host fibers for<br />

∗ Corresponding author. Tel.: +20 126418273/882325647; fax: +20 882325436.<br />

E-mail addresses: kamalaly2001@gmail.com (K.A. Aly),<br />

adahshan73@gmail.com (A. Dahshan).<br />

communications and photonic switches. Also, <strong>the</strong>se <strong>glasses</strong> exhibit<br />

a high refractive index, widely adjustable in <strong>the</strong> range from 1.8 to<br />

2.5 by <strong>the</strong> appropriate glass composition and have high energy gap<br />

attributable to <strong>the</strong> high polarizability <strong>of</strong> Bi 2 O 3 which allows using<br />

<strong>the</strong>m in wide spectra range from UV up to infrared [4–6].<br />

On <strong>the</strong> o<strong>the</strong>r hand, classical alkali borate <strong>glasses</strong> have been<br />

studied extensively because <strong>of</strong> <strong>the</strong>ir important advantages as solid<br />

electrolytes in storage batteries. In this type <strong>of</strong> glass, Na 2 O converts<br />

firstly, <strong>the</strong> three oxygen-coordinated trigonal borons (BO 3 ) to four<br />

oxygen-coordinated tetrahedral borons (BO 4 ). This procedure continues<br />

until <strong>the</strong> concentrations <strong>of</strong> BO 3 and BO 4 are nearly equal in<br />

<strong>the</strong> glass structure. Boron atoms in <strong>the</strong>se <strong>glasses</strong> are characterized<br />

by its high strength covalent B–O bonds, and with <strong>the</strong> modifierdependence<br />

B 4 /B 3 ratio which is affected in its turn by <strong>the</strong> nature<br />

and <strong>the</strong> concentration <strong>of</strong> <strong>the</strong> modifier oxides [7–16].<br />

In view <strong>of</strong> <strong>the</strong> aforementioned aspects, (100 − x)Na 2 B 4 O 7 –<br />

0.5Bi 2 O 3 –0.5MoO 3 , with 0 ≤ x ≤ 40 mol%, <strong>glasses</strong> have been syn<strong>the</strong>sized<br />

and <strong>the</strong> studied to explore <strong>the</strong> relationship between <strong>the</strong><br />

structure <strong>of</strong> <strong>the</strong> glass and its macroscopic behaviour for <strong>the</strong> design<br />

<strong>of</strong> materials suitable for specific applications. Infrared and UV spectroscopy<br />

have been used for <strong>the</strong> present investigations.<br />

2. Experimental procedures<br />

The glass samples having <strong>the</strong> general chemical formula (100 − x)Na 2B 4O 7–<br />

0.5Bi 2O 3–0.5MoO 3, with 0 ≤ x ≤ 40 mol%, have been prepared by <strong>the</strong> melt quenching<br />

technique. Required quantities <strong>of</strong> Analar grade Bi 2O 3,MoO 3,Na 2CO 3, and H 3BO 3<br />

were mixed toge<strong>the</strong>r by grinding <strong>the</strong> mixture repeatedly to obtain a fine powder. The<br />

0925-8388/$ – see front matter © 2009 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.jallcom.2009.11.123


Y.B. Saddeek et al. / Journal <strong>of</strong> Alloys and Compounds 494 (2010) 210–213 211<br />

Fig. 1. X-ray diffraction patterns <strong>of</strong> <strong>the</strong> glass system (100 − x)Na 2B 4O 7–<br />

0.5Bi 2O 3–0.5MoO 3, with 0 ≤ x ≤ 40 mol%.<br />

Fig. 3. Tauc’s plots for <strong>the</strong> glass system (100 − x)Na 2B 4O 7–0.5Bi 2O 3–0.5MoO 3, with<br />

0 ≤ x ≤ 40 mol%.<br />

is given by<br />

˛h = A 0 (h − E opt<br />

g ) p (2)<br />

Fig. 2. Spectral behaviour <strong>of</strong> (a) transmittance and (b) reflectance <strong>of</strong> <strong>the</strong> glass system<br />

