infrared and raman spectroscopy -...

38
CHAPTER 5 INFRARED AND RAMAN SPECTROSCOPY

Upload: buihanh

Post on 30-Jul-2018

239 views

Category:

Documents


0 download

TRANSCRIPT

CHAPTER 5

INFRARED AND RAMAN SPECTROSCOPY

125

5.1 INTRODUCTION TO INFRARED SPECTROSCOPY

Spectroscopy is the study of the interaction of electromagnetic radiation with a

chemical substance. The nature of the interaction depends upon the properties of the

substance. When radiation passes through a sample (solid, liquid or glass), certain

frequencies are absorbed which are unique for each molecule and is the characteristic

of a substance.

Infrared spectroscopy (IR) is one of the most useful techniques available for

structural studies of glasses. For the particular case of glasses modified by metal

oxides, IR is a powerful tool because it leads to structural aspects related to both the

local units constituting the glass network and the anionic sites hosting the modifying

metal cations. Borate glasses provide an ideal case in comparison to other glass

forming systems, to demonstrate the effectiveness of infrared spectroscopy in glass

science.

Infrared reflectance and transmission measurements are the most suitable

techniques for glass studies among the various sampling techniques employed in

infrared measurements. A key advantage of reflectance spectroscopy is the use of the

same sample for data acquisition over a broad and continuous frequency covering

both mid and far infrared region, without the need of changing sample form or its

thickness, a problem usually encountered in transmission measurements. The spectral

profiles obtained from reflectance studies are free of band shape distortions, which are

present in transmission spectra. IR method utilizes the optical excitation of the

localized vibrational modes of atoms, whose excitation energy is in the infrared

region. The transitions involved in Infrared absorption are associated with the

vibrational changes within the molecule. Different bonds have different vibrational

126

frequencies, and one can detect the presence of bonds in a compound by identifying

the characteristic frequency as an absorption band in the infrared spectrum. Infrared

spectroscopy is broadly divided into three regions as

1. Near infrared region (13,000 – 4,000 cm-1)

2. Mid infrared region (4,000 – 200 cm-1)

3. Far infrared region (200 – 10 cm-1)

Infrared spectra are usually plotted as percent of transmittance rather than

absorbance the ordinate. This makes absorption bands as dips in the curve rather than

as maxima in the case of UV and visible spectra. No two compounds except optical

isomers can have identical absorption bands.

To interpret the infrared spectra an understanding of the energy levels in the

molecule is required. The vibrational and rotational motions are quantized in a

molecule between the atoms. By the absorption of suitable energy the molecule

undergoes transition to higher quantized energy levels which causes an absorption in

the spectrum.

Infrared transmittance spectroscopy is the most suitable technique for glass

studies. Infrared spectra are usually recorded by measuring the transmittance of light

quanta through a continuous distribution of the sample. Interaction of infrared

radiation with a vibrating molecule is only possible if the electric vector of the

radiation field oscillates with the same frequency as does the molecular dipole

moment. A vibration is infrared active only if the molecular dipole moment (µ) is

modulated by the normal vibration,

0≠

∂∂

oqµ (5.1)

127

where q stands for the normal coordinate describing the motion of the atoms during a

normal vibration. If this condition is fulfilled, then the vibrations are said to be

allowed or active in the infrared spectrum, otherwise they are said to be forbidden or

inactive. The frequencies of the absorption bands are proportional to the energy

difference between the vibrational ground and excited states.

5.2 FUNDAMENTAL VIBRATIONS OF MOLECULES

Each molecule has certain natural vibrational frequencies. When light is

incident on the molecule, the frequency which matches the natural vibrational

frequency is absorbed by the molecule resulting in molecular vibrations.

Modes of vibrations

Stretching: Distance between two atoms increases or decreases

Bending: Position of the atom changes relative to the original bond axis

To interpret the IR spectra an understanding of the energy levels in the

molecule is required. The vibrational and rotational motions are quantized in a

molecule between the atoms. By the absorption of suitable energy, the molecule

undergoes transition to higher quantized energy levels which causes an absorption in

the spectrum. IR spectroscopy is the most suitable technique for glass studies. For a

particular case of glasses modified by metal oxides, infrared technique is a powerful

tool because it gives information about structural aspects related to both the local

units constituting the glass network and the anionic sited hosting the modifying metal

cations.

For a molecule to absorb IR radiation, it has to fulfill certain requirements

which are as follows:

a) Correct Wavelength of Radiation

128

Molecule absorbs radiation only when the natural frequency of vibration of

some part of a molecule (i.e., atoms or group of atoms comprising it) is same as the

frequency of radiation. After absorbing correct wavelength of radiation, the molecule

vibrates at increased amplitude. This occurs at the expense of the energy of IR

radiation that has been absorbed.

b) Electric dipole

This is another condition for a molecule to absorb IR radiation. A molecule

can only absorb IR radiation when its absorption causes a change in its electric dipole

moment (μ). A molecule is said to be electric dipole when there is a slight positive

and a slight negative electric charge on its component atoms.

When a molecule having electric dipole is kept in the electric filed (as in the

case when the molecule is kept in a beam of IR radiation), the field will exert forces

on the electric charges in the molecule. Opposite charges will experience force in

opposite directions. This tends to decrease separation. As the electric field of the IR

radiation is changing its polarity periodically, it means that the spacing between the

charged atoms (electric dipoles) of molecule also changes periodically. When the

vibration in the charged atoms is fast, the absorption of radiation is intense and thus,

the IR spectrum will have intense absorption bands. On the other hand, when the rate

of vibration of charged atoms in atoms in a molecule is slow, there will be weak

bands in the IR spectrum.

5.3 INTRODUCTION TO RAMAN SPECTROSCOPY

Raman spectroscopy has been successfully used for the structure

determination in crystalline materials. Due to the absence of symmetry in glasses, the

vibration analysis in these materials is not straightforward. In fact, no agreed theory of

129

vibration exists to explain all the aspects of vibrations in disordered materials even

though several steps in that direction do exist [1,2]. The method due to Brawer [3]

considers disorder to be a perturbation on the vibration of a structural group and hence

considers the width of the Raman peak as a measure of the disorder. The intensity of

Raman lines is a useful parameter since it is proportional to the number of scattering

centers and their scattering efficiency.

Raman spectroscopy is a measurement of the wavelength and intensity of in

elastically scattered light from molecules. A molecule with no Raman-active modes

absorbs a photon with the frequency vo. The excited molecule returns back to the

same basic vibrational state and emits light with the same frequency vo as an

excitation source. This type of interaction is called an elastic Rayleigh scattering. A

Photon with frequency vo is absorbed by Raman-active molecule which at the time of

interaction is in the basic vibrational state. Part of the photon’s energy is transferred to

the Raman-active mode with frequency vm and the resulting frequency of scattered

light is reduced to vo- vm. This frequency is called Stokes frequency or “stokes”. A

Photon with frequency vo is absorbed by a Raman-active molecule, which at the time

of interaction, is already in the excited vibrational state. Excessive energy of excited

Raman-active mode is released, molecule returns to the basic vibrational state and the

resulting frequency of scattered light goes up to vo+ vm. This frequency is called Anti

strokes frequency or “Anti-Stokes”.

