Sie sind auf Seite 1von 10

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

Spectroscopic Investigation of The Optical Properties of Rare Earth


Doped Phosphate and Tellurite Glasses
M. R. Sahar, M. S. Rohani, R. Ariffin And S. K. Ghoshal
Advanced Optical Material Research Group, Department of Physics, Faculty of Science,
Universiti Teknologi Malaysia, 81310 UTM Skudai, Johor
Email: mrahim057@gmail.com

Abstract-Improving the up-conversion efficiency is the key issue in rare earth doped glasses. The quantum efficiency, optical
gap, radiative transition rate and lifetimes of excited states are greatly influenced by the optical properties of the host
material, ligand field, multi-phonon relaxation processes, impurities, temperature and concentration of rare earth ions. This
presentation gives a panoramic view of the state of art of investigating experimentally the optical properties of rare earth
doped phosphate and tellurite glasses with different compositions and preparation techniques that our group has been carried
out in the recent past. We have prepared these glasses by using two different methods melt quenching and sol-gel, however in
the present paper we only report on the former method. Series of erbium-doped magnesium-phosphate glasses based on
(P2O5)0.5(MgO)0.5-x(Er2O3 / Nd2O3)x, ytterbium-doped sodium-tellurite glasses based on (80 x)TeO2 20Na2O-(x)Yb2O3
(x=0.0-2.0 mol%), samarium-doped glasses having composition of (P2O5)50-x(MgO) 50-(Sm2O3) x (0x4 mol%), and (50x)P2O5-50MgO-xSm2O3 (x= 0.0-3.0 mol%) have been studies. In addition, ytterbium-doped sodium-tellurite glasses having
composition of (80 x)TeO220Na2O-(x)Yb2O3 (x=0.0-2.0 mol%) is also presented. The amorphous nature of all these samples
is confirmed using the X-ray diffraction technique. The optical properties of the glass have been measured employing Infrared,
FTIR and UV-Visible spectroscopy. The vibrational frequencies are attributed predominately due to the OH- band, P=O, P-OP, and P-O- stretching vibrations respectively. Furthermore, it is found that the absorption cut-off wavelength very much
depending on the MgO contents and the integration mode area of absorption band is strongly affected by the phosphate
contents. Our systematic spectral analysis revealed that the substitution of P2O5 by a small amount of Sm2O3 has negligible
effect on the absorption band and hence on the glass structure. It is further observed that MgO does not affect the P-O bond
characters. The optical energy gap (Eg ) and the Urbach energy (E) has been estimated from the absorption edge studies. It is
found that Eg depend on the concentration of the non bridging oxygen in the glass network. Meanwhile, E is found to be
depending on the Yb2O3 concentration. Interestingly, samarium doped magnesium-phosphate glasses have been found to
change from colorless to light yellow on increasing the Sm2O3 content. The transmission spectra of ytterbium-doped sodiumtellurite glasses revealed that the predominant absorption peaks are due to the vibration of Te-O-Te and Yb3+ ions. The main
results on vibrational frequencies, optical gap, Urbach energy and absorption edge has been found to be quantitatively
consistent with other observations. Our detailed systematic spectroscopic studies provide useful information for further
development of up-conversion lasers.
Keywords: Phosphate glass, Tellurite glass, Melt-quenching, Sol-gel, Optical absorption, UV-VIS Spectroscopy, X-ray
diffraction, IR-spectroscopy, Optical properties, Cut-off wavelength, Urbach energy.
I.

INTRODUCTION
In the new millennium, there has been a renaissance in
the study of rare-earth doped glass materials for photonic
applications, e.g. phosphors, display monitors, X-ray
imaging, scintillators, lasers, up-conversion and amplifiers
for fiber-optic communications [110]. There are many
candidates in the family viz., borate, chalcogenide, fluoride,
germinate, oxynitride, silicate, phosphate, sulphide,
zirconate, and tellurite glasses. Rare-earth ions, especially
erbium, ytterbium and samarium, have played an important
role in the development of broadband fiber amplifiers in
optical communication technology during the past few
decades [5-7]. Phosphate glasses has attracted much
attention in recent years due to their unique high thermal
expansion, low melting temperature, high transmission in
the UV region and radiative properties [1, 2]. The main
advantage of a phosphate glass over other oxide glasses (e.g.
silicate and borate) is its ability to accommodate high
concentration of transition metal ions and remain
amorphous. In addition, phosphate glasses enjoy a range of
compositional and structural possibilities (ultra, meta, pyro,
and ortho) that facilitate tailoring chemical and physical

properties of interest for specific technological applications


[3, 4]. It is also an excellent material as host material due to
their good chemical durability, ion exchange ability, high
gain coefficient, wide bandwidth capability and low up
conversion emission.
Compared to silicates phosphate sol-gels studies
are very few in the literature. The solution chemistry of
phosphate materials is very different from silicate, which,
consequently, continues to make the phosphate systems
much more complex than the silicates [5]. Nowadays,
ytterbium (Yb3+) ion is regarded as the main dopant for the
application of high-power diode-pumped laser systems.
Since there are only two manifolds in the Yb3+ energy level
scheme namely the 2F7/2 ground state and 2F5/2 excitation
state, it is commonly believed that concentration quenching
and multi-phonon relaxation should not affect the excitation
wavelength [6]. However, metaphosphate glasses containing
rare-earth ions have potentially important applications in
optical communications and laser technologies. Glasses of
the R(PO3)3 formulas (where R represents one of the
lanthanide ions) have been reported to exhibit the largest
magnetic contributions to the low temperature specific heats