(100 − x)Na 2B 4O 7–0.5Bi 2O 3–0.5MoO 3, with 0 ≤ x ≤ 40 mol%.<br />

mixture was melted in a Porcelain crucible in an electrically heated furnace under<br />

ordinary atmospheric conditions at temperature about 1373 K for 2 h to homogenize<br />

<strong>the</strong> melt. The obtained glass samples were quenched into preheated stainless-steel<br />

mold preheated at a temperature <strong>of</strong> about 573 K for 2 h to remove any internal<br />

stresses. The samples were <strong>the</strong>n grounded and optically polished to have suitable<br />

dimensions (2 cm × 1cm× 0.15 cm) for <strong>the</strong> present optical measurements.<br />

X-ray measurements were performed to check <strong>the</strong> non-crystallinity <strong>of</strong> <strong>the</strong> glass<br />

samples using a Philips X-ray diffractometer PW/1710 with Ni-filtered, Cu K radiation<br />

( = 1.542 Å) powered at 40 kV and 30 mA. The XRD patterns <strong>of</strong> <strong>the</strong> <strong>glasses</strong> as<br />

shown in Fig. 1 did not reveal discrete or any sharp peaks, but <strong>the</strong> characteristic<br />

broad humps <strong>of</strong> <strong>the</strong> amorphous materials.<br />

The transmittance (T), and <strong>the</strong> reflectance (R) optical spectra <strong>of</strong> <strong>the</strong> <strong>glasses</strong><br />

(Fig. 2) were recorded at room temperature in <strong>the</strong> wavelength range 300–1100 nm<br />

using a computerized double beam spectrophotometer, type SHIMADZU UV-2100.<br />

where ˛0 is an energy-independent constant (band edge steepness<br />

parameter in Tauc’s picture [18]). For amorphous materials indirect<br />

transitions are valid according to Tauc [18], i.e. power part p = 2; so,<br />

<strong>the</strong> values <strong>of</strong> indirect optical band-gap energy (E opt<br />

g ) can be obtained<br />

from Eq. (2) by extrapolating <strong>the</strong> absorption coefficient to zero<br />

absorption in <strong>the</strong> (˛h) 1/2 –h plot as shown in Fig. 3. The respective<br />

values <strong>of</strong> E opt<br />

g are obtained by extrapolating to (˛h) 1/2 = 0 for <strong>the</strong><br />

indirect transitions [18,19]. The compositional dependence <strong>of</strong> E opt<br />

g<br />

shows a decrease trend as <strong>the</strong> Bi 2 O 3 and MoO 3 content as shown<br />

in Fig. 4. The decreasing values <strong>of</strong> E opt<br />

g can be understood in terms<br />

<strong>of</strong> <strong>the</strong> structural changes that are taking place in <strong>the</strong> studied glass<br />

system. According to <strong>the</strong> IR analysis [20], <strong>the</strong> stoichiometry <strong>of</strong> <strong>the</strong><br />

sodium diborate glass, by considering <strong>the</strong> association <strong>of</strong> Na 2 O with<br />

both MoO 3 and B 2 O 3 . Therefore, <strong>the</strong> BO 4 units will be destroyed and<br />

is converted into asymmetric BO 3 units with non-bridging oxygens.<br />

Also, <strong>the</strong> structure <strong>of</strong> <strong>the</strong> investigated <strong>glasses</strong> has distorted [BiO 6 ]<br />

octahedral units which shift <strong>the</strong> structural unit [BO 3 ] to lower wave<br />

numbers, and shifted <strong>the</strong> band <strong>of</strong> [BO 4 ] unit to higher wave number.<br />

This behaviour can be explained as to be due to <strong>the</strong> formation<br />

