Você está na página 1de 19

Renewable and Sustainable Energy Reviews 51 (2015) 566–584

Contents lists available at ScienceDirect

Renewable and Sustainable Energy Reviews


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

A review on optical and photoluminescence studies


of RE3 þ (RE ¼Sm, Dy, Eu, Tb and Nd) ions doped LCZSFB glasses
C. Madhukar Reddy a, B. Deva Prasad Raju b,n, N. John Sushma c, N.S. Dhoble d, S.J. Dhoble e
a
Department of Physics, AP Model School, Yerravaripalem 517 194, India
b
Department of Future Studies, Sri Venkateswara University, Tirupati 517 502, India
c
Department of Biotechnology, Sri Padmavathi Mahila University, Tirupati 517 502, India
d
Department of Chemistry, Sevadal Mahila Mahavidhyalaya, Nagpur 440 009, India
e
Department of Physics, R.T.M. Nagpur University, Nagpur 440 033, India

art ic l e i nf o a b s t r a c t

Article history: Trivalent rare earth ions show interesting optical properties and such properties have high technological
Received 27 February 2014 applications of rare earth doped materials such as energy saving lighting devices, optical displays, optical
Received in revised form fibers, amplifiers and lasers. Among these materials rare earth ions doped glasses are of great important
5 May 2015
to optoelectronics and are widely used in fiber amplifiers and solid state high power lasers for
Accepted 1 June 2015
telecommunications and light emitting diodes. Near-infrared luminescence for high power lasers is a
current requirement in modern fiber optic telecommunication network and emphasis has to put in
Keywords: recent advances of NIR emitting materials for amplifiers, fiber lasers and waveguides. We make an effort
Glasses to satisfy the above needs by preparing the LCZSFB glasses doped with Sm3 þ , Dy3 þ , Eu3 þ , Tb3 þ and
Rare earth ions
Nd3 þ rare earth ions. In this review, optical properties of LCZSFB glasses doped with Sm3 þ , Dy3 þ , Eu3 þ ,
Optical absorption
Tb3 þ and Nd3 þ rare earth ions and the current status of their applications is given.
Photoluminescence
Radiative properties & 2015 Elsevier Ltd. All rights reserved.
Lasers

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566
2. Research scheme for RE doped laser glasses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
2.1. Spectroscopic techniques for optical materials assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
2.1.1. Absorption measurements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
2.1.2. Luminescence measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569
2.1.3. Excited state decay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569
2.1.4. Non-radiative relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 570
3. Sm3 þ ions doped LCZSFB glass for high gain laser applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
4. Dy3 þ ions doped LCZSFB glass for simulation of white light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573
5. Eu3 þ ions doped LCZSFB glass for visible red lasers and display devices. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 576
6. Tb3 þ ions doped LCZSFB glass for green fiber lasers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 579
7. Nd3 þ ions doped LCZSFB glass for 1.06 mm laser applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 581
8. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 583
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 583
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 583

1. Introduction

n
Corresponding author at: Department of Future Studies, Sri Venkateswara
The field of luminescence from rare-earth ions has been one of
University, Tirupati 517 502, India. Tel.: þ91 94402 81769. steady growth during the past decade, principally due to the ever
E-mail address: drdevaprasadraju@gmail.com (B. Deva Prasad Raju). increasing demand for optical sources and amplifiers operating at

http://dx.doi.org/10.1016/j.rser.2015.06.025
1364-0321/& 2015 Elsevier Ltd. All rights reserved.
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 567