ISSN 0853-0823

10

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

known in oxide glasses. Even though there have been


studies of phosphate glasses doped with the lanthanide ions,
there have been limited studies of phosphate glasses codoped with the lanthanide ions [4]. They find a place in the
phosphor and luminescence materials applications, cathode
ray tube phosphors and scintillate phosphors, because of
their unique spectroscopic properties [1, 2, 4].
Recently, phosphate and tellurite glasses have been
exploited for various applications in optoelectronic devices,
fibre lasers, optical amplifiers, sensing and laser
technologies [1-2]. These glasses also exhibit reasonable
mechanical properties although may show some anomalous
characteristic due to the excessive amount of the modifier
[4]. The study on the effect of rare earth doped phosphate
glasses on the optical properties has also been done by many
workers especially in the ternary systems [5-6]. Our study
provides essential information on the band structure and the
energy gap in the non-crystalline material. While the
absorption in the lower energy part gives information about
the atomic vibrations, the higher energy part carries
information on the electronic state of the atom.
Tellurite glasses are considered as one of the best
hosts for doping with rare earth elements. It is a good
candidate for practical laser applications because of low
crystallization rate, excellent transparency in a wide spectral
range (318m), good mechanical stability and chemical
durability [7-9]. The optical properties of tellurite glasses
doped with rare earth have been investigated by several
groups [9-11]. Tellurite glasses doped with Nd3+ and Yb3+
are important for possible applications as luminescent solar
concentrator [11]. TeO2-based glasses were most studied in
the last 10 years, considering the scientific and technological
interest due to their high refractive indices, low melting
temperatures, high dielectric constants and good infrared
transmissions [12].
Topically important rare earth doped tellurite and
phosphate glasses in general are intensively studied, as they
are promising for widespread potential applications due to
excellent third-order nonlinear optical performance.
Recently, light energy up-conversion property of such
nonlinear optical glasses received special attention because
of their prospective use in biological labeling and solar near
infrared concentration for photovoltaic exploitation [3-9].
Meanwhile, erbium doped tellurite glass containing quantum
dots or metal nanoparticles (NPs) stimulate intense interest
in functionalizing tellurite glass by NPs. In that respect, NPs
dispersed up-converting glasses seem to be the ideal
candidates in terms of both efficiency and large area
coverage provided the absorption cross-section be enhanced.
To achieve enhanced optical characteristics in these glasses,
the concentration of rare-earth ions should be low enough to
avoid the quenching effect. Use of two or more rare-earth
ions together and energy transfer between them or doping
metallic NPs with rare-earth ions etc. has been found a
successful way to enhance luminescence. Therefore, glasses
containing metallic NPs doped with low concentration of
RE ions are of particular research interest to us. The
characterizations of the nonlinear optical and thermo-optical
properties of these glasses are very important for the
optimization and the nanophotonic applications. In spite of
some experiments on these glasses, the fundamental
understanding on the unusual nonlinear optical properties is

still lacking [7-15]. Interestingly, the quantum effect due to


metal NPs around the luminescent ion that possibly
enhances the nonlinear optical performance requires further
systematic experimental theoretical investigation. We will
focus on issues those are relevant to the fabrication of
tellurite
glass-based
nanophotonic
devices
and
photovoltaics. So far, there is no systematic theory or model
and not too many experiments exists to explain the influence
of embedded NPs in the erbium doped tellurite glasses.
Lanthanide doped crystal and glasses differ in their
physical properties as well, which influences the
manufacturing. Glasses can be produced much more
inexpensively and offer more flexibility in the size and
shape. They can be drawn into fibers that microns in
diameter and meters in length, or made into bulk rods that
are centimeters in diameter and meters long. Glasses also
have larger flexibility in their physical properties through
selection of the base material [13]. Analytical method using
X-rays are some of the most powerful techniques for
materials characterization. When X-rays encounter matter, a
variety of processes may take place and each of these
processes can be utilized to study particular properties of
material [14]. Infrared absorption spectra of glasses can give
valuable information about atomic configurations in glasses
even though quantitative analysis is rarely possible.
Experimental as well as theoretical investigations of
vibrational spectra for vitreous solids have been undertaken
[15]. In spite of vast literature, the detailed mechanism
behind the origin of linear and nonlinear optical properties is
far from being understood. In view of the topical importance
of these glasses we are tempted to investigate the optical
properties of samarium doped magnesium-phosphate
glasses, phosphate glass co-doped with the lanthanide ions
by using Infrared and UV-Visible spectroscopy following
two different sample preparation roots namely, melt
quenching and sol-gel respectively.
II. METHODOLOGY
A. Experimental
A.1 Glass Preparation
The glass samples have been prepared from starting
material constituents of P2O5, MgO, Er2O3 and Nd2O3 by
melt-quenching technique with constants phosphate content
at 50 mol% and by changing the MgO content as the Er2O3
and Nd2O3 change. Four samples co-doped with constant of
the Er2O3 contents at 2.5 mol% and 0.25, 0.50, 0.75 and 1
mol% of Nd2O3 has been obtained. In addition, another three
samples co-doped with constant of Nd2O3 contents at 0.50
mol% and 1.50, 2.00 and 3.00 mol% respectively has been
synthesized. An appropriate mixture of 20 gm. batch is place
in a silica crucible. The mixture was then mixed to become
homogenous by using milling machine at about 30 minutes.
The mixture was heated in an electrical furnace at 1100C
and then kept inside about 1 hour for the glass to melt. The
melts were quenched into a steel plate mould before
transferred to another furnace at 450 C for 3 hours and then
the sample was allowed to cool down to room temperature.
It is needless to mention that all our samples have been
prepared with starting materials from Aldrich products with
99.99% purity.
Series of samarium-doped magnesium phosphate glasses,
with composition of [P2O5]50-x-[MgO]50-[Sm2O3]x with

