You are on page 1of 6

Optical Materials 32 (2009) 339344

Contents lists available at ScienceDirect

Optical Materials
journal homepage: www.elsevier.com/locate/optmat

JuddOfelt parameters and radiative properties of Sm3+ ions doped zinc


bismuth borate glasses
A. Agarwal a, I. Pal a, S. Sanghi a,*, M.P. Aggarwal b
a
b

Department of Applied Physics, Guru Jambheshwar University of Science and Technology, Hisar 125 001, Haryana, India
Department of Physics, Guru Nanak Khalsa College, Yamunanagar 135 001, Haryana, India

a r t i c l e

i n f o

Article history:
Received 28 February 2009
Received in revised form 8 August 2009
Accepted 20 August 2009
Available online 22 September 2009
PACS:
42.70.Ce
42.70.Hj
32.70.Cs
78.40.q
78.55.m

a b s t r a c t
Glasses having composition 20ZnOxBi2O3(79  x)B2O3 (15 6 x 6 35 mol%) and doped with 1 mol% Sm3+
ions have been prepared by melt quench technique. Optical absorption and uorescence spectra have
been recorded. JuddOfelt approach has been applied for the f M f transition of Sm3+ ions to evaluate various intensity parameters (X2, X4, X6). The variations in intensity parameters, radiative transition probabilities and hypersensitive band positions with composition have been discussed. The variation of X2
with Bi2O3 content have been attributed to change in the asymmetry of the ligand eld at the rare earth
ion site and to the changes in their rare earthoxygen (REO) covalency, whereas the variation of X6 is
found to be strongly dependent on nephlauxetic effect. The shift of the hypersensitive band shows that
the covalency of the REO bond increases with increase of Bi2O3 content due to increased interaction
between the Sm3+ ion and non-bridging oxygens. The radiative transition probabilities for the Sm3+ ions
are large in zinc bismuth borate glasses, suggesting their suitability as laser material.
2009 Elsevier B.V. All rights reserved.

Keywords:
Laser materials
Glasses
Sm3+ ions
JuddOfelt theory

1. Introduction
Glasses doped with rare earth ions are used for solid state lasers,
solar concentrators, optical detectors, waveguide lasers, optical bers for uorescent display devices, etc. The investigations of
absorption and luminescent properties of the Sm3+ ions have indicated that the optical properties of these rare earth ions can be affected by varying the glass composition [14], which makes such
glasses as suitable host to achieve laser radiation due to appearance
of sharp and unambiguous absorption bands in their optical
absorption spectra. Focus on rare earth doped glasses is not limited
only to infrared optical devices but there is growing interest in visible optical systems [59]. Oxide glasses are stable hosts for obtaining efcient luminescence in rare earth ions. Borate glass is a
suitable optical material with high transparency, low melting point,
high thermal stability, and good rare earth ion solubility [10,11].
Tellurite and heavy metal oxide glasses catch much attention in recent years because their maximum phonon energy is lower than
those in borate, phosphate, silicates and germanate glasses. The
* Corresponding author. Tel.: +91 1662 263385; fax: +91 1662 276240.
E-mail address: sutkash@yahoo.com (S. Sanghi).
0925-3467/$ - see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.optmat.2009.08.012

emission quantum efciency from a given level depends strongly


on the phonon energy of the host medium, therefore, it is expected
that the non-radiative loss to the lattice will be small and the uorescence quantum efciency will be high in these glasses [12,13].
The intensity of the transitions for the rare earth ion can be calculated by using the JuddOfelt theory [14,15]. This theory dene a
set of three intensity parameters Xt (t = 2, 4, 6) which are sensitive
to the environment of the rare earth ion. These intensity parameters are used to calculate radiative transition probability for spontaneous emission, radiative life time of the excited state, branching
ratios (which predict the uorescence intensity of laser transitions)
and induced emission cross-section in order to optimize the best
conguration of the ion-host to improve the laser efciency of a
specic electronic transitions. The absorption and emission properties of rare earth ions for different glass compositions have been
investigated.
In particular, the optical properties of the Sm3+ ions have been
reported in large number of glass systems such as borate [16],
oxyuoro-borate [17], borotellurate [18] and borosulphate [19].
Saisudha and Ramakrishna [20] have reported a systematic Judd
Ofelt analysis of Sm3+ ions in binary PbOB2O3 glass systems and
showed that the intensity parameter X6, which is an indicator of