wavelengths compatible with fiber communications technology. the host glass for incorporation of rare earth ions and strongly
The 4f–4f electronic transitions of rare earth (RE) ions play an influenced by the presence of highly polarizable Pb2 þ ions owing to
important role in the applications like optical fiber amplifiers, solid the strong and directional nature of the Pb–O bond [10–13].
state lasers, planar waveguides and compact micro chip lasers Rare-earth elements have many energy levels and some com-
[1–4]. Glasses have been known for a long time as a suitable host binations of emissive rare-earth elements and host materials relax
for REs and have been widely used for the production of solid state the populations at the metastable state through a non-radiative
lasers. Glasses doped with rare earth ions have been widely relaxation process. The intensity of non-radiative transitions
examined and possess many advantages due to their laser action strongly depends on the phonon energy of the host crystal.
in visible and near infrared (NIR) regions that make them very Fluoride based glasses have been well established as the right
considerable materials to be used as laser media and are currently choice for laser emission in the visible as well as mid-infrared
best substitute to single crystals for all solid state laser emission regions due to their low phonon energies [15]. The most important
[5]. Generally, the optical homogeneity of glassy matrices make characteristic is that their phonon energies are lower than those
available RE ions to exhibit diverse latent laser transitions. Spec- for the oxide materials. Indeed, the phonon energy of fluoride
troscopic study of RE ions in glasses suggest information with glass is one-third or half that of oxide glass [14]. In fluoride
consider to transition probabilities, lifetimes, branching ratios of systems, huge amount of rare-earth ions can be introduced in
the excited states, which are vital in the design and growth of the matrix and they can be used for special optical fibers and fiber
diverse electro-optic and optical devices. Much research in the lasers. The low phonon energies of these systems result large
1960s was concerned with the improvement of solid state lasers luminescence efficiency, due to the small non-radiative rates of
utilize the luminescence of RE ions in glasses and crystals. The Nd: rare-earth ion excited states. In quite recent years glassy materials
YAG laser is probably the most central triumph from this period, based on ZnO have attracted the scientific community because of
and there is persistent interest in novel rare earth doped materials its optical, electrical and magnetic properties in combination with
for lasers. All the rare earth ions (Pr3 þ , Nd3 þ , Sm3 þ , Eu3 þ , Gd3 þ , its non toxicity, non hygroscopic nature. Its direct wide band gap,
Tb3 þ , Dy3 þ , Ho3 þ , Er3 þ , and Tm3 þ ) have their respective applica- large exciting binding energy and intrinsic emitting property
tions in the production of optical devices to extend advanced makes them as promising candidates for the development of
lasers and optical amplifiers to optoelectronic and optical com- optoelectronic devices, solar energy converters, ultra violet emit-
munication applications in the field of RE luminescence. More ting lasers and gas sensors [16–20].
recently, however, optoelectronics has emerged as the principal In the process of probing of possible glass compositions, by
area of research into RE luminescence, and the current article bearing in mind the significance of suitable host glass matrix for
therefore focus on the different ways in which RE luminescence various applications in the RE luminescence, the lead calcium zinc
has been exploited in this field. sodium fluoroborate (LCZSFB) [20PbO þ5CaO þ5ZnO þ 10NaFþ
As the fluorescence properties of RE3 þ ions depend on the host 59B2O3 þ 1RE2O3, (RE¼ Sm, Dy, Eu, Tb and Nd)] glass system has
environment, an immense quantity of research has to be carried out been chosen by doping Sm3 þ , Dy3 þ , Eu3 þ , Tb3 þ and Nd3 þ for the
to expand new glass materials containing RE3 þ ions with high present investigation. The paper reviews some studies on the
quantum efficiency. The possibility of controlling the physical spectroscopic properties, radiative and non-radiative processes,
properties of glasses such as refractive index and density by the and site structure of RE ions in glasses at each wavelength. We
variation of glass composition suggests the probability of usage of look forward to that this review will give confidence to young
different chemicals according to the requirements of respective investigators to join the emergent area who is already exploiting
applications. Enhancement in the quantum efficiency of the lumi- the interesting properties of lanthanides for present an improved
nescent levels of RE ions can be achieved by selecting an appro- living to the people of planet earth.
priate host material and by adjusting the local environment
surrounding them. Such modifications are frequently achieved by
breaking up the structure of atoms surrounding a rare earth ion 2. Research scheme for RE doped laser glasses
with other termed as network modifiers. Discussing the role of
modifiers requires an understanding of the basic structure of A laser works on the principle of light amplification by
different glass network formers. Oxide glasses such as borate, stimulated emission of radiation. It requires a pump to create a
phosphate, silicate, and telluride have proven to be the proper host population inversion in the active medium. When an incoming
materials for the advancement of opto-electronic components. signal photon stimulates an excited electron, the electron relaxes
Among the oxide glass hosts, borate glasses have concerned much back into a lower energy state and emits a second signal photon
concentration due to their high transparency, lower melting point with the same phase as the incoming photon. This process of
with good transparency, high chemical durability, thermal stability stimulated emission amplifies the signal. The success of a laser
and good rare earth ion solubility [6]. Usually, glass with B2O3 alone depends on the suitable energy level structure of RE ions can cause
possesses high phonon energies (1300 cm  1) and cannot suppress amplification to emit a photon with precisely the wavelength at
non-radiative decay process and hence rare earth ion emissions are which the network operates, with high efficiency. In general, RE
strongly reduced. Host glasses with low phonon energy provide less ions have an ideal energy level structure works at different
non-radiative relaxation rates and high quantum efficiencies [7,8]. pumping wavelengths available, and are used as a laser actives
The incorporation of heavy metal oxides such as PbO, PbF2, and when used in glasses and to design the novel optical devices. To
Bi2O3 in borate glass leads to decrease its phonon energy. However, understand the quantitative optical phenomena of rare-earth ions
a considerable enhancement in luminescence intensity and lifetime in glasses, it is important to evaluate radiative and non-radiative
has been observed for samples with PbO and/or PbF2. The PbO- decay process of related 4f levels. The induced electric dipole
based borate glasses doped with rare earth ions are mainly transitions are parameterized by the Judd–Ofelt theory [21–23].
fascinating due to combine luminescent properties of dopant ions The intensity of induced electric dipole transitions can be
with potential optical nonlinearity of borate group and promising described in terms of three phenomenological intensity para-
hosts to investigate the influence of chemical environment and the meters Ωλ (λ¼ 2, 4, 6). The Judd–Ofelt theory is usually adopted
structural variety of ligand groups on the spectroscopic properties to obtain the radiative transition probabilities including emission
of rare earth ions [9]. The lead borate glasses are optically by utilizing the data of absorption cross-sections of several f–f
transparent forming wide compositional range of PbO content in electric-dipole lines. The physical and chemical implement of
568 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