<strong>of</strong> new bridging bond <strong>of</strong> Bi–O–B. This new bond is formed due to <strong>the</strong><br />

inducement <strong>of</strong> <strong>the</strong> electrostatic field <strong>of</strong> <strong>the</strong> strongly polarizing Bi 3+<br />

ions. The stretching force constant <strong>of</strong> this bond tends to decrease,<br />

since <strong>the</strong> stretching force constant <strong>of</strong> Bi–O bonding is lower than<br />

that <strong>of</strong> <strong>the</strong> B–O bonding. Accordingly, <strong>the</strong> bond-length <strong>of</strong> Bi–O<br />

3. Results and discussion<br />

3.1. Determination <strong>of</strong> optical band gap<br />

The optical absorption coefficient, ˛, <strong>of</strong> a material can be evaluated<br />

from <strong>the</strong> optical transmittance, reflectance and <strong>the</strong> thickness<br />

<strong>of</strong> <strong>the</strong> sample (d) using <strong>the</strong> relation:<br />

˛ = 1 d ln 1 − R<br />

T<br />

(1)<br />

The absorption coefficient ˛() as a function <strong>of</strong> photon energy (h)<br />

for direct and indirect optical transitions, according to Pankove [17]<br />

Fig. 4. Dependence <strong>of</strong> <strong>the</strong> indirect optical band gaps <strong>of</strong> <strong>the</strong> glass system<br />

(100 − x)Na 2B 4O 7–0.5Bi 2O 3–0.5MoO 3, with 0 ≤ x ≤ 40 mol% on <strong>the</strong> parameter x.


212 Y.B. Saddeek et al. / Journal <strong>of</strong> Alloys and Compounds 494 (2010) 210–213<br />

dependence <strong>of</strong> <strong>the</strong> refractive indices <strong>of</strong> <strong>the</strong> studied <strong>glasses</strong> can be<br />

explained as follows. According to <strong>the</strong> Lorentz–Lorenz equation,<br />

<strong>the</strong> density <strong>of</strong> <strong>the</strong> material affects <strong>the</strong> refractive index in a direct<br />

proportion. Also, <strong>the</strong>re are some factors influence on <strong>the</strong> refractive<br />

index such as; <strong>the</strong> coordination number, Z, <strong>of</strong> <strong>the</strong> studied <strong>glasses</strong><br />

can explain its increase as <strong>the</strong> introducing <strong>of</strong> <strong>the</strong> structural units<br />

BiO 6 , and MoO 6 as stated previously, on <strong>the</strong> expense <strong>of</strong> ei<strong>the</strong>r BO 4<br />

or BO 3 will increase <strong>the</strong> coordination number <strong>of</strong> <strong>the</strong> <strong>glasses</strong>, and<br />

creating more NBOs, as noted in IR spectra [20]. This in turn leads<br />

to an increase in <strong>the</strong> average coordination number <strong>of</strong> <strong>the</strong> studied<br />

<strong>glasses</strong>, with respect to <strong>the</strong> base glass sample, which is increasing<br />

<strong>the</strong>ir index values. Also, <strong>the</strong> creation <strong>of</strong> NBOs creates more ionic<br />

bonds, which manifest <strong>the</strong>mselves in a larger polarizability over<br />

<strong>the</strong> mostly covalent bonds <strong>of</strong> bridging oxygen providing a higher<br />

index value.<br />

Fig. 5. The refractive index <strong>of</strong> (100 − x)Na 2B 4O 7–0.5Bi 2O 3–0.5MoO 3, with<br />

0 ≤ x ≤ 40 mol% <strong>glasses</strong> as a function <strong>of</strong> wavelength in a wavelength range from 400<br />

to 1100 nm.<br />

bonding increases, which in turn expand <strong>the</strong> glass network and<br />

replaces B–O linkage by <strong>the</strong> weaker Bi–O and Mo–O linkages. Thus,<br />