340

A. Agarwal et al. / Optical Materials 32 (2009) 339344

the covalency of the SmO bond, is sensitive to the structural


changes in glass due to the conversion of three coordinated boron
atoms (B3) to four coordinated boron atom (B4) and the changes in
the number of non-bridging oxygen ions (NBOs). In this paper, we
present a study of the optical absorption and emission spectra of
Sm3+ ions doped zinc bismuth borate glasses. These heavy metal
oxide based glasses have been chosen as these have large glass
forming region, high refractive index, good physical and chemical
stability and large transmission window.
2. The JuddOfelt Theory

mc2
pe2 N

amdm

where m and e are the mass and charge of the electron, respectively, c the speed of light and N the number of active rare earth
ions per cm3. The absorption coefcient a(m) is obtained from the
, where T
transmission spectra by using the relation am ln100=T
d
is the transmission andd the sample thickness. From the Judd
Ofelt theory, the oscillator strength of an electronic transition from
an initial state aJ to a nal state bJ0 is given by:

fcal

8p2 mn2 2m 
0
0 
Sed aJ; bJ n3 Smd aJ; bJ

2
27nhe 2J 1

where m is the frequency of transmission for J ? J0 , n is the refractive


index, h is the Plancks constant, Sed and Smd are the line strength for
electric and magnetic dipole transitions, given by

Sed e2
Smd

Xt jhaJjU t jbJ0 ij2 and

t2;4;6

eh
2pmc

2 X

2a
0

Xt jhaJjL 2SjbJ ij2

2b

t2;4;6

The JuddOfelt parameters, Xt, are closely related to the active ion
environment given by [20]

Xt 2t 1

(i)

the spontaneous emission probability, Arad, from an initial


state aJ to the nal state bJ0 for a particular electric dipole transition is given by

Arad

The absorption spectra of rare earth ions serve as a basis for


understanding their radiative properties. The sharp absorption line
arising from the 4f M 4f electronics transitions can be electric dipole, magnetic dipole or electric quadrupole in character. The
quantitative calculation of the intensity of these transitions was
developed independently by Judd [14] and Ofelt [15]. A brief outline of the JuddOfelt theory is given below.
For a free rare earth ion, the electric dipole transitions between
two states within the 4f conguration are parity forbidden, while
the magnetic dipole and electric quadrupole transitions are allowed. For an ion in a medium, the electric dipole transition becomes allowed due to the admixture of states from conguration
of opposite parity (e.g., 4fN1 5d) into the 4f conguration. The
transition probability depends on the extent of admixture. The
intensity of the absorption band is estimated by the oscillator
strength fexp, which is calculated from the absorption spectra by
using the following equation

fexp

for Sm3+, all transitions observed are predominantly induced by


electric dipoles. Therefore only Sed is considered in Eq. (2). Using
fexp values to fcal in Eq. (2), the JuddOfelt parameter Xt can be calculated. Using JuddOfelt theory, the following optical parameters
are also calculated:

jAs;p j2 N2 s; t2s 11

where the terms jAs,pj2 and N2(s, t) represent the crystal-eld


parameters and the nephlauxetic effect parameter b, respectively.
The N is proportional to b, which is an indicator of the SmO
covalent bond. The term is the square matrix element of the tensorial operator Ut which connect aJ to bJ0 states. The electronic conguration of rare earth ion is characterised by partially lled 4f
orbital shielded by lled outer electronic orbital 5s2 5p6. Therefore,
no signicant changes are observed in the values of jhaJjUtjbJ0 ij2
when these ions are incorporated into different hosts. The values
of jhaJjUtjbJ0 ij2 were reported by Carnall et al. [21]. In particular,

64p4 m3 nn2 22 e2

"