Fig. 2.1. Research scheme for efficient laser materials.

three Ωλ parameters (λ ¼2, 4, 6) are becoming clearer by combin- competitive fabrication of innovative devices. The fundamentals of
ing the information of the local ligand field of doped ions by other optical spectroscopy, both theoretical and experimental, can be
spectroscopic techniques and give the information about the rare- found in several textbooks and scientific papers [29–31]. The
earth environment in glass; bond covalency and symmetry spectroscopic properties, namely emission quantum efficiency,
[24,25]. The non-radiative decay rate can be evaluated experimen- lifetime of the excited electronic states, dynamical processes,
tally by combining the lifetime measurement, which includes etc., of systems activated by luminescent ions are investigated by
contributions of multiphonon decay, energy transfer such as cross luminescence spectroscopy.
relaxation, cooperative upconversion, etc. In order to distinguish
each contribution, systematic studies on composition and concen-
tration dependence of the decay rates are necessary. The above- 2.1.1. Absorption measurements
mentioned research scheme is shown in Fig. 2.1. When a RE3 þ ion is embedded in a solid matrix, the effect of
ligand environment is minimum on the 4f shell because it is
2.1. Spectroscopic techniques for optical materials assessment effectively shielded by the closed 4s and 5p shells. Although weak,
this perturbation is responsible for the rich electronic spectra and
The proposition of this part is to recollect some basics con- provides detailed fingerprint information of the surrounding
cerning the use of spectroscopic techniques such as absorption, arrangement of atoms and their interactions with the 4f electrons.
luminescence, and spectroscopy for the study of optical materials. Absorption spectra are fundamental to determine the factors
The significance of absorption and luminescence spectroscopy is governing several properties of optical materials, such as trans-
requiring in photonics. These two techniques, in particular, allow mission losses, absorption cross-sections and refractive index. The
one to determine the spectroscopic properties of the optical intensities of the absorption bands can be expressed in terms of
species embedded in a matrix [26–28]. The spectroscopic and oscillator strengths. The calculated oscillator strength of an f–f
optical techniques are currently used for optimization and pre- transition can be evaluated using the Judd–Ofelt theory. According
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 569