When a molecule is exposed to an electric field, electrons and nuclei are

forced to move in opposite directions. Thus, a dipole moment, which is proportional

to the electric field strength and to the molecular polarizability α, is induced. A

molecular vibration can be observed in the Raman spectrum only if there is a

modulation of the molecular polarizability by the vibration,

130

0≠

∂∂

oqα (5.2)

If this condition is fulfilled, then the vibrations are said to be allowed or active in the

Raman spectrum, otherwise they are said to be forbidden or inactive.

5.4 REVIEW OF EARLIER WORK

A brief review of the earlier work carried out on IR and Raman studies of

glasses is presented.

Kamitsos et al [4,5] studied the structure of cesium, rubidium and potassium

borate glasses by Raman and far infrared spectroscopy. It was demonstrated that the

creation of non-bridging oxygens (NBOs) even at very low modification levels of

alkali oxide. Konijnendijk [6] used Raman spectroscopy to interpret the molecular

structure of glasses in the systems CaO-Na2O-B2O3 and MgO-Na2O-B2O3 systems.

From experimental results it was confirmed that mainly magnesium connected BO4

tetrahedra and unconnected BO4 tetrahedra are formed.

Krogh-Moe [7] studied the infrared spectra of boron oxide and alkali borate

glasses extensively. It was concluded that the crystalline and vitreous system consists

of the same structural units and proposed a structural model. The borate glasses

mainly consists of boroxol rings, tetra borate and diborate groups.

Raman spectroscopy has been used effectively to study the formation of poly

borate groups in xR2O-(100-x) B2O3 (R = Li, Na, K, Rb & Cs) [8-11] and also in

xR2O-(100-x) B2O3 (R = Ba, Ca, Sr, Mg, Cd and Pb) [12-14,6]. Meera et al [15]

studied the Raman spectra of ZnO-B2O3 glasses. The role of ZnO in network

131

modification is not clearly understood, although the Raman data is consistent with the

glass forming tendency of ZnO.

Bale et al [16-21] reported spectroscopic studies such as EPR, Raman,

Infrared and optical absorption of pure Li2O–Bi2O3–B2O3–ZnO glasses and doped

with Cu2+ ions. The non-linear variation of the spin-Hamiltonian parameters with ZnO

content was attributed to the mixed former effect. This effect is also evident in the

optical absorption spectra. By correlating the EPR and optical absorption data, the

bonding parameters α2, β2 and β12 have been evaluated and correlated with optical

basicity. In addition, the bonding parameters also varied non-linearly suggesting a

mixed former effect. IR and Raman spectra shows that these glasses are made up of

[BiO3] pyramidal and [BiO6] octahedral units. The formation of Zn2+ in tetrahedral

coordination was observed. The average electronic polarizability of the oxide ion,

optical basicity and Yamashita-Kurosawa’s interaction parameter were also

examined.

Vijaya Kumar et al [22] determined theoretical basicity of Bi2O3-BaO-B2O3,

glasses. The infrared studies revealed that theese glasses are made up of [BO3], [BO4]

and [BiO6] structural units. Srinivasu et al [23] investigated bismuth based glasses

containing LiF, Li2O and SrO by different physical and spectroscopic techniques.

Infrared and Raman spectroscopic results indicate that the glass network consists of

BiO6 octahedral and BiO3 pyramidal units.

Rao et al [24] prepared and studied FT-IR and Raman spectra of Sm3+ and

Nd3+ co-doped magnesium lead borosilicate glasses. The FT-IR, Raman spectra

reveals the nature of bonding situation and different structural units in glass network.

132

Ivascu et al [25,26] reported FT-IR, Raman and thermo luminescence

investigations in ternary lithium phosphate glasses. FT-IR and Raman spectroscopy

studies revealed a local network structure mainly based on Q2 and Q3 tetrahedrons

connected by P-O-P linkages. Luminescence investigations show that by adding

modifier oxides to phosphate glass dose dependent TL signals results upon irradiation.

Thus P2O5-BaO-Li2O glass system is a possible candidate material for dosimetry in

the high range (>10 Gy). ESR study proves that the investigated glass system could be

successfully used as ESR dosimeter in medicine and industry.

Sodium zinc borate and lead bismuth borate glasses containing Nd3+ were

investigated by Karthikeyan et al [27,28] by FTIR, optical absorption and emission

studies. From FTIR spectral studies they concluded that the glass contains BO3, BO4

and ZnO4 units and zinc behaves as the network modifier. The formations of ZnO4

tighten the glass structure and this was confirmed by the nephelauxetic effect.

ZnO-K2O-B2O3-P2O5 glasses formulated with Sb2O3 were investigated by

Raman and NMR studies. Raman spectra with increasing Sb2O3 content reflect the

depolymerization of phosphate chains and NMR spectra reveal a steady

transformation of BO4 into BO3 units. Potassium zinc borophasphate glasses were

prepared and studied by NMR and Raman spectroscopy [29, 30].

FT-IR and EPR investigations have been done on CuO-B2O3-Bi2O3 glasses by

Ardelean et al [31]. They have reported that a part of bismuth ions are incorporated in

the glass network as [BiO6] octahedral units depends on CuO content and for all

concentration range, the [BO3] units are dominant. Both the structural units [BO3] and

[BO4] are depends on the CuO content.

133

Properties of unconventional lithium bismuthate glasses were investigated by

Hazra et al [32,33] by IR and Raman and optical absorption techniques. From the

Raman data they concluded the presence of [BiO6] octahedral units. Increase of

optical band gap and the increase of IR cutoff frequency with lithium content is

attributed to the decrease in the strength of the glass structure.

Structure and crystallization kinetics of Bi2O3 and B2O3 glasses were

investigated by Cheng et al [34] using IR and DSC. The composition dependence of

IR absorption suggests that addition of Bi2O3 results in a change in the short-range

order structure of the borate matrix. The increase of Bi2O3 content causes a

progressive conversion of [BO3] to [BO4] units. Bi2O3 exists in the form of [BiO6].

Upender et al [35] carried out infrared, Raman, electron paramagnetic resonance and

optical absorption studies on Li2O-P2O5-TeO2-CuO glasses. Both Raman and IR

studies showed that the present glass system consists of [TeO3], [TeO4], [PO3] and

[PO4] units. The spin-Hamiltonian parameters have been determined from EPR

spectra and it was found that the Cu2+ ion is present in tetragonal distorted octahedral

site with dx2-y2 as the ground state. Bonding parameters and bonding symmetry of Cu2+

ions have been calculated by correlating EPR and optical data and were found to be

composition dependent.

Infrared spectra of Na2O-B2O3-SiO2 and Al2O3-Na2O-B2O3-SiO2 glasses have

been analyzed to calculate the fraction N4 of four coordinated borons. A reasonable

agreement between the calculated from IR spectra and those determined from NMR

spectroscopy could be attained under certain condition. It has been proposed that the

absorption bands in the region 1000-1120 cm-1 arise from contribution of SiO2 and

B2O3 vibrations. Heat treatment of Na2O-B2O3-SiO2 glasses does not change the value

134

of N4 and this indicates that the structure of alkali borate phase is the same in the glass

obtained from melt and in the heat-treated one [36].