27hc 2J 1

#
t

Xt jhaJjU jbJ ij

t2;4;6

(ii) the total probability, AT, obtained by carrying out the summation of all the transitions to the nal states bJ0 is given by

AT

Arad

(iii) the values of Arad and AT can be used to calculate the uorescence branching ratio br given by

br

Arad
AT

(iv) the radiative lifetime of the excited states is represented by

sr

1
AT

(v) the induced emission cross-section r for each transition is


given by

k4p Arad
8pcn2 Dk

where kp is peak wavelength and Dk is the full width at half maxima


of the uorescent peak for different transitions obtained from the
emission spectra.
3. Experimental details
Glasses having composition 20ZnOxBi2O3(79  x)B2O31Sm2O3,
(15 6 x6 35) were prepared by using melt-quenching technique.
The 15 g batches of analar grade chemicals, as per given composition, were melted in a porcelain crucible in air at temperature of
1150 C for 40 min. The melt was stirred frequently for homogeneous mixing of all the constituents. The glass samples were obtained by quenching the melt in between two stainless steel plate
at room temperature (RT). The density (D) of each glass sample
was measured by the Archimedess principle using xylene as
immersing liquid. The refractive index (n) of the prepared samples
was measured by the Brewster angle method using HeNe laser
(632 nm). The glass transition temperature (Tg) was recorded with
the help of Differential Scanning Calorimeter (DSC) (Q-10, TA Instruments) at a heating rate of 10 C per minute, under N2-gas atmosphere. The optical absorption spectra of polished samples were
reordered on a Varian-Carry 5000 spectrophotometer in the range
of 3003200 nm. The emission spectra were recorded using FluoroMax-3 Fluorimeter with a Xe-arc lamp as the excitation source at a
wavelength of 450 nm.
4. Results
From the measured values of density (D), molar volume (Vm)
and refractive index (n), values of various other physical parameters such as samarium ion concentration (N), mean rare earth ion

341

A. Agarwal et al. / Optical Materials 32 (2009) 339344

@ RMS

v
!
u P
2
u

f

f
meas
cal
P
t
2
fmeas

where fcal and fmeas are the calculated and measured oscillator
strengths, respectively, and the summation is taken over all the
bands used to calculate Xt parameters.
In the present glasses, it has been observed that X2 > X4 > X6.
Same ordering in the JO parameters has been observed by Boehm
et al. [4] in borate glasses. However, different ordering is observed
in uorophosphates and ZBLAN glasses [22,23]. The values of X2
parameter in the present glasses are higher than that observed
for uorophosphates, ZBLAN and uoride glasses but less than
the X2 values for borate and phosphate glasses. The ratio
of X4/X6 is known as spectroscopic quality factor (SQF = X4/X6)
which is found to be greater than one in present study and is
higher than phosphate glasses [4]. According to Jorgensen et al.
[25,26], X2 is covalency-dependent parameter and X4 and X6 are
structure-dependent parameters. The ratio X4/X6 in the borate
glasses is 1.71, and in the present glasses it is 1.05 (ZBS2 glass
sample). This indicates that present glasses are fairly rigid. Nageno

0.98

H5/2

F7/2

0.96
6

H5/2

F9/2

0.94

Absorbance (a. u.)