to this theory, the calculated oscillator strength of an induced 2.1.2. Luminescence measurements
electric-dipole transition from the ground state Ψ J, to an excited Rare-earth-doped glasses are used in a large number of optical
state Ψ 0 J 0 , is given by devices because of the large number of absorption and emission
 2 bands available using the various rare-earth elements [34,35]. In
8 π 2 mcν n2 þ 2 X  2 material systems activated by ions, the spectroscopic properties,
f cal ¼ Ωλ Ψ J:U λ :Ψ0 J 0 ð1Þ
3 hð2 J þ 1Þ 9n λ ¼ 2;4;6 such as emission quantum efficiency, lifetime of the excited
electronic states, and dynamical processes such as non-radiative
where ‘n’ is refractive
 index2of the medium, ν is the energy of the relaxation mechanism, upconversion and cooperative processes
transition in cm  1, n2 þ 2 =9n is the Lorentz local field correc- [36], may be well investigated by luminescence spectroscopy. The
tion and accounts for dipole–dipole transition. J is the total angular J–O parameters along with refractive index (n) are used to predict
momentum of the ground state, Ωλ (λ¼ 2, 4 and 6) are J–O the radiative properties of excited states of RE3 þ ion. The radiative
2
intensity parameters and :U λ : are the doubly reduced matrix transition probability (AR ) for a transition Ψ J-Ψ 0 J 0 can be calcu-
elements of the unit tensor operator [32] evaluated in the inter- lated using the equation [23,37]
mediate coupling scheme for a transition Ψ J-Ψ 0 J 0 . The experi- " #
mental oscillator strength of an absorption (f exp ) is directly 0 0 64 π 4 ν3 nðn2 þ 2Þ2 0 0 3 0 0
AR ðΨ J; Ψ J Þ ¼ Sed ðΨ J; Ψ J Þ þ n Smd ðΨ J; Ψ J Þ
proportional to the area under the absorption curve and is 3h ð2J þ 1Þ 9
expressed as [23,33] ð5Þ
 Z Z
2:303 mc2 9
The total radiative transition probability (AT ) of an excited state
f exp ¼ ε ð νÞdν ¼ 4:318 10 εðνÞdν ð2Þ is the sum of the AR ðΨ J; Ψ 0 J 0 ) terms calculated over all the
Nπe2
terminal states
where m and e are mass and charge of an electron, c is the velocity X
of light, N is the Avogadro’s number, εðνÞ is the molar absorption AT ðΨ JÞ ¼ AR ðΨ J; Ψ 0 J 0 Þ ð6Þ
Ψ 0 J0
coefficient of absorption band corresponding to the energy ν
(cm  1) and dν is the half bandwidth. From the Beer–Lambert AT is related to the radiative lifetime ðτR ) of an excited state by
law, the molar absorption coefficient is given by
1
  τR ðΨ JÞ ¼ τcal ðΨ JÞ ¼ ð7Þ
1 I0 AT ðΨ JÞ
εðνÞ ¼ log ð3Þ
Cl I Strong emission probabilities and more transitions from a level
where C is the rare-earth ion concentration (in mol/l), l is the lead to faster decay and shorter lifetimes. The theoretical radiative
  lifetime τR ðΨ J Þ, calculated from the J  O intensity parameters (Ωλ ),
thickness of the glass sample and log I 0 =I is the optical density.
A least-squares fit method is then used for Eq. (1) to determine can be compared with the measured lifetimes, τm ðΨ J Þ. The dis-
Ωλ parameters, which gives the best fit between experimental and crepancy between predicted and experimental lifetimes is clearly
calculated oscillator strengths. The calculated oscillator strengths due to the manifestation of non-radiative process (W NR ) either by
(f cal ), are then obtained using Eq. (1) and Ωλ . multiphonon relaxation rate (W MPR ) or energy transfer rate (W ET ).
A measure of the accuracy of the fit between the experimental From the measured and calculated lifetimes, the quantum effi-
and calculated oscillator strengths is given by the root mean ciency (η) is estimated by the expression
square (rms) deviation τm AR
η¼ ¼ ð8Þ
2 2 31=2 τR AR þ W NR
P
6 f exp  f cal 7 The branching ratios can be used to predict the relative
δrms ¼ 4 5 ð4Þ
N intensities of all emission lines originating from a given excited
state Ψ J. The experimental branching ratios can be found from the
relative areas of the emission bands. The branching ratio ðβR )
where N is the number of levels included in the fit. The small rms
corresponding to the emission from an excited level (Ψ J) to its
deviation indicates a good fit between experimental and calcu-
lower level (Ψ 0 J 0 ) is given by
lated oscillator strengths.
AR ðΨ J; Ψ 0 J 0 Þ
βR ðΨ J; Ψ 0 J 0 Þ ¼ ð9Þ
AT ðΨ JÞ
The peak stimulated emission cross-section, σ e ðΨ J; Ψ 0 J 0 )
between the states Ψ J and Ψ 0 J 0 having a probability of
AR ðΨ J; Ψ 0 J 0 ) can be expressed as
λ4p
σ e ðΨ J; Ψ 0 J 0 Þ ¼ AR ðΨ J; Ψ 0 J 0 Þ ð10Þ
8 πcn2 Δλp
where λp is the transition peak wavelength and Δλp is its effective
linewidth found by dividing the area of the emission band by its
average height. Good laser transitions are characterized by large
cross-sections for stimulated emission.