Krishna Kumari et al [37] investigated mixed alkali zinc borate glasses by

UV-VIS absorption, EPR and FT-IR spectroscopic studies. FT-IR measurements of

all glasses revealed that the network structure of glass system are mainly based on

BO3 and BO4 units placed in different structural groups in which the BO3 units being

dominant. The EPR spectra of Mn2+ ions doped glasses exhibited a characteristic

hyperfine sextet around g = 2. The spectroscopic analysis of the obtained results

confirmed near octahedral site symmetry for the Mn2+ impurity ions. Crystal field and

Racah parameters are evaluated from optical absorption spectra. The optical band gap

and Urbach energies are determined the mixed alkali effect.

Risen et al [38,39] reported far infrared and Raman spectra of several series

of single, mixed alkali meta phosphate and mixed penta silicate glasses in both the

annealed and annealed forms. The frequencies of the cation-motion bands in the far

infrared and Raman spectra, which correspond to cation site vibrations, do not shift

with alkali content, including that the vibrationally significant local geometry and

forces associated with a particular cation are unaffected by the introduction of the

second cation into the glass structure or by annealing. A simple vibrational model is

presented which shows that the cation- dependent shifts are due to small changes in

the network bond angles and variation of the cation site forces.

Ahamed et al [40,41] prepared and characterized lead containing barium zinc

lithium fluoroborate glasses doped with different concentrations of trivalent Dy3+ and

Sm3+ ions through the XRD, DSC, FTIR, Raman, optical absorption,

photoluminescence and decay curve analysis. Coexistence of trigonal BO3, tetrahedral

135

BO4 units non-bridging oxygen and strong OH bonds was evidenced by IR and

Raman spectroscopy. The bonding parameters and the oscillator strengths were

determined from the absorption spectra. It was proved that the present glass is more

suitable for generation of white light for blue LED chips.

Srinivasa Rao et al [42, 43] reported, for the first time the study of mixed

alkali effect (MAE) in boroarsenate and borobismuthate glasses through density,

DSC, DC electrical conductivity and IR studies. The strength of the MAE in Tg. DC

electrical conductivity and activation energy has been determined. It was observed

that the strength of MAE in DC electrical conductivity is less pronounced with

increase in temperature, supporting molecular dynamic results. The IR studies shows

that the glass system contains BO3 and BO4 units in the disorder manner. Soppe et al

[44] observed the mixed alkali effect in Raman spectra of Li2O-Cs2O-B2O3 glasses.

Kamitsos et al [45] employed Infrared and Raman study to investigate the

structure of borate glasses. The results presented for single alkali glasses illustrate the

strong dependence of the network structure on the nature and content of the oxide

modifier. Variations of the cation-motion frequencies in the far-infrared spectra of

mixed alkali glasses have been interpreted as suggesting changes in the alkali-oxygen

interaction upon alkali mixing.

Kistaiah et al [46,47] reported mixed alkali bismuth borate glasses of

composition Li2O-K2O-Bi2O3-B2O3:V2O5. The spectroscopic properties of glass

samples were studied using infrared and Raman spectroscopic techniques. Acting as

complimentary spectroscopic techniques, both types of measurements IR and Raman

revealed that the network structure of the studied glasses is mainly based on BO3 and

BO4 units placed in different structural groups, BO3 units being dominate and bismuth

136

exists as BiO3 and BiO6 octahedral units. The observation of disappearance and

reappearance of some IR and Raman bands and non-linear variation of the peak

positions of some of these bands with alkali content is an important result pertaining

to the mixed alkali effect in this glass system.

Infrared spectra of mixed-alkali diborate glasses, Li2O-Na2O-B2O3 and Li2O-

K2O-B2O3, have been investigated by Selvaraj and Rao [48]. B-O stretching and B-O-

B bending frequencies exhibit nonlinear shifts which can be described as a mild

mixed alkai effect. Both shifts and nonlinear variations of frequencies may be

explained on the basis of alteration of the structure of the diborate groups in the

presence of Li+ ions.

Efimov [49] discussed various methods for the quantitative analysis of the IR

and Raman spectra of various inorganic glasses to determine physically meaningful

optical functions and individual band parameters. Resent data on band intensities or

frequencies obtained with these methods for phosphate, borate and germinate glasses

were considered. Trends in the IR and Raman band assignments deduced from the

current state of vibrational spectroscopy of glasses were analyzed and structural

information obtained by different authors for binary phosphate, borate and germinate

glasses was compared.

Sharada and Suresh babu [50] prepared Li2O-B2O3-Ta2O5-Bi2O3 glasses via

normal melt quenching technique and these glasses were characterized by FTIR,

optical absorption and AC conductivity studies. FTIR spectra of the samples recorded

in the frequency range 400-1500cm-1 exhibited characteristic bands corresponding to

BO3, BO4 stretching vibrations and BO bending vibrations. Tightening of the

structure is indicated by increase in the vibration of BO3 at the cost of BO4.

137

Mansour [51,52] reported infrared absorption study of Al2O3-PbO-B2O3-SiO2

and Li2O-CeO2-B2O3 glasses. The IR band located near 700 cm-1 was suggested to be

due to the vibrations of bridging oxygens between trigonal boron atoms. The

depolymerization of the whole glass skeleton increases with increasing the content of

Al2O3. There is a competition between the role of PbO and Al2O3 in changing the

value of N4 and the cross-linking of the glass network. An essential change in the role

of PbO (CeO2) in these glasses from glass modifier to glass former was also

suggested.

Raghavendra Rao et al [53,54] studied and correlated the physical and

structural properties of Co2+ (Ni2+) doped ZnO-Li2O-Na2O-B2O3 glasses. The

structural parameters of all the glasses are evaluated and a non-linear behavior is

observed. FT-IR spectra of ZLNB glasses reveal diborate units in borate network. The

optical absorption spectra suggest the site symmetry of Co2+ in the glasses is near

octahedral. Crystal field and inter-electronic repulsion parameters are also evaluated.

The optical band gap and Urbach energies exhibit the mixed alkali effect.

Naresh and Buddhudu [55] carried out optical absorption, FT-IR and Raman

spectral studies on transparent and stable glasses in the chemical composition of Li2O-

LiF-B2O3-MO (M = Zn and Cd) glasses. The LFB glasses with the presence of ZnO

and CdO an extended UV- transmission ability has been achieved. The measured FT-

IR and Raman spectra have exhibited the vibrational bands of B-O from [BO3] and

[BO4] units and Li-O bonds.

Infrared reflectance spectra of B2O3-Li2O-Cs2O glasses [56] have been

measured in the frequency range of 10-5000 cm-1. The mid-infrared parts of the

spectra are discussed in connection with B-O network vibrational modes. The far-

138

infrared parts of the spectra yields cation vibrational modes, which distinctly depend

on the glass composition.

Kashif et al [57] studied the effect of heat treatment and doping the transition

metal on LiNbO3 and LiNb3O8 nano-crystallite phases in lithium borate glass system

through XRD and FTIR studies. The FTIR data propose for these glasses and heat-

treated glass network structures mainly built by: di-, tri-, tetra-, penta- and ortho-

borate groups. It was found that the quntitative evaluation of these various borate

species in the glass structures was influenced by the transition metal. A detailed

discussion relating to the N4 evaluation with the transition metal content was made.