ISSN 0853-0823

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

11

0x4 mol% has been prepared from chemically pure raw


materials by employing the melt quenching technique.
Analytical grade of P2O5 MgO and Sm2O3 (purity 99.99%)
were used as starting materials. The corresponding weights
of the starting materials were mixed thoroughly in an
alumina crucible and then placed in air before heated inside
the electric furnace at temperature 900oC-1300oC depending
on composition. Higher temperature was needed for higher
Sm2O3 content. The melts was then poured on a brass plates
before being annealed at 300oC for 3 hours and then allowed
to cool down to room temperature.
The raw materials of TeO2-Na2O- Yb2O3 glass
were obtained in the powder form. All samples were
synthesized using the conventional melt-quenching
technique. An appropriate amount of TeO2, Na2O and Yb2O3
powder were mixed properly in a silica crucible to the
corresponding composition. 10 gm batches of mixture were
heated at 900 C for about 1 hour. After the mixture was
completely melted, it was quenched on a mould of stainless
steel and then allowed to anneal at 300 C for 2 hours.
Finally, the furnace was switched off to allow the sample to
cool down to the ambient temperature to remove thermal
strains. All samples obtained are transparent and have good
optical quality.
The gel with composition of P2O5-Al2O3-Na2O glass was
prepared by dissolving aluminum isopropoxide in
isopropanol at about 150C for 2 hours. Then, a mixture of
HNO3 and NaNO3 in water is added to the alcoholic solution
of aluminum isopropoxide and stirred again at 150C until a
clear solution is attained before another alcoholic solution of
H3PO4 (obtain by dissolving H3PO4 in isopropanol) is
added. To the resulting solution, YbCl3 solution is added
and was stirred together under reflux at the same
temperature for 3 hours, where a completely clear and stable
solution formed. The gel was obtained after all the solvent
and water evaporated. The formed gels were then left in a
furnace at 150C for 72 hours so that the xerogel powders
were formed. The powder as finally melted in air at
temperature around 950C. The melts was quickly poured on
a stainless steel plate to obtain the colorless glass.
A.2. X-Ray Diffraction
The X-ray diffraction studies of all the samples (in
powder form) has been carried out using an automatic
Phillips made X ray powder diffractometer to verify the
amorphosity of the samples. The running voltage of 30 kV
and current 20 mA has been employed. A radiation source
of Cu-K (=1.5418) is used with a step scan of 0.52. The
degree of crystallinity of the samples was also determined
by using a Siemens made diffractometer D5000 model,
equipped with diffraction software analysis. Diffraction
patterns were collected in the 2 range from 10 to 80o, in
steps of 0.04o and 4s counting time per step. Meanwhile, the
actual glass composition is determined using Energy
Dispersive X-ray microanalysis (EDX).
A.3. Infrared Spectral Studies
The infrared absorption spectra of the glass samples were
recorded
using
a
Perkin-Elmer
double
beam
spectrophotometer in conjunction with the KBr disc
technique, over the spectral range of 4000450 cm1 at room
temperature. Glass powdered samples of 4mg were

thoroughly mixed and grounded with 200 mg KBr, after


which the mixtures were pressed at 10 tonnes/cm2 for 5 min
into a pellet with a surface area of 1 cm2. The IR spectra
were recorded in the spectral range 4000-400 cm-1 using
Perkin Elmer Fourier Transform Infra Red (FTIR)
spectrophotometer. The UV-Vis spectroscopy is recorded to
determine the absorption characteristic at 200700 nm.
A.4. Optical Absorption
The optical absorption spectra are recorded at room
temperature. These curves are traced for highly polished
glass samples of ~3mm thickness using a Perkin-Elmer
spectrophotometer in the wavelength range of 200 - 800nm.
The optical spectra of the sample are recorded using the
UV-Vis spectroscopy in the region of 300 700 nm at
normal incidence. The transmission cut-off spectra in the
visible and ultraviolet region were recorded at room
temperature. These curves have been traced for highly
polished glass samples of ~3mm thickness using a PerkinElmer spectrophotometer in the wavelength range of 200800 nm
.
B. Theoretical
The optical absorption coefficient () for each photon
energy was calculated by using the relation
A
( ) =
(1)
d
where A is the absorbance and d is the thickness of the
sample.
In amorphous materials the absorption due to the
electronic transition within the band in relation with the
optical band gap is described by Mott and Davis [16]
equation given by,

( )h = Constant h E g

)n

(2)

where () is the absorption coefficient, the photon


energy (Eg ) the optical band gap and n is an index which
can have any values between and 3 depending on the
nature of interband electronic transition. It has been found
out that for most amorphous material n is equal to 2 (indirect
band gap) will gives reasonable fit to Equation (2) [17]. This
means that at these allowed indirect transition, the
interactions of photons with lattice vibrations will take
place. The value of Eg can be obtained by extrapolating the
linear part of the (() ) versus photon energy,
graph to the x-axis. The absorption coefficient, () in the
optical region near the absorption edge, at certain
temperature, always obey the Urbach equation [18] can be
written as,
( ) = Constant exp h E
(3)
where E is called Urbach energy that can be interpreted
as the width of the localized state in the normally forbidden
gap. This energy can be obtained by the inverse of the slope
ln() against photon energy ().