separation (ri), reection loss (RL), etc. of these glasses were calculated from the conventional formula and are presented in Table 1.
The glass transition temperature, Tg of various glasses are also
listed in Table 1. Tg decreases from 500 to 464 C with increase
in bismuth oxide content.
The optical absorption spectra of Sm3+ doped ZBS1 glass sample
is shown in Fig. 1. The oscillator strength of Sm3+ are arranged in
two groups [4], one low energy group corresponding to transitions up to 10,700 cm1 and the other high energy group corresponding to transitions in the energy range 17,60032,000 cm1.
Therefore, JuddOfelt equation (Eq. (2)) is applicable to cases
where the high f-splitting is small as compared to the fd energy
gap since it is not appropriate to use in the high-energy levels for
the calculation of Xt. Thus, in present study, the Xt parameters
are determined only for the low energy region. In the high energy
region, due to strong ultraviolet absorption by Bi3+ ions, Sm3+ peaks
are not well resolved. All the transitions of Sm3+ are electric dipolar
with signicant intensities in the observed range, i.e., 850
1650 nm.
Fig. 1 exhibits the following absorption bands for Sm3+ ions:
6
H5/2 ? 6F11/2,6F9/2, 6F7/2,6F5/2,6F3/2, and 6F1/2. The intensity parameters Xt (t = 2, 4, 6) are obtained from the experimental oscillator
strengths and the calculated double reduced matrix elements
using the expression for oscillator strengths, Eq. (2), by least square
analysis. The theoretically and experimentally calculated values of
oscillator strength of all the prepared glasses are presented in Table 2. The intensity parameters determined for Sm3+ ions for all
the samples are listed in Table 3. Reported values of Xt parameters
for borate [4], phosphate [4], uorophosphates [22], ZBLAN [23]
and uoride [24] glasses have also been given in Table 3 for comparison. In order to determine the accuracy of the obtained intensity parameters, the root-mean-square deviations (@ RMS) are
calculated by using the following relation:

H5/2

F5/2

0.92
6

0.9

H5/2

H5/2

F11/2

F3/2

H5/2

F1/2

0.88

0.86

0.84
850

1050

1250

1450

Wavelength (nm)
Fig. 1. Optical absorption spectra of Sm3+ ions in 20ZnO15Bi2O364B2O31Sm2O3
(ZBS1) glass.

et al. [27] have also studied various alkali borate and phosphate
glasses and reached to similar conclusion that X4 and X6 parameters are related to the rigidity of the glass matrix. The intensity
parameter X2 in present glasses increases from 1.93  1020 cm2
to 3.03  1020 cm2, with increase in Bi2O3 content from 15 to
35 mol%. X4 and X6 show decrease up to 25 mol% of Bi2O3 content
and then these increase. The position and intensity of certain electric dipole transition of rare earth ions are found to be very sensitive to the environment of the rare earth ion. Such transitions are
known as hypersensitive transitions [25].
The hypersensitivity of a transition has been shown to be proportional to the nephlauxetic ratio b, which is a measure of the covalency of the REO bond [28]. In present glasses, the positions of the
peak wavelength of the hypersensitive band of Sm3+ (6H5/2 ? 6F1/2;
1516 nm) is investigated as a function of glass composition to study
the nature of REO bond. The peak wavelength for Sm3+ is found to
shift towards longer wavelength from 1516 to 1524 nm with increase in Bi2O3 content from 15 to 35 mol%.
The uorescence spectra of Sm3+ recorded at room temperature
at excitation wavelength of 450 nm is shown in Fig. 2. This gure

Table 1
Density (D), molar volume (Vm), glass transition temperature (Tg), number density of Sm3+ ions (N), refractive index (n), dielectric constant (e), reection loss (RL), inter-ionic
distance (ri) and polaron radius (rp) for 20ZnOxBi2O3(79  x)B2O31Sm2O3 glasses.
Sample code

x (mol%)

D (g/cm3)

Vm (cm3/mol)

Tg (C)

N (1020 ions/cm3)

e (n2)

RL (%) (n  1/n + 1)2

ZBS1
ZBS2
ZBS3
ZBS4
ZBS5

15
20
25
30
35

4.09 0.01
4.44 0.02
4.71 0.01
5.00 0.01
5.30 0.02

32.82
34.69
36.91
38.73
40.28

500 1
483 1
479 1
473 1
464 1

4.22 0.02
4.16 0.01
4.04 0.03
3.96 0.02
3.89 0.01

1.72 0.01
1.79 0.01
1.84 0.01
1.90 0.01
1.99 0.01

2.95
3.20
3.38
3.61
3.96

7.00
8.01
8.74
9.63
10.96

ri (
A)

rp (
A)