2.1.3. Excited state decay


The study of excitation and de-excitation of RE ions due to intra
4f electron transitions provides a deeper insight into mechanisms
involved in excitation. A typical time-resolved intensity spectrum
is shown in Fig. 2.2. The curve is described by integral solutions to
appropriate rate equations which account for the possible excita-
Fig. 2.2. A typical time-resolved excitation and de-excitation curve of RE3 þ ions. tion and de-excitation mechanism. An excited RE3 þ ion may relax
570 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

to the initial ground state through radiative transition or phonon


emission or by transferring its excess energy to a nearby RE3 þ ion
or by a combination of relaxations.
At very low concentrations of dopant ions, when the interac-
tion between the optically active RE3 þ ions is negligible, the
fluorescence decay curves can be fitted to a single exponential
function. A perfect single exponential decay indicates that the
energy transfer between luminescent ions is not dominant and the
lifetime of the excited level can be simply determined by finding
the first e-folding times. The fluorescence intensity as a function of
time can be described by the following equation.
 
I ðt Þ ¼ I 0 exp  t=τ ð11Þ
where I0 is the fluorescence intensity when t¼0. The value τ
represents the lifetime of the excited state and is reciprocal to the
probability of a spontaneous emission from the excited state to the
ground state. A logarithmic plot of the intensity versus time
should help to determine lifetime. It is evident from Fig. 2.2 that
after time τ the intensity of the excited state has decreased to 1/e.
Fig. 3.2. Near-infrared absorption spectrum of LCZSFBSm10 glass.

2.1.4. Non-radiative relaxation


An excited RE3 þ ion produced by absorption of a photon, also
relaxes to the lower energy state via a non-radiative process. Any
deviation from a single exponential decay is unambiguously
attributable to non-radiative decay due to energy transfer to the
4f electron system after the excitation source is turned off,
provided only a single local environment is present. Typical
examples are the energy transfer between different RE3 þ ions
such as cross-relaxation and excitation migration between the
same types of RE3 þ ions. Another example would be the popula-
tion of a multiplet by another multiplet, which is energetically
located above, but within the same ion.
If the measured lifetime of the emitting state is denoted by τm,
the total decay rate (1/τm) is the sum of radiative (AR) and non-
radiative (WNR) decay rates.
1
¼ AR þ W NR ð12Þ
τm
The non-radiative decay rates play an important role on the
quenching of lifetime. There are mainly four non-radiative decay
processes, contributing to the reduction of measured lifetime of Fig. 3.3. Excitation spectrum of LCZSFBSm10 glass.
the emitting level [38].
W NR ¼ W MPR þW ET þ W CQ þ W OH ð13Þ
where WMPR, WET, WCQ and WOH denote the non-radiative decay
Table 3.1
rates corresponding to the multi-phonon relaxation, energy
Experimental and calculated oscillator strengths of LCZSFBSm10 glass.

Transition (6H5/2-) Energy (cm  1) f exp 10  6 f cal 10  6

6
F1/2 6,293 0.57 1.03
6
H15/2 6,540 1.74 0.04
6
F3/2 6,757 3.74 3.01
6
F5/2 7,252 4.77 5.08
6
F7/2 8,130 7.32 7.53
6
F9/2 9,268 4.97 4.78
6
F11/2 10,582 1.26 0.76
δrms ¼ 7 0.7610  6

transfer between donor to donor or donor to acceptor, concentra-


tion quenching and hydroxyl (OH) groups, respectively.
From the measured and calculated lifetimes, the quantum
efficiency (η) is estimated by the expression

τm AR
η¼ ¼ ð14Þ
Fig. 3.1. Visible absorption spectrum of LCZSFBSm10 glass. τR AR þ W NR
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 571

Table 3.2
Comparison of J–O intensity parameters, root mean square deviation for LCZSFBSm10 glass with different glass hosts.

Glass system Ω2 (10  20 cm2) Ω4 (10  20 cm2) Ω6 (10  20 cm2) Trend δrms (10  6)

LCZSFBSm10 3.29 9.16 5.28 Ω4 4 Ω6 4Ω2 7 0.76


PKFBASm10 [39] 1.50 3.75 1.89 Ω4 4Ω6 4 Ω2 7 0.39
N4BS [40] 2.52 3.62 1.14 Ω4 4Ω2 4 Ω6 7 0.49
N5BS [40] 2.86 4.13 1.92 Ω4 4Ω2 4 Ω6 7 0.78
PKMASm10 [41] 6.83 2.97 2.03 Ω2 4Ω4 4 Ω6 7 0.63
TZKCSm10 [42] 2.48 2.99 1.82 Ω4 4Ω2 4 Ω6 7 0.30
PKAZLFSm10 [43] 4.50 7.40 3.93 Ω4 4Ω2 4 Ω6 7 0.46
PKAlCaFSm10 [44] 0.41 5.65 4.57 Ω4 4Ω6 4 Ω2 7 0.62