( (

))

C. Result and Discussion


The compositions and the XRD patterns of the Er3+/ Nd3+
co-doped magnesium phosphate glasses obtained are
depicted in Table 1 and Figure 1 respectively. The
amorphous nature of all the glass samples is confirmed by a

ISSN 0853-0823

12

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

broad halo as that is a characteristic of amorphous structure


[19].
TABLE 1. THE COMPOSITIONS (MOL %) OF THE
MAGNESIUM
Sample
No.

P2O5
(mol%)

MgO
(mol%)

Er2O3
(mol%)

Nd2O3
(mol%)

S1

50.00

47.25

2.50

0.25

S2

50.00

47.00

2.50

0.50

S3

50.00

46.75

2.50

0.75

S4

50.00

46.50

2.50

1.00

S5

50.00

48.00

1.50

0.50

S6

50.00

47.50

2.50

0.50

S7

50.00

46.50

3.00

0.50

the non-bridging oxygen. A small absorption bands occur


around 764 cm-1-769 cm-1 are due to the symmetric
stretching modes of the P-O-P linkages and (P-O-P)s. The
other absorption bands around 471 cm-1474 cm-1 can be
assigned to the P-O-P bending vibration are actually the
signature of the phosphate glass IR transmission spectra
[23].
Figure 3 shows the UV-Vis spectra of the sample 1 with
the thickness ~2.90 mm. Using Figure 3 the cut-off
wavelengths is obtained and is summarized in Table 2.

Figure 1. X-ray diffraction pattern of the glass samples. phosphate


glasses.

The absorption spectra of all 7 samples are shown in


Figure 2. It is clear that the absorption bands are around
3420 cm-1-3560 cm-1 and that can be assigned to the
vibration peaks of OH- band. The strong absorption reflects
the higher degree of hygroscopicity of glass samples and can
be minimized using high purity materials.
Figure 3: Optical transmittance spectra for Sample 1
TABLE

Figure 2. FTIR spectra for the glass samples.

This problem can also be avoided by preparing the


samples in high vacuum or in the environment of nitrogen
gas. The occurrence of strong absorption bands around 1312
cm-1-1319 cm-1 can be assigned to the stretching vibration of
P=O [20, 21] that appear in the branching group of Q3
tetrahedral site exist in most phosphate glass system.
However, these bands shifted towards a lower wave number
as the amount of Er3+/Nd3+ content increase indicating that
there are some structural changes occur in the glass network.
The bands around 1060 cm-1-1083 cm-1 is assigned to the
asymmetrical stretching vibration of P-O-P bonds. The
bands for (P-O-P)as-s shifts to lower wave number with
increasing of Er2O3 and Nd2O3 content, presumably, due to
the changes in the phosphate chain length and the chain PO-P band angle. Another absorption bands are located
around 933 cm-1-938 cm-1 are due to the stretching vibration
of P-O- [22]. A shift towards lower bands is believed to be
due to the existence of the bonds between Er3+/Nd3+ ions and

2.

CUT-OFF WAVELENGTH
CONTENTS.

VERSUS

Sample No.

MgO (mol%)

cut-off (nm)

S1

47.25

322

S2

47.00

321

S3

46.75

319

S4

46.50

318

S5

48.00

321

S6

47.50

320

S7

46.50

318

MGO

A plot for the cut-off wavelength versus sample number


generated from Table 2 is depicted in Figure 4.

Figure 4. Cut-off wavelength versus sample number obtained from


Table 2.

ISSN 0853-0823

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

13

It is clear that the cut-off wavelength shows gradual


decrement with decreasing MgO contents that is due to the
change in glass network structures.
A series of samples of samarium doped magnesium
phosphate glass having a composition of [P2O5]50-x-[MgO]
50-[Sm2O3] x, with 0x4 mol% is represented in Table 3
and the corresponding X-ray diffraction pattern Is shown in
Figure 5.
TABLE 3. NOMINAL COMPOSITION OF
[P2O5]50-x[MgO]50[Sm2O3] x WITH 0X4 MOL% GLASSE SAMPLES.
Sample
No.

Nominal composition (mol%)


P2O5

MgO

Sm2O3

S1

50

50

S2

49

50

S3

48

50

S4

47

50

S5

46

50

Figure 5. X-ray diffraction pattern of samples

The Table 3 shows the possibility of wide glass-forming


region is phosphate glasses. The absence of sharp peaks in
the measured X-ray spectra confirms the amorphous nature
of the sample.