<strong>the</strong> available oxygen environment around <strong>the</strong> bismuth cations will<br />

increase, and this will decrease <strong>the</strong> values <strong>of</strong> E opt<br />

g .<br />

3.2. Determination <strong>of</strong> <strong>the</strong> refractive index<br />

According to <strong>the</strong> <strong>the</strong>ory <strong>of</strong> reflectivity <strong>of</strong> light, <strong>the</strong> refractive<br />

index (n) as a function <strong>of</strong> <strong>the</strong> reflectance (R), and <strong>the</strong> extinction<br />

coefficient (k) is given by <strong>the</strong> quadratic equation:<br />

R = (1 − n)2 + k 2<br />

(1 + n) 2 (3)<br />

+ k 2<br />

The extinction coefficient can be determined from <strong>the</strong> wavelength<br />

() and <strong>the</strong> calculated values <strong>of</strong> ˛ according to <strong>the</strong> relation:<br />

˛ = 4k<br />

(4)<br />

<br />

The refractive index can be well described by a modified Sellmeier<br />

dispersion relation (solid lines in Fig. 5):<br />

n() = A − B + C 2 − D 3 + E 4 (5)<br />

where A, B, C, D and E are <strong>the</strong> Sellmeier coefficients are averaged<br />

and given in Table 1 for all <strong>glasses</strong>.<br />

As shown in Eq. (5), <strong>the</strong> changes in <strong>the</strong> refractive index are<br />

affected by wavelength changes. Fig. 5 shows <strong>the</strong> plots <strong>of</strong> <strong>the</strong> refractive<br />

index as a function <strong>of</strong> <strong>the</strong> wavelength for <strong>the</strong> studied <strong>glasses</strong>.<br />

From this figure <strong>the</strong> refractive decrease with increase <strong>the</strong> wavelength<br />

and increase as <strong>the</strong> Bi 2 O 3 and MoO 3 content increases in<br />

<strong>the</strong> wavelength range <strong>of</strong> 200–2000 nm. Thus, <strong>the</strong> compositional<br />

3.3. Polarizability and optical basicity <strong>of</strong> <strong>the</strong> <strong>glasses</strong><br />

The average electronic polarizability <strong>of</strong> ions and <strong>the</strong> optical<br />

basicity are considered to be from <strong>the</strong> most important <strong>properties</strong> <strong>of</strong><br />

<strong>the</strong> materials, which is closely related to <strong>the</strong>ir applicability in <strong>the</strong><br />

field <strong>of</strong> optics and electronics. The optical basicity proposed was<br />

used as a measurement <strong>of</strong> acid–base <strong>properties</strong> <strong>of</strong> <strong>the</strong> oxide <strong>glasses</strong>,<br />

and is expressed in terms <strong>of</strong> <strong>the</strong> electron density carried by oxygen.<br />

It was found that, optical non-linearity is caused by <strong>the</strong> electronic<br />

polarization <strong>of</strong> a material upon its exposure to intense light beams,<br />

so, <strong>the</strong> non-linear response <strong>of</strong> <strong>the</strong> material is governed by <strong>the</strong> electronic<br />

polarizability. For this purpose, materials with higher optical<br />

non-linearity have to be found or designed on <strong>the</strong> basis <strong>of</strong> <strong>the</strong> correlation<br />

between <strong>the</strong> optical non-linearity and with some o<strong>the</strong>r easily<br />

understandable and accessible electronic <strong>properties</strong> [21–24].<br />

For isotropic substance such as <strong>glasses</strong>, <strong>the</strong> average molar refraction<br />

(R m ) was given by <strong>the</strong> Lorentz–Lorenz equation:<br />

[ ] [ ]<br />

n 2 − 1 M n 2 R m =<br />

n 2 + 2 = − 1<br />

n 2 V m (6)<br />

+ 2<br />

where <strong>the</strong> quantity (n 2 − 1)/(n 2 + 2) is called <strong>the</strong> reflection loss<br />

[24,25]. The molar refraction is related to <strong>the</strong> structure <strong>of</strong> <strong>the</strong><br />

glass and it is proportional to <strong>the</strong> molar electronic polarizability<br />

<strong>of</strong> <strong>the</strong> material (˛m) (in ×10 −24 cm 3 ) through <strong>the</strong> following<br />