FTIR spectra of three MgO-PbO-B2O3 glass series have been analyzed by

Doweidar et al [58]. There is a decrease in the fraction of N4 of four coordinated

boron with increasing the MgO content, at the expanse of PbO. A new technique has

been presented to make use of the N4 data and follow the change in the modifier and

former fractions of PbO and MgO. These fractions change markedly, at different

rates, with the glass composition. The ability of the glass to include MgO increases

with increasing PbO content.

Martino et al [59] reported polarized Raman and infrared absorption studies

for GeO2-M2O (M = Na and Cs) glasses, as a function of the alkali content. Infrared

reflectivity measurement confirm the presence of a fraction of higher coordinated Ge

atoms, either five-fold or six-fold.

Gaafar et al [60] investigated Na2O-B2O3-P2O5-Fe2O3 glasses prepared by the

melt quenching technique. Elastic properties and FT-IR spectroscopic studies have

been employed to study the role of P2O5 on the structure of the glass system. Infrared

139

spectra of the glasses reveal that the borate network consist of diborate units and is

affected by the increase in the concentration of P2O5 content as a second network

former. These results are interpreted in term of the replacement of the borate units

with B-O-B bridges by phosphate units with non-bridging oxygens (NBOs).

Gajanan et al [61] investigated room temperature Raman specra on Oxyfluro

Vanadate glass containing various proportions of lithium fluoride and rubidium

fluoride to see an effect of mixture of alkali on vanadium-oxygen bond length. The

variation in bond length and its distribution about a most probable values was

correlated to the alkali environment present in these glasses. They observed that all

rubidium environment around the network forming units is more homogenous than all

lithium environment.

Akagi et al [62] studied the structure of K2O-B2O3 glasses and melts by high-

temperature Raman spectroscopy. With an increase in the K2O content and with

increasing temperature, the boroxal rings, which only consist of BO3 triangular units,

were converted into pentaborate groups which consist of BO4 tetrahedral (O =

bridging oxygen atom) and BO3 units. The fraction of four-coordinated boron atoms,

N4, obtained from deconvolution of the Raman bands was gradually reduced above

the glass transition temperature and converged to a constant value over 1400K.

Silver oxide doped lead lithium borate (LLB) glasses have been prepared and

characterized by XRD, SEM, EDS, FTIR and Raman[63].Results from FTIR and

Raman spectra indicate that Ag2O acts as a network modifier even at quantities by

converting three coordinated to four coordinated boron atoms. Optical basicity was

also evaluated which was affected by the silver oxide composition.

140

Spectroscopic technique such as ESR, optical absorption, Raman and IR were

synthesized to determine the structural groups and bonding parameters in alkali boro-

tellurite glasses doped with CuO [64]. From ESR spectra, the spin Hamiltonian

parameter values indicate that the ground state of Cu2+ is dx2-y2 and the site symmetry

around the Cu2+ ion is tetragonally distorted octahedral coordination. Bonding

parameters calculated from optical absorption and ESR data are found to change with

alkali oxide and TeO2 content. Both Raman and IR results show that glass network

consists of TeO3, TeO4, BO3 and BO4 group as basic structural groups. BO3→ BO4

transition is also observed, which correlates with the transition of TeO4 → TeO3 via

TeO3+1 .

Pure and copper doped multi-component lithium borosilicate glasses [65] have

been investigated as a function of Al2O3 concentration by a variety of spectroscopic

(optical absorption, IR, Raman and ESR) and dielectric properties (over a range of

frequency and temperature). The results of optical absorption and ESR spectral

studies have indicated that a part of copper ions do exist in Cu+ state in addition to

Cu2+ state especially in the samples containing low concentration of Al2O3. The IR

and Raman spectral studies have revealed that there is a decreasing degree of disorder

in the glass network with increase in the concentration of Al2O3.

Different concentrations of dysprosium doped strontium lithium bismuth

borate glasses we synthesized and characterized through Raman, absorption and

visible luminescence spectroscopy’s [66]. These Dy3+ doped glasses are studied for

their utility for white light emitting diodes. Coexistence of triangle BO3 and

tetrahedral BO4 units was evidenced by Raman spectroscopy. From the emission

spectra, a strong blue emission that corresponds to the transition, 4F9/2→6H15/2, was

141

observed and it also shows combination of blue, yellow and red emission bands for

these glasses. In addition to that, white light emission region have been observed from

these studies.

Maheshvaran et al [67] reported structural and optical behavior of the Er3+/

Yb3+ co-doped boro-tellurite glasses through FTIR, Raman, absorption, luminescence,

upconversion luminescence and lifetime measurements for green leaser applications.

Through the absorption spectra, bonding parameters, oscillator strengths and Judd-

Ofelt (JO) parameters were calculated and reported.

Different glasses in the system xCeO2-(1-x)B2O3 were prepared and

characterized by thermal expansion and infrared measurements [68]. IR results of

CeO2-rich glasses (40-60 mol%) confirm that CeO2 affects the glass network as both a

network modifier and a network former. The modifier part of CeO2 is consumed in

transformation of BO3 units to BO4 groups. The rest of CeO2 can participate in the

network in the form of CeO4 units.

5.5 AIM AND SCOPE OF PRESENT WORK

The aim of the present study is to obtain specific data regarding the local

structure of xLi2O-(30-x)Na2O-10WO3-60B2O3 (0 ≤ x ≤ 30) quaternary glass system

by means of Raman and infrared spectroscopy. The correlation between spectral

assignment and the physical properties of the glasses will be discussed. The presence

of mixed alkali oxides in the glass system increases the mixed alkali effect in the

present study.

142

5.6 RESULTS AND DISCUSSION

5.6.1 IR spectra

The IR absorption spectra of the present glasses were recorded in the range

300- 2000 cm-1. Figure 5.1 shows the normalized FTIR absorption spectra of xLi2O–

(30-x)Na2O–10WO3–60B2O3 glasses. The observed infrared spectra of these glasses

arise largely from the modified borate networks and are mainly active in the spectral

range 400-1600 cm-1; therefore the spectra are shown in 350-1800 cm-1 range for

better clarity.

Each spectrum was deconvelated by using 12 Gaussian functions considering

peak assignment as reported earlier [69-71]. An example of the fitting for 5Li2O–

25Na2O–10WO3–60B2O3 glass composition is shown in figure 5.2. The infrared spectra

of the present glasses show 11-12 absorption peaks. All the glass compositions show

absorption peak at 467 cm-1, 540 cm-1, 697 cm-1, 762 cm-1, 874 cm-1, 940 cm-1, 1020

cm-1, 1080 cm-1, 1228-1266 cm-1, 1346-1380 cm-1, 1438 cm-1 and 1637 cm-1. The peaks

are sharp, medium and broad. Broad bands are exhibited in the oxide spectra, most

probably due to the combination of high degeneracy of vibrational states, thermal

broadening of the lattice dispersion band and mechanical scattering from powder

samples. For the present glasses the IR band positions and area under the peak are

presented in Table 5.1.