50.0

45

40

S5
35

30

%T

25

S4
S3
S2

20

15

S1

10

0.0
4000.0

3600

3200

2800

2400

2000

1800

1600

1400

1200

1000

800

600

400.0

cm-1

Figure 6. IR absorption spectra of the [P2O5]50-x -[MgO]50-[Sm2O3] x glass system with x=0, 1, 2, 3 and 4 mol%.

The room temperature infrared spectra for all five samples is


shown in Figure 6.The spectra reveals broad, strong and
weak absorption bands over the investigated range of wave
numbers namely, 4000 cm-1-400 cm-1. The comparison
shows the sample modifications with the increment of
Sm2O3 contents in the sample. The frequencies of
predominant absorption peaks are characterized and
presented in Table 4. Figure 6 clearly shows the existence of
a low frequency envelope around 462-474 cm-1consist of
one absorption band. This band is assigned as a bending
vibration of O-P-O units, (PO2) modes of (PO-2) chain

groups [24, 25], while a fundamental band at ~500 cm-1 can


be ascribed as a fundamental frequency of (PO43-) [26] or as
harmonics of P=O bending vibration [26]. The absorption
band around 753-769 cm-1is assigned to the symmetric
stretch of P-O-P bridges and vs(P-O-P) those are generally
the characteristics of cyclic meta-phosphates. Furthermore,
Shikerkar et al [27] reported the occurrence of P-O-P
symmetric stretch absorption band around 731-778 cm-1.

ISSN 0853-0823

14

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

TABLE 4. THE IR PEAKS POSITIONS OF [P2O5]50-x-[MgO]50[Sm2O3]x GLASSES WITH 0X4 MOL%


IR absorption Peaks Positions (cm-1)

Sample
no.
S1

472.05

752.99

931.82

1086.31

1341.07

3425.98

S2

462.99

757.00

927.77

1076.92

1333.67

3413.79

S3

467.03

762.94

931.59

1068.53

1323.49

3413.87

S4

474.95

764.80

931.36

1074.12

1315.05

3414.16

S5

474.81

768.84

927.80

1082.51

1307.34

3414.20

The frequency of vibration of P-O-P bonds enhances


with increasing Sm2O3 content as clearly depicted in Figure
7(a). However, at about 3.0 mol% of Sm2O3, the rate of
increment in frequency starts decreasing. This phenomenon
indicates that at this point Sm2O3 may act as a modifier.
The report of Moustafa [28] on IR absorption in the
region 850-1200 cm-1 showed the sensitiveness for the
different meta-phosphate groups in the form of chain-, ring-,
and terminal groups [29, 30]. The spectral analysis of this
region shows the existence of seven bands: 927, 931, 1068,
1074, 1076, 1082 and 1086 cm-1. The asymmetric stretch of
P-O-P bridges, vas (P-O-P) occur around 931-928cm-1, while
the absorption band around 1050-1071cm-1can be assigned
as symmetric stretch of PO2. It is also found that PO2
absorption band occur at 1068-1086 cm-1 and increases with
increasing of Sm2O3. The bands around 1307-1341cm-1 are
assigned to P=O asymmetric stretching vibration modes, in
which the bands become sharper by the increment of Sm2O3.
This attribution is based on the assumption that increasing
Sm2O3 content leads to a breakdown of some terminal bonds
(P=O) in the glass network. Such alteration in the structure
decreases the bond strength and consequently, the band
centre is shifted towards lower wave number as observed
[28]. The Sm2O3 concentration dependence of P=O vibration
frequency mode is shown in Figure 7(b). It can be seen that
as the Sm2O3 content increases the frequency of vibration
decreases and the behavior is similar to the P-O-P mode
explained in Figure 7(a).
The absorption around 3400 cm-1 is due to OH- ion that
demonstrates the presence of small amount of water in the
sample. The report of Bridge and Patel [31] indicated that
the intake of water occurred mostly during the preparation
of the pellets. However, the width of the bands, which
become progressively weaker with increasing Sm2O3
content, shows that the OH- groups are an integral part of
the glass network and causes considerable absorption.
Although, the presence of small amount of OH- does not
dramatically change the structure of phosphate glasses.
However, the calculation on the relative integrated area for
these bands show that it is reduced as the Sm2O3 content is
increased which is shown in Figure 8. Relatively broad
absorption band of OH- bonds reflects the amount of these
groups present in a sample.

(a)

(b)
Figure 7. Variation of frequency of vibration of P-O-P mode (a)
and P=O mode (b) as a function of Sm2O3 concentration.

Figure 8. Relative area of OH- absorption band (cm2) of [P2O5]50x[MgO] 50-[Sm2O3] x, with 0x4 mol% glass system.