Clasius–Mosotti relation:<br />

( 3<br />

)<br />

˛m = R m (7)<br />

4N<br />

where N is Avogadro’s number. For various ternary oxide <strong>glasses</strong><br />

with <strong>the</strong> general formula y 1 A p O q –y 2 B r O s –y 3 C n O m , where y’s<br />

denote <strong>the</strong> molar fraction <strong>of</strong> each oxide, <strong>the</strong> electronic polarizability<br />

<strong>of</strong> oxide ion can be calculated on <strong>the</strong> basis <strong>of</strong> refractive indices<br />

using <strong>the</strong> following equation:<br />

[ ∑ ]<br />

(Vm ˛O2−(n) /2.52)((n 2 − 1)/(n 2 + 2)) − ˛cat<br />

=<br />

(8)<br />

N O<br />

2−<br />

Table 1<br />

∑<br />

The chemical composition, <strong>the</strong> refractive index parameters, <strong>the</strong> molar cation polarizability ˛cat and <strong>the</strong> number <strong>of</strong> oxide ions in <strong>the</strong> chemical formula (N O 2− ).<br />

∑<br />

Composition (mol%)<br />

Refractive index parameters<br />

˛cat<br />

N O 2−<br />

A B C (×10 6 ) D (×10 9 ) E (×10 11 )<br />

0 2.59 2203.46 4.19 2.64 5.72 0.06 2.34<br />

5 2.60 2145.27 4.35 2.84 6.41 0.14 2.37<br />

10 2.99 3627.31 6.72 4.36 9.86 0.21 2.41<br />

15 3.33 4815.91 8.58 5.50 12.36 0.29 2.44<br />

20 3.10 3698.12 7.37 5.04 12.1 0.37 2.47<br />

25 2.48 1232.81 3.96 3.08 8.15 0.44 2.51<br />

30 2.87 2802.81 7.24 5.74 16.01 0.52 2.54<br />

35 2.38 737.817 3.59 3.09 8.96 0.60 2.57<br />

40 3.61 6072.79 12.9 9.80 26.59 0.67 2.60


Y.B. Saddeek et al. / Journal <strong>of</strong> Alloys and Compounds 494 (2010) 210–213 213<br />

A = 1.67<br />

(<br />

1 − 1<br />

˛O2−<br />

)<br />

(9)<br />

Fig. 6. The oxide ion polarizability ˛O2− <strong>of</strong> (100 − x)Na 2B 4O 7–0.5Bi 2O 3–0.5MoO 3,<br />

with 0 ≤ x ≤ 40 mol% <strong>glasses</strong> as a function <strong>of</strong> wavelength in a wavelength range from<br />

400 to 1100 nm.<br />

The A values were estimated according to Eq. (8). Fig. 7 shows a<br />

plot <strong>of</strong> optical basicity versus <strong>the</strong> wavelength for different glassy<br />

compositions. According to Fig. 7, <strong>the</strong> values <strong>of</strong> A decrease with an<br />

increase in <strong>the</strong> wavelength, and increase with an increase in <strong>the</strong><br />

Bi 2 O 3 and MoO 3 content (at any value <strong>of</strong> <strong>the</strong> wavelength). Zhao et<br />

al. [28] reported that B 2 O 3 is a strong acidic oxide with low optical<br />

basicity (0.42), while Bi 2 O 3 is an oxide with a significant basicity<br />

(1.19). The increased optical basicity <strong>of</strong> <strong>the</strong> <strong>glasses</strong> with large Bi 2 O 3<br />

and MoO 3 content indicates that <strong>the</strong> acidic oxide <strong>of</strong> Bi 2 O 3 has a<br />

significant effect. The low optical basicity means a reduced ability<br />

<strong>of</strong> oxide ions to transfer electrons to <strong>the</strong> surrounding cations. On<br />

<strong>the</strong> o<strong>the</strong>r hand, borate <strong>glasses</strong> with a large amount <strong>of</strong> Bi 2 O 3 content<br />

possess high optical basicity. High optical basicity means low<br />

electron donor ability <strong>of</strong> <strong>the</strong> oxide ions.<br />