7.49
7.47
7.39
7.34
7.29

5.37
5.39
5.45
5.49
5.52

342

A. Agarwal et al. / Optical Materials 32 (2009) 339344

Table 2
Oscillator strength for some transitions from the indicated levels to the ground level 6H5/2, and their root-mean-square deviation (@ RMS), which indicates the t quality of
theoretical and experimental results for 20ZnOxBi2O3(79  x)B2O31Sm2O3 glasses.
Transitions from ground level ?6H5/2

Oscillator strength, (f  108)

k (nm)

ZBS1

F1/2
6
F3/2
6
F5/2
6
F7/2
6
F9/2
6
F11/2
@ RMS (107)

1516
1472
1368
1226
1076
938

ZBS2

ZBS3

ZBS4

ZBS5

fmeas

fcal

fmeas

fcal

fmeas

fcal

fmeas

fcal

fmeas

fcal

93
126
118
234
172
61

144
104
125
241
171
28

106
118
92
204
178
62

142
96
94
221
167
29

122
107
102
221
171
78

143
88
105
232
169
29

152
124
118
215
217
85

163
128
101
248
190
32

187
139
125
231
220
89

170
143
109
260
198
33

5.03

5.76

6.65

9.91

11.41

Table 3
JuddOfelt intensity parameters (X2, X4, X6) of various Sm3+ doped glasses.
Glass

X2 (1020 cm2)

X4 (1020 cm2)

X6 (1020 cm2)

X4/X6

References

ZBS1
ZBS2
ZBS3
ZBS4
ZBS5
Borate
Phosphate
Fluorophosphate
ZBLAN
Fluoride

1.93 0.16
2.03 0.15
2.49 0.24
2.88 0.17
3.03 0.19
6.36
4.31
2.18
2.06
1.44

1.87 0.03
1.75 0.06
1.64 0.12
1.83 0.06
1.72 0.17
6.02
4.28
3.80
2.55
2.87

1.79 0.11
1.66 0.12
1.61 0.11
1.74 0.13
1.69 0.19
3.51
5.78
2.15
1.63
1.44

1.04
1.05
1.02
1.05
1.02
1.71
0.74
1.76
1.56
1.99

Present
Present
Present
Present
Present
[4]
[4]
[22]
[23]
[24]

Fluorescence Intensity (a.u.)

G5/2

work
work
work
work
work

very intense than others. The peaks 4G5/2 ? 6H7/2, 9/2 and 11/2 seem
to have two components. These components are most likely due
to the Stark splitting of the ionic levels.
The Stark splitting reduces with increasing Bi2O3 content in the
host glass. The uorescence spectrum has been used to derive radiative properties like radiative transition probability (Arad), branching ratio (br), stimulated emission cross-section (r) and the
radiative life time of excited state (sr), etc. and these parameters
are presented in Table 4. It is also noted that the intensity of uorescence bands decreases as the Bi2O3 increases.

H9/2

5. Discussion
4

520

G5/2

G5/2

H7/2

G5/2

H11/2

H5/2

570

620

670

Wavelength (nm)
Fig. 2. Emission spectra of Sm3+ ions in 20ZnO15Bi2O364B2O31Sm2O3 (ZBS1) glass
(kext = 450 nm).

shows four broad peaks near 536, 562, 598 and 646 nm with rst
one being very weak. The uorescence peaks are assigned to arise
due to 4G5/2 ? 6H5/2, 7/2, 9/2 and 11/2 transitions, respectively. The
uorescence peak due to transition 4G5/2 ? 6H9/2 appears to be

Jorgensen [29] has proposed that the positions of the hypersensitive bands change due to nephlauxetic effect. The nephlauxetic effect [30] arises from an expansion of the partially lled f-shell
due to charge transfer from the ligands to the core of the central
ion. Hence, this effect may be used as a measure of the covalency
of the bonds, between the rare earth ion and the surrounding oxygens in the glass. With increase in the overlap of the oxygen orbitals and the 4f-orbitals, the energy level structure of the rare earth
ion contracts, and leads to a shift of hypersensitive band transitions
towards longer wavelength (lower energy).
In present glasses, the peak wavelength of the hypersensitive
transitions for Sm3+ ion shifts towards longer wavelength with
Bi2O3 content, indicating that an increasing covalent nature for
the REO bond. Here, close packing of the structure takes place
with the formation of BO4 and BiO4 units and results in increased
interaction between the rare earth and charged non-bridging oxygens (NBO). Bi2O3 is a strong network former and oxygens are
packed closely in bismuth borate matrix [31]. The covalency of
the REO bond increases with increase in Bi2O3 content as indicated by shift of the hypersensitive bands towards longer
wavelengths and by the increase in the intensity of the transition
6
H5/2 ? 6F1/2. This implies that N(s, t) increases with Bi2O3 content.
The intensity parameter X2 is found to increase with increase of
Bi2O3 content for Sm3+ doped glasses. This indicates that the As,p