According to the Krogh Moe’s the structure of the boron oxide glass consists of

a random network of planer BO3 triangles with a certain fraction of six membered

(boroxol) rings [7]. X-ray and neutron diffraction data suggests that glass structure

consists of a random network of BO3 triangles without boroxol rings. The vibrational

modes of the borate network are active mainly in three regions: the first region lies

between 600 and 800 cm-1 and is due to bending vibration

143

144

The vibrational modes of the borate network are active mainly in three

regions: the first region lies between 600 and 800 cm-1 and is due to bending vibration

of various borate segments, the second region lies between 800 and 1200 cm-1 and is

due to stretching vibrations of tetrahedral BO4 units and third region lies between

1200 and 1600 cm-1 and is due to stretching vibrations of B-O in BO3 triangles [72-

74]. Alkali oxides like Li2O and Na2O are well known glass modifiers and may enter

the glass network by transforming sp2 planar BO3 units into most stable sp3

tetrahedral BO4 units and may also create non–bridging oxygens. Both BO3 and BO4

units co-exist in these glasses, which is evident from Figure 5.1.

The broad IR bands as shown in figure 5.2 are the overlapping of some

individual bands with each other. Each individual band has its characteristic

parameter such as its center which is related to some type of vibration of a specific

structural group. A weak IR band around 467 cm-1 is assigned to the vibrations of Li

cations through glass network. This IR band increases and then decreases in intensity

as Li2O content is increasing. Padmaja and Kistaiah [75] observed the vibrations of

Li cations in mixed alkali zinc borate glasses doped with transition metal ions at

around 460 cm-1 in the IR spectra. Moustafa et al [76] observed weak IR band around

438 cm-1 in Li2O-B2O3-Bi2O3 glass system which was attributed to Li-O-Li bonds.

The present IR spectra showed non-existence of band at 806 cm-1, which reveals the

absence of boroxol rings in glasses and hence it consists of only BO3 and BO4 groups

[77, 78]. In Li2O-B2O3-Bi2O3 glasses the peak at around 700 cm-1 is assigned to

pentaborate units [79]. In alkali boro-tungstate glasses, the IR band around 700 cm-1

stands for B-O-B bond bending vibrations of bridging oxygen atoms [80,81]. In the

present study, the IR peak around 697 cm-1 is assigned to bending vibrations of

145

pentaborate groups, which are composed of BO4 and BO3 units in the ratio 1:4. The

intensity of this band increases and then decreases with Li2O content.

Since WO3 is a conditional glass former, with the substitution of WO3 with

alkali oxides in borate glass network the intensity of vibrational band due to the BO3

groups is observed to increase at the expense of BO4 structural units [82]. From

XANES and FTIR studies in TeO2-WO3 glasses it was observed that W6+ prefers six-

coordination and exhibits an absorption band at 930 cm-1 [83]. In the present IR

spectra the peak at around 940 cm-1 is assigned to the stretching vibrations of B-O

linkages in BO4 tetrahedra overlapping with the stretching vibrations of WO6 units.

Boudlich et al [70] reported a mixture of WO4 and WO6 units at 880 cm-1 and

at 900-950 cm-1 respectively, in alkaline tungsten phosphate glasses. In the present

study the IR peak around 874 cm-1 is assigned to starching vibration of tri-, tetra- and

penta- borate groups and also due to the starching vibration of non-bridging oxygens

of BO4 groups overlapping with the stretching vibrations of WO4 units. This peak was

observed around 866 cm-1 and at around 850 cm-1 in single alkali boro-tungstate

glasses [8,84].

A broad band around 1020 cm-1 is assigned to stretching vibrations of B-O

bonds in BO4 units from tri, tetra and penta borate groups. The weak peak at about

762 cm-1 can be attributed to B-O-B bending vibrations of BO3 and BO4 groups with

W-O-W vibrations in the borate network [85,86]. This indicates that tungsten enter

the glass structure. The IR band around 1080 cm-1 is assigned to penta borate groups

[87]. The peak lying in 1346-1380 cm-1 is attributed to asymmetric stretching

vibrations of the B-O of trigonal (BO3)3- units in meta-, pyro- and ortho-borate units

[85]. The band around 1438 cm-1 is assigned to antisymmetrical stretching vibrations

146

with three non-bridging oxygens of B-O-B linkages [88-92]. The weak band

observed around 1637 cm-1 indicates a change from BO3 triangles to BO4

tetrahedra, and this peak may also be

Table 5.1 Infrared absorption band positions of the xLi2O-(30-x)Na2O-10WO3-60B2O3

glass system

Glass IR band positions in wave number (cm-1)

x=0 465 539 697 762 874 934 1020 1074 1266 1377 1407 1637

X=5 464 532 697 764 874 938 1015 1050 1266 1363 1460 1629

X=10 456 533 694 761 868 945 1010 1075 1228 1346 1433 1638

X=15 467 535 695 762 877 944 1015 1081 1234 1350 1438

1636

X=20

467 539 697 764 873 940 1017 1080 1261 1371 1442 1632

X=25

467 537 698 764 876 940 1023 1075 1254 1360 1448 1632

X=30 444 540 695 759 879 946 1021 1076 1246 1381 1452 1641

assigned to OH bending mode of vibrations [93,94]. The IR band in the range 1228-

1266 cm-1 is assigned to B-O stretching vibrations of (BO3)3- unit in metaborate

chains and orthoborates and these groups contain large number of non-bridging

oxygens (NBO’s) [84]. This suggests the conversation of the BO4 tetrahydral to the

non-bridging oxygen containing BO3 trangles. The peak at around 540 cm-1 can be

attributed to the borate deformation modes such as the in-plane bending of boron-

oxygen triangles [95]. Figure 5.3(a) and figure 5.3(b) shows the compositional

dependence of various peak position of the IR bands in the present study. The above

figures depicts a non linear variation in peak positions for IR band centered around

697 cm-1, 874 cm-1, 1020 cm-1and 1346 cm-1 exhibiting mixed alkali in present

glasses. The assignments of IR bands are given in Table 5.2.

147

5.2 Infrared band assignments of the present glasses.

Wavenumber

( cm-1 )

IR band assignments

~ 467 Li cation vibrations

~ 540 Borate deformation mode such as the in-plane bending of B–O triangles

~ 697 Bending vibrations of pentaborate groups, which are composed of BO4

and BO3 units in the ratio 1:4

~ 762 B–O–B bending vibrations of BO3 and BO4 groups with W–O–W

vibrations in the borate network

~ 874 Stretching vibration of tri-, tetra- and pentaborate groups also due to the

stretching vibration of non-bridging oxygen’s of BO4 groups overlapping

with the stretching vibration of WO4 units

~ 940 Stretching vibration of B–O linkages in BO4 groups tetrahedral

overlapping with the stretching vibration of WO6 units

~1020 Stretching vibration of B–O bands in BO4 units from tri, tetra and penta

borate groups

~1080 Pent borate groups

~1266 B-O stretching vibration of (BO3)3- units in metaborate chains and ortho

borates

~ 1346 Asymmetric stretching vibration of the B-O of triangle (BO3)3- units in

meta-, pyro- and ortho-borate units

~ 1438 Anti symmetrical starching vibrations with three non-bridging oxygen’s

of B-O-B linkages

~1637 OH bending mode of vibration and change from BO3 triangle to BO4

tetrahedra

148

To quantify the inter-alkali variation effect in the relative population of

tetrahedral and triangular borate units we have calculated the fraction of four-

coordinated boron atoms, N4 and three coordinated boron atoms containing NBOs, N3

which were estimated as follows [96]

N4 = [A4] / {[A3] + [A4]} and N3 = 1- N4 (5.3)

where A3 and A4 denotes the areas of BO3 units (the areas of component IR bands

from 1200- 1600 cm-1) and BO4 units (the areas of component bands from 800- 1200

cm-1), respectively. The amount of four coordinate boron atoms, N4 and three

149

coordinate boron atoms is plotted as a function of inter alkali variation in figure 5.4. It

is clear from the above figure that the non-bridging oxygen containing BO3 units, N3

varies non linearly. There is some sort of ordering that occurs which leads to a

lessening of NBOs at RLi=0.5.