4. Conclusions<br />

Increasing <strong>the</strong> Bi 2 O 3 and MoO 3 content on <strong>the</strong> expense <strong>of</strong><br />

Na 2 B 4 O 7 in <strong>the</strong> studied glass system, reveals some remarkable features;<br />

as <strong>the</strong> Bi 2 O 3 and MoO 3 content increase, more NBOs were<br />

created and <strong>the</strong> bond-length <strong>of</strong> <strong>the</strong> borate structural units increase<br />

which in its turn increase <strong>the</strong> refractive indices, <strong>the</strong> polarizability,<br />

and <strong>the</strong> optical basicity <strong>of</strong> <strong>the</strong> studied <strong>glasses</strong>. Also, <strong>the</strong> observed<br />

decrease in <strong>the</strong> optical band gap was related to <strong>the</strong> weaker bond<br />

strength <strong>of</strong> <strong>the</strong> Bi–O compared with that <strong>of</strong> B–O.<br />

Acknowledgements<br />

The authors thank Dr. Shaban I Hussien, Egyptian Atomic Energy<br />

Authority for his help in preparing <strong>the</strong> samples.<br />

References<br />

Fig. 7. <strong>Optical</strong> basicity for <strong>the</strong> (100 − x)Na 2B 4O 7–0.5Bi 2O 3–0.5MoO 3, with<br />

0 ≤ x ≤ 40 mol% <strong>glasses</strong>.<br />

where ˛O2−(n) <strong>the</strong> ∑ electronic polarizability <strong>of</strong> oxide ion based on<br />

refractive indices. ˛cat denotes <strong>the</strong> molar cation polarizability<br />

given by y 1 p˛A + y 2 r˛B + y 3 n˛C and N O<br />

2− denotes <strong>the</strong> number <strong>of</strong><br />

oxide ions in <strong>the</strong> chemical formula ∑ given by y 1 q + y 2 s + y 3 m. For <strong>the</strong><br />

studied <strong>glasses</strong>, <strong>the</strong> values <strong>of</strong> both ˛cat and N O<br />

2− were calculated<br />

using <strong>the</strong> values <strong>of</strong> ˛Na = 0.0291 Å 3 for Na + ions, <strong>of</strong> ˛B = 0.002 Å 3<br />

for B 3+ , ˛Bi = 1.508 Å 3 for Bi 3+ ∑ ions and <strong>of</strong> ˛Mo = 0.169 Å 3 for Mo 3+<br />

[26]. The calculated values ˛cat and N O<br />

2− are listed in Table 1.<br />

Then <strong>the</strong> electronic polarizability ˛O2−(n) for different glass compositions<br />

easily calculated by using Eq. (8) and figured as a function<br />

<strong>of</strong> wavelength as shown in Fig. 6, this figure shows that, <strong>the</strong> ˛O2− (n)<br />

decreases with <strong>the</strong> increase in <strong>the</strong> wavelength and increases with<br />

<strong>the</strong> increase in Bi 2 O 3 and MoO 3 content. It was reported that Bi 3+<br />

cation possesses a very high polarizability, which is due to its large<br />

ionic radii and small cation unit field strength. Moreover, Bi 3+ ions<br />

possess a lone pair in <strong>the</strong> valence shell. So, <strong>the</strong> increase in <strong>the</strong> average<br />

polarizability <strong>of</strong> <strong>the</strong> studied <strong>glasses</strong>, on a molar basis, can be<br />

attributed to <strong>the</strong> replacement <strong>of</strong> a less polarizable boron oxide by<br />

a highly polarizable oxides [27,28].<br />

An alternative approach for <strong>the</strong> calculation <strong>of</strong> <strong>the</strong> optical basicity<br />

is <strong>the</strong> determination <strong>of</strong> <strong>the</strong> state <strong>of</strong> polarization <strong>of</strong> oxide ions in a<br />

glass matrix on <strong>the</strong> basis <strong>of</strong> refraction data. Duffy [21,22,29] has<br />

established <strong>the</strong> following correlation:<br />

[1] K. Rao, Structural Chemistry <strong>of</strong> Glasses, Elsevier, North Holland, 2002.<br />