343

A. Agarwal et al. / Optical Materials 32 (2009) 339344

Table 4
The peak wavelength (kp), radiative transition probability (Arad), branching ratio (br), stimulated emission cross-section (r), total radiative transition probability (AT), radiative life
time (sr) and the total emission cross-section (rt) for 20ZnOxBi2O3(79  x)B2O31Sm2O3 glasses.
Transitions
from 4G5/2?

kp
(nm)

536
562
598
646

H5/2
H7/2
6
H9/2
6
H11/2
AT (s1)
sr (ms)
rt (1022 cm2)
6

ZBS1

ZBS2

ZBS3

ZBS4

ZBS5

Arad
(s1)

br (%) r
Arad
(1022 cm2) (s1)

br (%) r
Arad
(1022 cm2) (s1)

br (%) r
Arad
(1022 cm2) (s1)

br (%) r
Arad
(1022 cm2) (s1)

50
0.172
94
0.324
110
0.379
36
0.124
290 11
3.45 0.07
23.87

1.24
10.50
9.01
3.12

67
0.175
101
0.264
126
0.329
88
0.230
382 17
2.61 0.04
28.15

1.14
10.43
9.53
7.05

75
0.168
127
0.284
141
0.316
103
0.231
446 25
2.24 0.08
31.52

should increase with increase in Bi2O3 content. The addition of


Bi2O3 in place of B2O3 converts three coordinated boron (B3) to four
coordinated boron (B4), resulting in the conversion of boroxol units
to pentaborate groups [30]. It has been proposed [32] that in oxide
glasses, a rare earth ion is surrounded by eight neighbouring oxygens belonging to the corners of BO4, BiO4 or any other oxygen
atoms, forming an edge of the cube. Bi2O3 can form its own network without involving the borate groups. The presence of oxygen
unbounded to any boron as in 3Bi2O3.5B2O3 glasses supports this
conclusion [33].
The asymmetry of the crystal eld at the rare earth site with increase in Bi2O3 content for Sm3+ ions indicates that these ions in
present glasses might be surrounded only by bismuth groups and
the appearance or disappearance of borate groups do not affect
the symmetry of the ligand eld at the rare earth site. The X4
parameter for Sm3+ ions shows random behaviour. This behaviour
of X4 could not be attributed to either As,p or N(s, t) independently.
It must be dependent on both. X6 parameter also shows random
behaviour with increase of Bi2O3 content. Thus in present glasses,
As,p plays an important role in the variation of X2 with Bi2O3 content. X6 mainly depends on N(s, t) of Sm3+ ion. Generally, it is observed that in silicate and borate glasses [1,3437], X2 is
determined by the asymmetry of the ligand eld at the rare earth
site (i.e., As,p) and nephlauxetic effect [N(s, t)], while, X6 depends
only on the nephlauxetic effect. On the other hand in phosphate
glasses, all the three parameters depend strongly on the ionic radius of the modier [27].
The radiative transition probability (Arad) in present glasses is
large and the refractive index of the host glass plays a dominant
role. Refractive index, n of present glasses increases with Bi2O3
content from 1.72 to 1.99 (Table 1). The increase of Arad with
n can be explained as follows: the ratio of the effective eld at
the ion site in glass to the applied eld E0 in the simple isotropic
tight binding approximation is given by [38]