It is clear from the plot that as the concentration of


Sm2O3 is increased, the glass becomes more stable.
Moreover, there is an indication that if the dopant
concentration is higher than some threshold, the glass
becomes less stable. This is expected, because the Sm3+ ions
ISSN 0853-0823

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

15

either enter in the lattice sites by replacement of OH- ions or


reside between the lattices. In case of low Sm3+ content, the
Sm3+ ions would replace the OH- from the lattices, and thus,
the glass becomes more stable. On the other hand, for higher
Sm3+ content (>3 mol%) the rate of OH- absorption is
reduced but the PO4 structure may experience changes in
their structural network that may reduce the stability of the
glass. In this case, the dopant may be featured as an oxide
modifier confirming our earlier explanation.
Now we turn our attention to the absorption edge
studies for the estimation of the optical energy gap (Eopt)
and the Urbach energy (E) and their dependence on the
nature of samples with varying Sm2O3 concentration.
Table 5 shows the composition of our prepared
samples. All glasses are found to be very stable except the
one with more than 3 mol% of samarium oxide where the
glass is easily devitrified. Figure 9 shows the plot of
(()) versus photon energy () for the sample S2. By
extrapolating the linear part of the graph to the x-axis, the
optical energy gaps are obtained and are listed in Table 5.

which indicates the increasing amount of non-bridging


oxygen in the glass network. In this sense, the MgO seem
successfully opens up the chain by breaking up the oxygen
bond which finally forming the bridge between the corners
of the PO4 tetrahedra.

Figure 10. Plot for the variation of Eg versus Sm2O3 mol%.

Figure 11 shows the plot of ln() against photon energy


() of S2. The Urbach energy, were found to be in the
range of 0.28-0.37 eV and is depicted in Table 5.

TABLE 5. NOMINAL COMPOSITION OF THE P2O5-MgOSm2O3 GLASSES AND THE VALUES OF Eg AND
E EXTRACTED FROM THE PLOT.
Sample
number

Nominal composition
(mol%)
P2O5 MgO Sm2O

Optical band
gap (Eg) in eV

Urbach energy
(E) in eV

S1

50

50

4.35

0.32

S2

49

50

4.45

0.28

S3

48

50

4.30

0.29

S4

47

50

4.30

0.37

Figure 11. Graph ln() versus photon energy () of S2.

Figure 9. Typical plot of (() ) versus photon energy ()


for S2.

The existence of the linear dependence of the graph on


photon energy suggested that the transition is of indirect
nature. The dependence of Eg on Sm2O3 content is shown in
Figure 10. It can clearly be seen that the optical band gap
increases gradually as the Sm2O3 content is increased.
However, as the amount of Sm2O3 is about 1 mol%, the gap
is decreased. This indicates that the addition of Sm2O3 into
the glass network in the presence of MgO that acts as a
modifier impinged the network into a more compact
structure through the formation of P-O-Mg cross-links. As a
result, the hardness as well as the Youngs modulus also
increased with the MgO content [32]. However, if the
amount of Sm2O3 is further increased, the effect is lessening

The variation of Urbach energy (E) as a function of


Sm2O3 contents is presented in Figure 12. It is observed that
the Urbach energy decreases as the Sm2O3 concentration
increases indicating the fact that the indirect nature of the
band gap becomes weaker. However, at 1.0 mol% of
Sm2O3, the effect gets weaker and thereby the energy is
increased. This result shows that the width of the localized
state in the forbidden gap become smaller and indicates that
the density of electrons with higher energy becomes smaller
thus reducing the probability of electronic transition from
the valence band to the conduction band. As a result, we
observe a decrease in the optical band gap as has been
shown in Figure 10. However, with an increase of 1.0 mol%
of Sm2O3 the probability of electronic transitions is
increased. At this point, the absorption coefficient is slightly
increased as has also been observed by Sahar et al [32] in
the samarium doped Mn phosphate glasses.
Now we focus our attention to the tellurite glass. In this
study, Ytterbium doped sodium-tellurite glasses having
composition of (80x)TeO2 20Na2O-(x)Yb2O3 (x = 0.0-2.0
mol%) are prepared and X-ray diffraction technique and the
transmission spectroscopy is employed for characterization.
The absorption vibrational spectra is analyzed is detail.
Table 6 shows the nominal composition of the prepared
samples. The X-ray diffraction patterns for the samples with
Yb2O3 concentration from 0.5 to 2.0 mol% are shown in
Figure 13.

ISSN 0853-0823

16

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

oscillations of the TeO bonds. The position of all peaks in


TeO2 spectrum showed a close agreement with those
previously reported [33]. Rare earths connected to the
chains of TeO4 groups are identified based on the
simultaneous presence of the bands at ~721 cm-1-732 cm1.
For rare earth-doped glasses, new bands were detected
around 592, 599, 600, 580 and 606 cm1 respectively and
those are attributed to YbO stretching vibrations.
TABLE 7. INFRARED ABSORPTION PEAKS FOR ALL
SAMPLES.
IR peaks(cm-1)

Sample Number
1

3423

1643

1382

732

606

Figure 12. Variation of the Urbach energy (E) as a function of


Sm2O3 contents at room temperature.

3423

1640

1381

727

600

3417

1637

1380

727

599

The X-ray diffraction pattern shows expected broad halo


characteristic of amorphous samples. The broad peaks
around 30 show the characteristic of most tellurite glasses.

3412

1632

1377

724

592

3405

1635

1377

721

589

TABLE

6.

THE NOMINAL COMPOSITION


PREPARED SAMPLES.

Sample
number

OF

THE
54
50
45

Nominal Composition
(mol%)

T%

40

TeO2

Na2O

Yb2O3

S1

70.0

30

35

S2

69.5

30

0.5

30

S3

69.0

30

1.0

25

S4

68.5

30

1.5

20

S5

64.0

30

2.0

15
10
5

4000

3000

2000

1500

1000

450

Figure 14. IR absorption spectra for TeO2RE glass system.