3. Sm3 þ ions doped LCZSFB glass for high gain laser


applications

The room temperature absorption spectra of LCZSFBSm10 glass


consists of several inhomogeneously broadened bands assigned to
f-f transitions from the ground 6H5/2 ground state to various
excited states of Sm3 þ ions in the host glass. Optical absorption
spectra of LCZSFBSm10 glass in UV–vis and NIR regions are shown in
Figs. 3.1 and 3.2 respectively. The characteristic absorption bands are
assigned to 6H5/2-6F1/2, 6H15/2, 6F3/2, 6F5/2, 6F7/2, 6F9/2, 6F11/2, 4G5/2,
4
F3/2, 4G7/2, 4M15/2, 4I11/2 transitions at 6293, 6540, 6757, 7252, 8130,
9268, 10,582, 17,762, 18,939, 20,000, 21,052 and 21,645 cm  1
respectively, that fall in the UV, Vis and NIR regions. The observed
absorption transitions are similar to other Sm3 þ ions doped glasses
[39–44], except in peak positions and intensities due to variations in
the glass compositions. The excitation spectrum recorded at 602 nm
is most useful and helpful to create and assign few more levels at
higher energy side shown in Fig. 3.3, which are not observed in the
absorption spectrum. In general, the oscillator strength is expressed Fig. 3.4. Emission spectrum of LCZSFBSm10 glass. (For interpretation of the
in terms of intensity of absorption bands. The experimental oscilla- references to color in this figure legend, the reader is referred to the web version
of this article.)
tor strengths of the absorption bands are estimated by measuring
the area under the absorption bands. These results have been used
to calculate the Judd–Ofelt (JO) parameters Ωλ (λ ¼ 2, 4 and 6) environment around the Sm3 þ ion sites, whereas Ω4 and Ω6 are
[21,22], along with the calculated oscillator strengths using doubly related to the viscosity and rigidity of the host medium in which the
reduced matrix elements by least square fit approximation. The ions are situated. Moreover, among the three Ωλ parameters the
calculated oscillator strengths can be expressed as a function of the value of Ω2 is relatively larger for oxide glasses, smaller for fluoride
reduced matrix elements of the unit tensor U ðλÞ that are almost host glasses while intermediate values are noticed for oxyfluoride glasses.
independent and insensitive to the local environment. Reasonably This implies that RE3 þ –O covalency decreases when pure oxide
small root mean square deviation (δrms ) of 7 0.76 10  6 between glasses are modified with fluorine content [38]. Thus in the present
experimental (f exp ) and calculated (f cal ) oscillator strengths indicates glass system, lower magnitude of Ω2 shows a decrease of covalency
the goodness of fit. The assignment of bands, their peak positions, of the Sm3 þ ion environment than the other reported [41,43] Sm3 þ
experimental and calculated oscillator strengths are presented in ions doped glass hosts. The lesser the value of Ω2 , the more centro
Table 3.1. A comparison of δrms value of the present glass with other symmetrical the ion site and the more ionic its chemical bond with
reported glasses [39–44] is shown in Table 3.2. The oscillator the ligands. The higher magnitude of Ω2 parameter is an indicator
strengths of the absorption bands of LCZSFBSm10 glass are higher for the Sm3 þ ion sites are highly asymmetric and Sm3 þ –O2 bonds
than those of other reported Sm3 þ ions doped glasses [39,41–44] posses higher covalency than the other reported [39,40,42,44] Sm3 þ
except in the N5BS [40] glass. The higher magnitude of oscillator doped glass hosts. The value of Ω4 in LCZSFBSm10 glass is higher
strengths in the present glass host may be due to the higher value of compared to those of Sm3 þ ions doped [39,40,42,44] glasses and
the asymmetric component of the electric field induced by the lesser compared to those of [41,43] glasses. The higher magnitude of
ligand environment surrounding the Sm3 þ ion in LCZSFBSm10 glass. Ω4 indicates that the present glass posses more rigidity.
Table 3.2 compares the JO parameters obtained in the present work LCZSFBSm10 glass emits bright reddish-orange luminescence under
with those of the other Sm3 þ doped glass systems [39–44]. JO 402 nm excitation wavelength. The emission spectra of LCZSFBSm10
analysis yields different sets of intensity parameters in different glass is shown in Fig. 3.4 and it consists of potential green yellow and
hosts as compared in Table 3.2, since they depend on the intensities reddish-orange and red emission bands at 565, 602, 648 and 710 nm
of energy levels in that particular glass host. As seen from Table 3.2, which correspond to the 4G5/2-6H5/2, 6H7/2, 6H9/2 and 6H11/2 transitions,
it is observed that the magnitude of trends of Ωλ parameters for respectively. Among these four, 4G5/2-6H5/2, 7/2 transitions contain
LCZSFBSm10 glass is found to be Ω4 4Ω6 4Ω2 , where as a mixed magnetic dipole contributions obeying the selection rules ΔJ¼0, 71
trend is observed in the other Sm3 þ doped glasses [39–44]. These [46], while the other transitions 4G5/2-6H9/2, 11/2 are purely electric
trends may be attributed to the presence of assimilar sites around dipole. In order to evaluate the intensity of probable lasing transition,
the Sm3 þ ions depending upon the host glass composition. Accord- from the areas under the emission bands of Sm3þ ion the effective
ing to Jorgensen and Reisfeld [45], Ω2 parameter is related to glass bandwidth (Δλp ), radiative transition probabilities (AR ), peak stimulated
structure that depends on RE–O covalency in the vicinity of RE3 þ emission cross section (σ e ), optical gain parameter (σ e  τm ), experi-
ions, because Ω2 is connected to the asymmetry of the local mental and calculated branching ratios (βm ; βR ) are determined and
572 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