5.6.2 Raman spectra

Raman spectroscopy is one of the techniques used to investigate the structure

of a glass. The room temperature Raman spectra of the present glass system is shown

in figure 5.5. There are three regions clearly visible in the Raman spectra : (i) 250 –

500 cm-1, (ii) 500 – 1100 cm-1 and (iii) 1250 – 2000 cm-1. In all the present glasses

150

studied the total alkali content is 30 mol%. At this concentration of the alkali content

in borate containing glasses , the boroxol rings get converted mostly into pentaborate

groups. This is observed clearly by the strong presence of peaks around 787 and 684

cm-1 resembling the localized breathing motions of oxygen atoms in the boroxol ring.

Each Raman spectrum was deconvoluted by using 7-8 Gaussian functions to identify

the exact position of peak and their intensity variation. Deconvoluted Raman spectra of

xLi2O–(30-x)Na2O–10WO3–60B2O3 glass system is shown in figure 5.6. All the glass

compositions show Raman peak at around 333 cm-1, 554 cm-1,650cm-1, 684 cm-1, 787 cm-1,

825cm-1, 873 cm-1, 957 cm-1 and 1464 cm-1. The Raman band positions of all the glasses

under study are given Table 5.3.

The vibrational Raman bands at 329 -340 cm-1, 873-904 cm-1 and 951- 960 cm-1

belonging to tungstate groups undergo complex changes. In the Raman spectra of all the

studied glasses, there is a strong peak observed at ~957 cm-1 which is assigned to W–O–

stretching vibrations in WO4 tetrahedral. The peaks around 329-340 cm-1 are due to the

bending vibrations of W–O–W in the WO6 units [97]. In TeO2-WO3 glasses this peak was

observed around 340-360 cm-1 [98]. The Raman band around 873-904 cm-1 is assigned to

stretching vibrations of W–O–W in the WO4 or WO6 units.

In the Raman spectra of all the glassy specimens, there is a peak observed

around 774-793 cm-1 which is characteristic of a six membered ring with one or

two BO4 tetrahedra. Earlier, Brill [99] assigned this peak to the formation of six

membered rings containing one BO4 tetrahedron, and the shift of this peak towards

lower frequency has been assigned to six memebered rings with two BO4 tetrahedra.

The six membered rings with one BO4 tetrahedron can be in triborate, tetraborate or

pentaborate forms, and rings with two BO4 tetrahedra can be in diborate, di-triborate

or di-pentaborate forms. In alkali borate glasses [100] and also in PbO-B2O3 glasses

151

this peak was observed at 806 cm-1 and 760-780 cm-1, respectively [101]. In the

studied glasses, the presence of Raman band in the range at 675-692 cm-1 has been

attributed to the pentaborate groups in the borate glasses. Similar results were

observed in mixed alkali zinc borate glasses [62].

.

152

Table 5.3 Observed Raman band positions of the present glass system.

Glass Raman band positions in wave number (cm-1)

30Na2O-10WO3-60B2O3 339 546 652 677 788 812 874 934 1407

5Li2O-25Na2O-10WO3-60B2O3 334 564 664 675 764 808 882 938 1460

10Li2O-20Na2O-10WO3-60B2O3 339 555 649 691 776 810 874 946 1433

15Li2O-15Na2O-10WO3-60B2O3 340 554 638 692 779 815 867 944 1438

20Li2O-10Na2O-10WO3-60B2O3 329 559 649 698 792 820 873 940 1442

25Li2O-5Na2O-10WO3-60B2O3 337 564 644 684 764 817 886 940 1448

30Li2O-10WO3-60B2O3 333 561 650 686 776 825 879 942 1452

Table 5.4 Observed Raman band positions in xLi2O-(30-x)Na2O-10WO3-60B2O3 glass system.

Band positions

( cm-1 )

Raman band assignments

~329 Bending vibrations of W-O-W in the WO6 units

~ 546 In plane bending mode of BO33- units

~ 675 Denotes the existence of BO2O23- units

~ 697 Stretching of B-O- bonds attached to large number of borate groups

~ 774 Ring breathing vibration of six membered ring contains both BO3

triangles and BO4 tetrahedral

~ 836 Symmetric stretching Vibration of planar orthoborate units (BO3)3-

~ 873 Starching vibrations of W-O-W in the WO4 or WO6 units

~ 951 W-O- stretching vibrations in WO4 tetrahedra

~1464 Stretching of B–O- bonds attached to large number of borate groups

153

The Raman bands in the high frequency range 1429-1547 cm-1 has been

assigned to stretching of B-O- bonds attached to large number of borate groups by

Kamitsos [9]. In K2O-B2O3 glasses the Raman band around 1490 cm-1 is attributed to

BO2O- triangles linked to other borate triangular units [62]. In the present study 1464

cm-1 Raman band was assigned to stretching of B-O- bonds attached to large number

of borate groups. The Raman band in the range 546-564 cm-1 is assigned to in plane-

bending mode of BO33- units. In R2O-B2O3 (R=Li, Na and K) glasses the Raman

band centered at 548 cm-1 is assigned to in plane-bending mode of BO33- units [100].

In all the glasses studied the total alkali content is of 30 mol%. At this concentration

of alkali content in borate glasses, the boroxol rings get converted mostly into

pentaborate groups. This is observed clearly by the strong presence of a weak peak ~

650 cm-1 resembling the localized breathing motions of oxygen atoms in the boroxol

ring [75]. The Raman band around 825 cm-1 is assigned to W-O single bond

stretching vibrations within W–O–W bonded units [102]. The assignments of

Raman bands are given in Table 5.4.

Raman spectroscopic studies of alkali borate glasses for different

concentrations of R2O reveal the possibility of two chemical processes by which the

alkali ion can be dispersed in the glasses. The first process, operative at lower

concentrations of R2O, leads to the formation of boron in fourfold coordination, i.e.

BO4– units, with the positive alkali ion (R+) adjacent to the negative BO4– unit to

provide local charge neutrality. The second process is the formation of a non-bridging

oxygen (O– ) adjacent to the positive alkali ion.

154

5.7 CONCLUSIONS

Mixed alkali tungsten borate glasses in the form of xLi2O–(30-x)Na2O–

10WO3–60B2O3 (0 ≤ x ≤ 30) were prepared, and their structural properties have been

studied. The following conclusions were made:

(i) The infrared studies indicate the presence of BO3, BO4, WO3, WO6 and Li units in

the structure of the studied glasses. The intensities and their peak position were

affected by the alkali concentrations in each glass. The peak positions of few IR

bands showed non-linear variation with alkali content manifesting mixed alkali

effect.