The main absorption frequencies around 730 cm-1600


cm as shown in Figure 14 are assigned to the stretching
vibrations of the TeO bonds. The spectra of the all studied
glasses showed the typical broadening of the observed
bands. In most cases, they are very similar to the spectra of
the crystalline phases. This is the direct proof for the
similarities of the structural units and of the short-range
order believed to be crystalline. Due to this reason, the
spectra of tellurite glasses are interpreted based on their
crystalline phases. The vibrations of a specific group of
atoms in a lattice are regarded as relatively independent
from motions of the rest of the atoms, as mentioned earlier
[34, 35], according to an assumption in vibrational
spectroscopy of the solid-state [36]. The concept of separate
vibrations for glasses was discussed in detail for the first
time in the work of Tarte [35-40]. A similar approximation
is used in the present investigation. This empirical analysis
can give some useful information concerning the
arrangement of atoms in glasses.
1

Figure 13. X-ray diffraction patterns for the glass system. (80 x)
TeO2 20Na2O - (x)Yb2O3 glass system.

The positions of the absorption bands of these glasses are


listed in Table 7 with the specification of their
corresponding attributed vibrational modes. For all glasses
presented in this work, the spectra occurs at range 3405 cm1
-3423 cm-1, 1632 cm-1-1643cm-1, 1377 cm-1-1382 cm-1, 721
cm-1-732 cm-1 and 589 cm-1-606 cm-1. The predominant
peaks around700 cm-1 is attributed to symmetrical vibration

ISSN 0853-0823

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

17

III. CONCLUSION
In conclusion, the phosphate glass sample has
successfully been made and their amorphous nature has
been confirmed by X-ray diffraction techniques. The IR
spectra show that there are six broad absorption peaks
corresponding to the OH- band, P=O, P-O-P, P-O- stretching
vibrations. The absorption cut-off wavelength in the range
318 nm to 322 nm depends strongly on the concentration of
MgO in the sample.
In a separate study the [P2O5]50-x-[MgO]50-[Sm2O3]x
glass with 0x4 mol% has been prepared again by melt
quenching technique. XRD shows all samples are glass in
nature. While, IR shows that the glass network is dominated
by the existence of asymmetric and symmetric of P-O-P
bridges around 931 cm-1-928 cm-1 and 753 cm-1-769 cm-1
respectively. While, P=O asymmetric stretching vibration
modes exists around 1307 cm-1-1341cm-1. For low Sm3+
content, the glass becomes more stable but only up to 3
mol% when the glass start to be precisely influence by the
existence of broad OH peak. In this case, the Sm2O3 may act
as a modifier.
To estimate the relevant optical parameter a series of
samarium doped magnesium phosphate glasses have been
made and the optical absorption characteristics showed that
the transition is of indirect type with n=2. The doping of
Sm2O3 into the glass network with the present of MgO as
modifying oxides does not alter very much the optical
nature of band gap as well as the Urbach energy tail. Only
after the further addition of 1 mol% Sm2O3 the optical band
gap is affected because the formation range of the glass
become very small beyond a certain threshold concentration
of the samarium ion.
We further investigate the optical behavior of a series of
ytterbium doped sodium tellurite glasses. Glasses with more
than 2 mol% of Yb2O3 are found to be chemically unstable.
The X-ray diffraction pattern confirms the amorphous
nature of the samples. The infrared absorption spectra for all
the glasses showed bands around 3500, and 600 cm1. Rare
earths connected to the chains of TeO4 groups identified
based on the simultaneous presence of the bands at 589 to
606 cm1. Our spectroscopic investigations on phosphate
and tellurite samples prepared by sol-gel techniques will be
reported in future communications.
Temperature and concentration dependence of the
multi-phonon relaxation rates, radiative decay and upconversion luminescence is worth to look at. It is hoped that
our detail systematic experimental study may provide useful
information for exploiting rare earth doped tellurite and
phosphate glasses in fabricating up-conversion lasers. In
addition, there is a need to extend our studies to examine the
temporal behavior of up-conversion emission and Stokes
emission for red and green transition. A complete
microscopic picture on optical characteristics, however,
require rigorous theoretical model and simulation on the
local bonding environment and the local vibrational density
of states of the rare earth ion within the matrix because
some of these vibrational modes may or may not be coupled
to electronic excited states. The effect of embedded metallic
nanoparticles on the optical properties of these glasses will
be reported elsewhere.