Eeff
1
1

3n2  1
E0

10
2

This introduces a factor nn 92 in the radiative transition proba2 12


in the stimulated emission cross-secbility in Eq. (4) and n 9n
tion of the uorescence line of rare earth ions. The refractive
2
2
index ranges from 1.72 to 1.99, the factor nn 92 in the radiative
transition probability ranges from 4 to 8. Because of the strong
dependence on refractive index, the radiative decay rates for
equal Xt values can be signicantly different. Because of large value of Arad, r is quite high for the lasing transitions of the Sm3+
ions in present glasses.
Stimulated emission cross-section (r) for the uorescence lines
of Sm3+ ions in present glasses is calculated from the uorescence
spectra (Fig. 2) using radiative transition probability. In present
glasses, the stimulated emission cross-section is found to be large
which is desired attractive feature for low-threshold, high-gain

1.21
12.41
10.09
7.81

91
0.175
152
0.292
165
0.317
112
0.215
520 29
1.92 0.09
34.35

1.38
13.93
11.08
7.96

Fig. 3. Lasing transitions of Sm3+ (4G5/2 ? 6H11/2,


borate glasses.

br (%) r
(1022 cm2)

104
0.175
171
0.287
190
0.319
129
0.217
594 32
1.68 0.03
36.52

9/2, 7/2, 5/2)

1.43
14.29
11.63
9.17

ions in zinc bismuth

applications and can be utilized to obtain CW laser action. In tellurite glasses [39] and in lead borate glasses [20], doped with Nd3+,
and also in oxyuoro-borate glasses [17] and lead uoro-borate
glasses [40], doped with Sm3+, the radiative transition probabilities
are large and hence the stimulated emission cross-section. Laser
action has been observed in lead uoro-borate glass as well in tellurite systems. The radiative transition probabilities in present
glasses are of the order of those reported in lead uoro-borate
and tellurite glasses, therefore, these glasses may be utilized as laser host material. The branching ratios are evaluated for each transition and a probable lasing transition of the rare earth ion is
shown in Fig. 3.
6. Conclusions
Using the JuddOfelt theory, the three intensity parameters,
spontaneous emission probabilities, radiative life time and stimulated emission cross-section of Sm3+ ions doped in zinc bismuth
borate glasses are determined. The change in the positions and
intensity parameters of the transitions in the optical absorption
spectra of the ions are correlated to the structural changes in the
host glass matrix. The shift of the hypersensitive bands of Sm3+
ions (6H5/2 ? 6F1/2; 1516 nm) shows that the covalency of the
REO bond increases with increase of Bi2O3 content, due to the increased interaction between rare earth ions and NBOs. The variation of intensity parameters X2 with Bi2O3 content for Sm3+
implies that the As,p plays an important role in determining the
intensity. X6 mainly depends on nephlauxetic effect [N(s, t)] for
the Sm3+ ions. The radiative emission probability for Sm3+ ions in
zinc bismuth borate is high as in lead uoro-borate glasses, which
have been successfully used as laser host material. This indicates
that the zinc bismuth borate glasses can act as potential laser host
material.

344

A. Agarwal et al. / Optical Materials 32 (2009) 339344

Acknowledgement
Authors are thankful to CSIR, New Delhi, for providing nancial
support.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]