(ii) The Raman spectra of the investigated glasses exhibits several bands which are

attributed to BO3 , BO4 tetrahedra and pentaborate groups linked to BO4

tetrahedra. Raman spectra confirms the IR results regarding the presence of

tungsten ions mainly as WO6 groups

(iii) The amount of four coordinate boron atoms, N4 and the non-bridging oxygen

containing BO3 units, N3 varies non linearly as a function of inter alkali variation.

155

5.8 References

[1] R. Shuker, R.W. Gammon, Phys. Rev. Lett. 25 (1970) 222.

[2] P. H. Gaskell, Trans. Faraday Soc. 62 (1966) 1493.

[3] S. Brawer, Phys. Rev. B11 (1975) 3173.

[4] E. I. Kamitsos, M. A. Karkassides, G. D. Chryssikos, Phys. Chem. Glasses, 30

(1989) 229.

[5] E. I. Kamitsos, A. P. Patsis, G. D. Chryssikos, Phys. Chem. Glasses 32 (1989)

219.

[6] W. L. Konijnendijk, Phys. Chem. Glasses 17(6) (1976) 205.

[7] J. Krogh-Moe, J. Phys. Chem. Glasses 6(2) 1965 46.

[8] E. I. Kamitsos, M. A. Karkassides, G. D. Chryssikos, J. Phys. Chem. 90 (1986)

4528.

[9] E. Kamitsos, M. Karakassides, G. Chryssikos, J. Phys. Chem. 91 (1987) 1073.

[10] E. I. Kamitsos, M. A. Karakassides, G. D. Chryssikos, Phys. Chem. Glasses 28

(1987) 203.

[11] E. I. Kamitsos, M. A. Karakassides, Phys. Chem. Glasses 30 (1989) 19.

[12] G. D. Chryssikos, E. I. Kamitsos, W. M. Risen Jr., J. Non-Cryst. Solids 93

(1987) 155.

[13] Y. Tang, Z. Jiang, X. Song, J. Non-Cryst. Solids 112 (1989) 131.

[14] W. L. Konijnendijk, H, Verweij, J. Am. Ceram. Soc. 59 (1976) 459.

[15] B. N. Meera, J. Ramakrishna, J. Non-Cryst. Solids 159 (1993) 1.

[16] S. Bale, M. Purnima, Ch. Srinivasu, Syed Rahman, J. Alloys compds. 457

(2008) 545.

[17] S. Bale, Syed Rahman, Opt. Mater. 31 (2008) 333.

[18] S. Bale, N. S. Rao, Syed Rahman, Solid Sate Sciences 10 (2008) 326.

156

[19] S. Bale, Syed Rahman, J. Non-Cryst. Solids 355 (2009) 2127.

[20] S. Bale, Syed Rahman, Physica B: Condens. matter 418 (2013) 52.

[21] S. Bale, Syed Rahman, Modren Physics Letters B 23 (22) (2009) 2665.

[22] R. V. Kumar, S. Bale, Ch. Srinivasu, M. A. Samee, K. S. Kumar, Syed Rahman,

J. Alloys compd. 490 (2010) 1.

[23] Ch. Srinivasu, V. Sathe, A. M. Awasti, Syed Rahman, J. Non-Cryst. Solids 357

(2011) 1051.

[24] T. G. V. M. Rao, A. R. Kumar, N. Veeraiah, M. R. Reddy, J. Phys. Chem.

Solids 74 (2013) 410.

[25] C. Ivascu, A. T. Gabor, O. Cozar, L. Daraban, I. Ardelean, J. Mol. Struct. 993

(2011) 249.

[26] C. Ivascu, I. B. Cozar, L. Daraban, G. Damian, J. Non-Crysts. Solids 359

(2013) 60.

[27] B. Karthikeyan, S. Mohan, M. L. Baesso, Physica B 337 (2003) 249.

[28] B. Karthikeyan, S. Mohan Physica B 334 (2003) 298.

[29] L. Koudelka, J. Subcik, P. Mosner, L. Montagne, L. Delevoye, J. Non-Cryst.

Solids 353 (2007) 1828.

[30] L. Koudelka, P. Mosner, L. Montagne, M. Zeyer-Dusterer, C. Jager, J. Phys.

Chem. Solids 68 (2007) 173.

[31] I. Ardelean, S. Cora, R. Ciceo-laucacel, Modren Phys. Lett. B 18 (2004) 803.

[32] S. Hazra, S. Mandal, A. Ghosh, Physical Review B 56 (1997) 8021.

[33] S. Hazra, A. Ghosh, Physical Review B 51 (1995) 851.

[34] Y. Cheng, H. Xiao, W. Guo, Thermochimica Acta 444 (2006) 173.

[35] G. Upender, J. Chinna Babu, V. Chandr Mouli, Spectrochemica Acta Part A:

Molecular and Biomolecular Spectroscopy 89 (2012) 39.

157

[36] K. El-Egili, Physica B 325 (2003) 340.

[37] G. K. Kumari, Ch. Rama Krishna, SK. Muntaz Begum, V. P. Manjari, P. N.

Murthy, R. V. S. S. N. Ravikumar, Spectrochemica Acta Part A: Molecular and

Biomolecular Spectroscopy 101 (2013) 140.

[38] W. M. Risen Jr., G. B. Rouse Jr. and P. J Miller, J. Non-Cryst. Solids 28 (1978)

193.

[39] E. I. Kamitsos, W. M. Risen Jr., J. Non-Cryst. Solids 65 (1984) 333.

[40] S. Z. A. Ahamed, C. M. Reddy, B. D. P. Raju, Opt. Mater. 35 (2013) 1385.

[41] S. Z. A. Ahamed, C. M. Reddy, B. D. P. Raju, Spectrochemica Acta Part A:

Molecular and Biomolecular Spectroscopy 103 (2013) 246.

[42] N. S. Rao, S. Bale, M. Purnima, K. Siva Kumar, Syed Rahman, J. Phys. Chem.

Solids 68 (2007) 1354.

[43] N. S. Rao, S. Bale, M. Purnima, K. Siva Kumar, Syed Rahman, Bull. Mater.

Sci. 29 (2006) 365.

[44] W. Soppe, J. Kleerebezem, H. W. den Hartog, J. Non-Cryst. Solids 93 (1987)

142.

[45] E. I. Kamitsos, G. D. Chryssikos, A. P. Patsis, Solid State Ionics 67 (1992) 331.

[46] G. Padmaja, P. Kistaiah, Soild State Sciences 12 (2010) 2015.

[47] G. Padmaja, P. Kistaiah, Vib. Spectrosc. 62 (2012) 23.

[48] U. Selvaraj, K. J. Rao, Spectrochemica Acta Part A: Mol. Spectroscopy 40

(1984) 1081.

[49] A. M. Efimov, J. Non-Cryst. Solids 253 (1999) 95.

[50] Sharada and Suresh babu, Physica B: condens. Matter 407 (2012) 3945.

[51] E. Mansour, J. Non-Cryst. Solids 358 (2012) 454.

[52] E. Mansour, J. Non-Cryst. Solids 357 (2011) 1364.