[2] A. Wells, Structural Inorganic Chemistry, 4th ed., Clarendon Press, Oxford, 1975.<br />

[3] W. Vogel, Chemistry <strong>of</strong> Glass, American Ceramic Society, Westerville, OH, 1985.<br />

[4] G. Sharma, K. Thind, Monika, H. Singh, Manupriya, L. Gerward, Phys. Status<br />

Solidi (A) 204 (2007) 591.<br />

[5] I. Oprea, H. Hesse, K. Betzler, Phys. Status Solidi (B) 242 (2005) R109.<br />

[6] I. Kityk, A. Majchrowski, Opt. Mater. 25 (2004) 33.<br />

[7] M. Kodama, T. Matsushita, S. Kojima, Jpn. J. Appl. Phys. 34 (1995) 2570.<br />

[8] M. Shibata, C. Sanchez, H. Patel, S. Feller, J. Stark, G. Sumcad, J. Kasper, J. Non-<br />

Cryst. Solids 85 (1986) 29.<br />

[9] S. Feller, W. Dell, P. Bray, J. Non-Cryst. Solids 51 (1982) 21.<br />

[10] K. El-Egili, A. Oraby, J. Phys.: Condens. Matter 8 (1996) 8959.<br />

[11] E.I. Kamitsos, G.D. Chryssikos, Solid State Ionics 105 (1998) 75.<br />

[12] Y. Moustafa, A. Hassan, G. El-Damrawi, N. Yevtushenko, J. Non-Cryst. Solids 194<br />

(1996) 34.<br />

[13] C. Reddy, R. Chakradhar, Mater. Res. Bull. 42 (2007) 1337.<br />

[14] V. Gowda, R. Anavekar, Ionics 10 (2004) 103.<br />

[15] B. Chowdari, Z. Rong, Solid State Ionics 90 (1996) 151.<br />

[16] M. Ganguli, K. Rao, J. Solid State Chem. 145 (1999) 65.<br />

[17] J. Pankove, <strong>Optical</strong> Processes in Semiconductors, Prentice-Hall, Englewood<br />

Cliffs, NJ, 1971.<br />

[18] J. Tauc, Amorphous and Liquid Semiconductor, Plenum Press, New York, 1974.<br />

[19] M. Nocun, W. Mozgawaa, J. Jedliński, J. Najman, J. Mol. Struct. 744–747 (2005)<br />

603.<br />

[20] Y. Saddeek, A. Abousehly, S. Hussien, J. Phys. D: Appl. Phys. 40 (2007) 4674.<br />

[21] J. Duffy, M. Ingram, J. Non-Cryst. Solids 21 (1976) 373.<br />

[22] J. Duffy, M. Ingram, J. Am. Chem. Soc. 93 (1971) 6448.<br />

[23] Y. Saddeek, E. Shaaban, E. Moustafa, H. Moustafa, Physica B 403 (2008) 2399.<br />

[24] E. Moustafa, Y. Saddeek, E. Shaaban, J. Phys. Chem. Solids 69 (2008) 2281.<br />

[25] M. Abdel-Baki, F. Abdel-Wahab, A. Radi, F. El-Diasty, J. Phys. Chem. Solids (2007)<br />

1016.<br />

[26] V. Dimitrov, S. Sakka, J. Appl. Phys. 79 (1996) 1736.<br />

[27] T. Honma, R. Sato, Y. Benino, T. Komatsu, V. Dimitrov, J. Non-Cryst. Solids 272<br />

(2000) 1.<br />

[28] X. Zhao, X. Wang, H. Lin, Z. Wang, Physica B 390 (2007) 293.<br />

[29] J. Duffy, Phys. Chem. Glasses 320 (1989) 1.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!