Table 3.3
Radiative properties of LCZSFBSm10 glass.

Transition 4G5/2- λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  22 cm2) σ e  Δλp (10  28 cm3) σ e  τ m (10  25 cm2 s)

6
H11/2 710 20.60 72.59 0.12 0.06 4.70 9.70 4.42
6
H9/2 648 19.32 173.50 0.29 0.19 8.30 16.0 7.80
6
H7/2 602 14.49 264.88 0.44 0.59 12.60 18.3 11.85
6
H5/2 565 13.85 31.02 0.05 0.21 1.20 1.70 1.13

Table 3.4
Comparison of radiative properties of 4G5/2-6H7/2 emission transition in LCZSFBSm10 glass with different glass hosts.

Glass λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  22 cm2) σ e  Δλp (10  28 cm3) σ e  τ m (10  25 cm2 s)

LCZSFBSm10 602 14.49 264.88 0.44 0.59 12.60 18.3 11.85


PKFBASm10 [39] 597 12 99 0.42 0.56 5.92 7.1 14.2
N5BS [40] 597 17.5 246.47 0.44 0.6 8.74 15.3 19.05
N4BS [40] 600 15 207.57 0.43 0.54 8.78 13.17 21.86
N3BS [40] 598 22.5 145.86 0.39 0.57 4.09 9.2 9.24
PKMASm10 [41] 598 11.2 88 0.32 0.47 5.80 6.5 10.4
TZKCSm10 [42] 603 15 232.76 0.38 0.56 6.68 10.02 5.14
PKAZLFSm10[43] 601 14.63 189 0.45 0.56 9.53 13.94 16.1
PKAlCaFSm10 [44] 602 13.78 174 0.55 0.56 9.30 12.8 1.45