158

[53] T. R. Rao, Ch. Venkat Reddy, Ch. R. Krishna, U. S. U. Thampy, P. S. Rao,

R. V. S. S. N. Ravi Kumar J. Non-Cryst. Solids 357 (2011) 3373.

[54] T. R. Rao, Ch. R. Krishna, Ch. Venkat Reddy, U. S. U. Thampy, Y. P. Reddy,

P. S. Rao, R. V. S. S. N. Ravi Kumar Spctrochimica Acta Part A: Molecular and

Biomolecular Spectroscopy 79 (2011) 1116.

[55] V. Naresh, S. Buddhudu Ceramics International, 38 (2012) 2325.

[56] A. H. Verhoef, H. W. den Hartog, J. Non-Cryst. Solids 182 (1995) 221.

[57] I. Kashif, A. Ashia, Soliman, M. Sakr Elham, A. Ratep, Spctrochimica Acta Part

A: Molecular and Biomolecular Spectroscopy 113 (2013) 15.

[58] H. Doweidar, G. El-Damrawi, E. Mansour, R. E. Fetouh, J. Non-Cryst. Solids

358 (2012) 941.

[59] D Di Martino, L. F Santos, A. C Marques, R. M Almeida J. Non-Cryst. Solids

293 (2001) 941.

[60] M. S. Gaafar, H. A. Afifi, M. M. Mekaway, Physica B: Cond. Mat. 404 (2009)

1668.

[61] Gajanan V. Honnavar, S. N. Prabhava, K. P. Ramesh J. Non-Cryst. Solids 370

(2013) 6.

[62] R. Akagi, N. Ohtori, N. Umesaki, J. Non-Cryst. Solids 293 (2001) 471.

[63] J. Coelho, C. Freire, N. S. Hussain, Spctrochimica Acta Part A: Molecular and

Biomolecular Spectroscopy 86 (2012) 392.

[64] S. Suresh, P. G. Pavani, V. C. Mouli Mater. Res. Bull. 47 (2012) 724.

[65] P. R. Babu, R. Vijay, P. S. Rao, P. Suresh, N. Veeraiah, D. K. Rao, J. Non-

Cryst. Solids 370 (2013) 21.

[66] D. Rajesh, Y. C. Ratnakaram, M. Seshadri, A. Balakrishna, T. S. Krishna

J. Lumin. 132 (2012) 841.

159

[67] K. Maheshvaran, S. A. Kumar, V. Sudarsan, V. Natarajan, K. Marimuthu,

J. Alloys Compd. 561 (2013) 142.

[68] G. El-Damrawi, K. El-Egili, Physica B: Cond. Mat. 299 (2001) 180.

[69] L. Bih, L. Abbas, S. Mohdachi, A. Nadiri, J. Mol. Struct. 891 (2008) 173.

[70] D. Boudlich, L. Bih, M. El Hassane Archidi, M. Haddad, A. Yacoubi, A.

Nadiri, B. Elouadi, J. Am. Ceram. Soc. 85 (3) (2002) 623.

[71] M. S. Gaffar, S. Y. Marzouk, Physica B 388 (2007) 294.

[72] M. M. El-Desoky, H. Farouk, A. M. Abdalla, M. Y. Hassaan, J. Mater. Sci.

Mater. Elect. 9 (1998) 77.

[73] K. El-Egili, A. H. Oraby, J. Phys.: Condens. Matter 8 (1996) 8959.

[74] S. G. Motke, S. P. Yawale, S. S. Yawale, Bull. Mater. Sci. 25(1) (2002) 75.

[75] G. Padmaja, P. Kishtaiah, J. Phys. Chem. A 113 (2009) 2397.

[76] El Sayed Moustafa, Y. B. Saddek, E. R. Shaaban, J. Phys. Chem. Solids 69

(2008) 2281.

[77] F. L. Galeener, G. Lucovsky, J. C. Mikkelsen Jr., Phys. Rev. B 22 (1980) 3983.

[78] M. A. Kaneshisa, R. J. Elliot, Mater. Sci. Eng. B 3 (1989) 163.

[79] Y. B. Saddeek, E. R. Shaaban, E. S. Moustafa, H. M. Moustafa, Physica B 403

(2008) 2399.

[80] Manal Abdel-Baki, Fouad El-Diasty, J. Solid State Chem 184 (2011) 2762.

[81] A. Sheoran, A. Agarwal, S. Sanghi, V. P. Seth, S. K. Gupta, M. Arora, Physica

B 406 (2011) 4505.

[82] P. Charton, L. Gengembre, P. Armand, J. Solid State Chem. 168 (2002) 175.

[83] D. M. Martin, M. A. Villegas, J. Gonzalo, J. M. F. Navarro, J. Eur. Ceram. Soc.

29 (2009) 2903.

160

[84] M. S. Gaffar, Y. B. Saddeek, L. Abd El-Latif, J. Phys. Chem. Solids 70 (2009)

173.

[85] I. Shaltout, Y. Tang, R. Braunstein, E. E. Shaisha, J. Phys. Chem. Solids 57

(1996) 1223.

[86] R. C. Lucacel, I. Ardelean, J. Non-Cryst. Solids 353 (2007) 2020.

[87] M. Milanova, R. Iardanova, K. L. Kostov, J. Non-Cryst. Solids 355(6) (2009)

379.

[88] S. Rada, M. Culea, M. Neumann, E. Culea, Chem. Phys. Lett. 460 (2008) 196.

[89] S. Rada, P. Pascta, M. Culea, V. Maties, M. Rada, M. Barlea, E. Culea, J. Mol.

Struct. 924 (2009) 89.

[90] S. Rada, M. Culea, E. Culea, J. Non-Cryst. Solids 354 (2008) 5491.

[91] S. Rada, P. Pascuta, M. Bosca, M. Culea, L. Pop, E. Culea, Vib. Spectrosc.

48(2) (2008) 255.

[92] G. Yahya, Turk. J. Phys. 27 (2003) 255.

[93] B. V. R. Chowdari, Z. Rong, Solid State Ionics 90 (1996) 151.

[94] E. I. Kamitsos, G. D. Chryssikos, Solid State Ionics 105 (1998) 75.

[95] B. Bendow, P. K. Banerjee, M. G. Drexhage, J. Lucas, J. Am. Ceram. Soc. 65

(1985) C92.

[96] S. Prabakar, K. J. Rao, C. N. R. Rao, Proc. R. Soc. London, Ser. A, 429 (1990)

1.

[97] B. V. R. Chowdari, P. Pramoda Kumari, Solid State Ionics 113-115 (1998) 665.

[98] G. Upender , V. G. Sathe , V. C. Mouli, Physica B 405 (2010) 1269.

[99] T.W. Brill, Philips Res. Rep. Suppl. no 2 (1976) 117.

[100] B. P. Dwivedi, B. N. Khanna, J. Phys. Chem. Solids 56 (1995) 39.

161

[101] B. N. Meera, A. K. Sood, N. Chandrabhas, J. Ramakrishna, J. Non-Cryst.

Solids 126 (1990) 224.

[102] Carla C. de Araujo, Wenzel Strojek, Long Zhang, Hellmut Eckert, Gael

Poirier, Sidney J. L. Ribeiro, Younes Messaddeq, J. Mater. Chem. 16 (2006)

3277.