ACKNOWLEDGEMENTS
S.K. Ghoshal especially thanks to the Physics Department,
Universiti Teknologi Malaysia for providing research
facilities to complete this work.
REFERENCES
[1] Jong-Oh Byun, Byong-Ho Kim, Kun-Sun Hong , Hyung-Jin Jung ,
Sang-won Lee, and A.A. Izyneev, J. Non-Cryst. Sol. 190 (1995) 288.
[2] A. Mogus-Milankovic, V. Licina, S.T. Reis, and D.E. Day, J. NonCryst. Sol. 353 (2007) 2659.
[3] E. Metwalli, M. Karabulut, D.L. Sidebottom, M.M. Morsi, and R.K.
Brow, J. Non-Cryst. Sol, 344 (2004) 128.
[4] H. Desirena, E. De la Rosa, L.A. Daz-Torres, and G.A. Kumar, Opt.
Mater. 28 (2006) 560.
[5] Sherief M. Abo-Naf, N.A. Ghoneim, and H. A. El-Batal, J. Mater. Sc.:
Mater. in Elect. 15 (2004) 273.
[6] Chun Jiang, Fuxi Gan, Junzhou Zhang, Peizhen Deng, and Guosong
Huang, Mater. Lett. 41 (1999) 209.
[7] M.J. Weber, J.D. Myers, and D.H. Blackburn, J. Appl. Phys. 52 (1981)
2944.
[8] H. Nii, K. Ozaki, M. Herren, and M. Morita, J. Lumin. 7677 (1998)
116.
[9] J.S. Wang, E.M. Vogel, and S. Snitzer, Opt. Mater. 3 (1994) 187.
[10] S. Xu, D. Fang, Z. Zhang, and Z. Jiang, J. Solid State Chem. 178
(2005) 1817.
[11] W. Ryba-Romanwski, S. Golab, L. Cichosz, and B. JezowaskaTrzebiatawska, J. Non- Cryst. Sol. 105 (1988) 295.
[12] S. Biswal, J. Nees, A. Nishimura, H. Takuma,and G.
Mourou,Opt.Comm. 160 (1999) 92.
[13] Brian M. Walsh, Norman P. Barnes, and Russell and J. DeYoung.
Lath. Glass Spectroscopy and Fiber Laser, 2008.
[14] O. Glatter, and O.Kratky Small Angel X-ray Scattering Academic
Press, 1982.
[15] I. Simon, Modern Aspects of the Vitreous State, Butterworth, London,
1964.
[16] N.F. Mott, and E.A. Davis, Phil. Mag. 28 (1970) 903.
[17] A.A. Higazy, B.Y. El-Baradic, and M.I. Abd El-Ati, J. Mater. Sci.
Lett. 11 (1992) 581.
[18] F. Urbach, Phys. Rev. 92 (1953) 1324.
[19] F.F. Sene, J.R. Martinelli, and L. Gomes, J. Non-Cryst. Sol. 348
(2004) 63.
[20] M.R. Sahar, A. Wahab M.A. Hussein, and R. Hussin, J. Non-Cryst.
Sol. 353 (2007) 1134.
[21] S.M. Abo-Naf , M.S. El-Amiry, and A.A. Abdel-Khalek, Opt. Mat. 30
(2008) 900.
[22] P.Y. Shih, S.W. Yung, and T.S. Chin. J. Non-Cryst. Sol. 244 (1999)
211.
[23] Dora Ilieva, Bojidar Jivov, Georgi Bogachev, Christo Petkov, Ivan
Penkov, and Yanko Dimitriev, J. Non-Cryst. Sol. 283 (2001) 195.
[24] P. Znik, M. Jamnick, J. Non-Cryst. Solids 146 (1992) 74.
[25] A. Abdel-Kader, A.A. Higazy, and M.M. Elkholy, J. Mater. Sci.:
Mater. Electron. 2 (1991) 157.
[26] K. Nakamoto, Infra-Red Spectra of Inorganic and Coordination
Compounds, Wiley, New York, 1963.
[27] A.G. Shikerkar, J.A.E. Desa, P.S.R. Krishna, and R.Chitra, J. NonCryst. Sol. 270 (2000) 234.
[28] Y.M. Moustafa, and K.El. Egili, J. Non-Cryst. Sol. 240 (1998) 144.
[29] E.I. Kamitsos, A.P. Patsis, M.A. Karakassides, and G.D. Chryssikos, J.
Non-Cryst. Sol. 126 (1990) 52.
[30] E.I. Kamitsos, J.A. Kapoutsis, G.D. Chryssikos, J.M. Hutchinson,
A.J.Pappin, M.D. Ingram, and J.A. Duffy, Phys. Chem. Glasses 36 (3)
(1995) 141.
[31] B. Bridge and N.D. Patel, J. Non-Cryst. Sol. 91 (1987) 27.
[32] M.R. Sahar, and A.Z. Abidin, J. Mater. Sci. Lett. 13 (1994) 227.
[33] M.R. Sahar, and B. Astuti, Optical Review (2007).
[34] M. Arnaudov, V. Dimitrov, and Y. Dimitriev, Mater. Res. Bull. 17
(1982) 1121.
[35] P. Tarte, and J.A. Prins (Ed.), Physics of Non-Cystalline Solid, North
Holland, Amsterdam, 1964, p. 549.
[36] O. Lindqvist, Acta Chem. Scand. 22 (2007) 977.
[37] R.K. Brow, R.J. Kirkpatrick, and G.L. Turner, J. Non-Cryst. Sol. 116
(1990) 39.
[38] X. Fang, C.S. Ray, A. Mogus-Milankovic and D.E. Day, J. NonCryst. Sol. 283 (2001) 162.

ISSN 0853-0823

18

Prosiding Pertemuan Ilmiah XXV HFI Jateng & DIY

[39] C. Nelson and D.R. Tallant, Phys. Chem. Glass. 26 (1985) 119.
[40] R.Hussin, A.S. Musdalilah., A. Nur Shahira, A. Mutia Suhaibah, A.
Suhailah, A.F. Siti Aishah, H.

Sinin and M.Y. Mohd Noor, J. Fund. Sc. 2008 (accepted for publication).

ISSN 0853-0823

Das könnte Ihnen auch gefallen