J.L. Adam, A.D. Docq, J. Lucas, J. Solid State Chem. 75 (1988) 403.
R. Reisfeld, A. Bornstein, L. Boehm, J. Solid State Chem. 14 (1975) 14.
M. Canalejo, R. Cases, R. Alcala, Phys. Chem. Glass. 29 (1988) 187.
L. Boehm, R. Reisfeld, N. Spector, J. Solid State Chem. 28 (1979) 75.
S. Schweizer, L.W. Hobbs, M. Secu, J. Spaeth, A. Edgar, G.V.M. Williams, Appl.
Phys. Lett. 83 (2003) 449.
P.E.-A. Mobert, E. Heumann, G. Huber, B.H.T. Chai, Appl. Phys. Lett. 73 (1998)
139.
D.A. Turnbull, S.Q. Gu, S.G. Bishop, J. Appl. Phys. 80 (1996) 2436.
T. Tsuboi, Eur. Phys. J. Appl. Phys. 26 (2004) 95.
D. Ruter, W. Bauhofer, Appl. Phys. Lett. 69 (1996) 892.
N. Soga, K. Hirao, M. Yoshimoto, H. Yamamoto, J. Appl. Phys. 63 (1988) 4451.
S.M. Kaczmarek, Opt. Mater. 19 (2002) 189.
Z. Pan, S.H. Morgan, K. Dyer, A. Ueda, H. Liu, J. Appl. Phys. 79 (1996) 8906.
S. Tanabe, X. Feng, T. Hanada, Opt. Lett. 25 (2000) 817.
B.R. Judd, Phys. Rev. 127 (1962) 750.
G.S. Ofelt, J. Chem. Phys. 37 (1962) 511.
J.C. Joshi, J. Joshi, R. Belwal, B.C. Joshi, N.C. Pandey, J. Phys. Chem. Solids 39
(1978) 581.
K.K. Mahato, D.K. Rai, S.B. Rai, Solid State Commun. 108 (1998) 671.

[18] C.V. Reddy, Y.N. Ahammed, R.R. Reddy, T.V.R. Rao, J. Phys. Chem. Solids 59
(1998) 337.
[19] C.K. Jayasankar, E. Rukmini, Opt. Mater. 8 (1997) 93.
[20] M.B. Saisudha, J. Ramakrishna, Phys. Rev. B 53 (1996) 6186.
[21] W.T. Carnall, P.R. Fields, K. Rajnak, J. Chem. Phys. 49 (1968) 4424.
[22] K. Binnemans, R.V. Deun, C. Gorller-Walrand, J.L. Adam, J. Alloys Compds. 275
(1998) 455.
[23] R.V. Deun, K. Binnemans, C. Gorller-Walrand, J.-L. Adam, SPIE 3622 (1999) 175.
[24] R. Cases, M.A. Chamarro, J. Solid State Chem. 90 (1991) 313.
[25] C.K. Jorgensen, B.R. Judd, Mol. Phys. 8 (1964) 281.
[26] C.K. Jorgensen, R. Reisfeld, J. Less Common Met. 93 (1983) 107.
[27] Y. Nageno, H. Takebe, K. Morinaga, J. Am. Ceram. Soc. 76 (1993) 3081.
[28] C.K. Jorgensen, Modern Aspects of Ligand Field Theory, North-Holland,
Amsterdam, 1971.
[29] C.K. Jorgensen, Oxidation Numbers and Oxidation States, Springer, New York,
1969.
[30] C.K. Jorgensen, Prog. Inorg. Chem. 4 (1962) 73.
[31] B.N. Meera, Ph.D. Thesis, Indian Institute of Science, Bangalore, 1992.
[32] R. Reisfeld, Y. Eckstein, J. Solid State Chem. 5 (1972) 174.
[33] A. Vegas, F.H. Cano, S. Garcia-Blanco, J. Solid State Chem. 17 (1976) 151.
[34] T. Izumitani, H. Toratani, H. Kuroda, J. Non-Cryst. Solids 47 (1982) 87.
[35] M.J. Weber, L.A. Boatner, B.C. Sales, J. Non-Cryst. Solids 74 (1985) 167.
[36] M.J. Weber, R.A. Saroyan, R.C. Ropp, J. Non-Cryst. Solids 44 (1981) 137.
[37] S. Tanabe, T. Ohyagi, N. Soga, T. Hanada, Phys. Rev. B 46 (1992) 3305.
[38] D.L. Dexter, in: F. Seitz, D. Turubull (Eds.), Solid State Physics, vol. 6, Academic
Press, New York, 1958, p. 353.
[39] M.J. Weber, J.D. Myers, D.H. Blackburn, J. Appl. Phys. 52 (1981) 2944.
[40] A.G.S. Filho, J.M. Filho, F.E.A. Melo, M.C.C. Custodio, R. Lebullenger, A.C.
Hernandes, J. Phys. Chem. Solids 61 (2000) 1535.

You might also like