Fig. 3.5. Energy level diagram of LCZSFBSm10 glass showing cross relaxation channels. Fig. 3.6. Decay curve of the 4G5/2 level of LCZSFBSm10 glass. Inset shows the
logarithmic intensity plot with time
are presented in Table 3.3. A comparison of λp , Δλp , AR , βm ; βR and σ e
for 4G5/2-6H7/2 transition of Sm3þ ions doped glasses [39–44] shown highly useful in under sea communication, optical data storage, color
in Table 3.4. From the values of radiative transition probabilities of displays and medical diagnostics. The partial energy level diagram of
Table 3.4, it is noted that 4G5/2-6H7/2 has highest radiative transition Sm3þ ions in LCZSFBSm10 glass showing observed emission transitions
rates compared to other transitions. Hence this transition is very useful under 402 nm excitation wavelength as well as the possible cross
for laser emission. The predicted branching ratios are found to be high relaxation channels are shown in Fig. 3.5. When Sm3þ ions are excited
for those transitions having maximum AR values. The levels having the to any other levels above the 4G5/2 metastable state, then there is a
relatively large values of AR , βR and energy gap to the next lower level quick non-radiative relaxation to this fluorescent level due to the small
may exhibit laser action. The luminescence branching ratio is another energy gap between them and consequently emission takes place from
4
important parameter that characterizes the lasing power of a transition G5/2 level to lower levels. The 4G5/2 level posses purely radiative
and it is well established that an emission transition having the relaxation as this level has sufficient energy gap of  7250 cm  1 with
luminescence βR greater than 50% is considered to have more potential respect to next lower level 6F11/2. The radiative relaxation of an excited
for laser emission [47]. The large stimulated emission cross section is state to all its lower levels depends upon AR values. The values of AR
very attractive feature for low threshold, high gain laser applications, depend upon JO parameters and energy gap between initial level and
which are utilized to obtain continuous wave laser action. Among the terminal level.
four observed transitions, the peak stimulated emission cross-section is Fig. 3.6 represents the experimental decay curve of
found to be high for the 4G5/2-6H7/2 transition. Also it is worth noting LCZSFBSm10 glass. The experimental lifetime of fluorescent 4G5/2
that the present glass exhibits higher stimulated emission-cross section level has been determined by taking first e-folding times of decay
for 4G5/2-6H7/2 transition than any other reported [39–44] glasses curves [48,49]. The decay curve is found to be non-exponential
shown in Table 3.4. The optical gain is also one of the critical parameter that is may be due to non-radiative energy transfer (energy
to predict the amplification of the medium in which the RE3þ ions are transfer from donor (excited state ion) to acceptor (ground state
situated. From Tables 3.3 and 3.4, it is concluded that 4G5/2-6H7/2 ion)) through cross relaxation. The experimental lifetime (τm ) is
transition of LCZSFBSm10 glass has the higher value of σ e and βR which significantly smaller than the predicted lifetime (τR ) 1.66 ms
suggests that it might be used for high gain laser applications, which are obtained using the JO theory. This change is may be due to the
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 573

energy transfer through cross relaxation but not due to the multi Table 4.1
phonon relaxation as it is negligible in case of Sm3 þ , since the Experimental and calculated oscillator strengths of LCZSFBDy10 glass.
energy gap is very large ( 7250 cm  1) between next lower level
6 Transition (6H15/2-) Energy (cm  1) f exp 10  6 f cal 10  6
F11/2. The magnitude of energy transfer rate (W ET ) is found to be
462 s  1 for the 4G5/2 level of LCZSFBSm10 glass. The cross relaxa- 6
H11/2 5,945 1.93 2.56
tion channels in the LCZSFBsm10 glass may be estimated to be 6
F11/2 7,856 11.29 11.22
A: 4G5/2-6F11/2 (E ¼7180 cm  1)  6H5/2-6F5/2 (E ¼7252 cm  1),
6
F9/2 9,140 4.54 4.73
6
F7/2 11,173 4.90 4.34
B: 4G5/2-6F9/2 (E ¼8494 cm  1)  6H5/2-6F7/2 (E ¼8130 cm  1), 6
F5/2 12,500 3.65 2.16
C: 4G5/2-6F7/2 (E ¼9532 cm  1)  6H5/2-6F9/2 (E ¼9268 cm  1), 6
F3/2 13,298 0.99 0.40
D: 4G5/2-6F5/2 (E¼ 10,510 cm  1)  6H5/2-6F11/2 (E¼ 10,582 cm  1)
as the energy difference between these transitions is negligible. δrms ¼ 7 0.5510  6
The cross relaxation is due to the energy transfer from the Sm3 þ
ion in an excited 4G5/2 state, to a nearby Sm3 þ ion in the ground
6
H5/2 state. This transfer of cross relaxation occurs via 4G5/2-6F5/2 range 400–1800 nm contains 7 absorption bands originated from
transition in one ion and 6H5/2-6F11/2 transition on the other. This the 6H15/2 ground level to the various excited levels belonging to the
transfer leaves the first ion in the intermediate level of the 6F11/2 at 4f9 electronic configuration of the Dy3 þ ion, centered at 1682, 1273,
around 10,582 cm  1 and the second one in the 6F7/2 at around 1094, 895, 800, 752 and 453 nm corresponding to the 6H11/2, 6F11/2,
8130 cm  1 to which resonance occur with the 4G5/2-6F5/2 and 6
F9/2, 6F7/2, 6F5/2, 6F3/2 and 4I15/2 excited states and are similar to
4
G5/2-6F9/2 transitions at around 10,510 cm  1 and 8494 cm  1 other reported glass hosts [50–57] though there is a slight variation
respectively. After that both ions quickly decay non-radiatively to in band positions and intensities of the bands due to the change in
the ground state. The energy resonance of these transitions can be glass composition. The band identified at 453 nm (6H15/2-4I15/2)
clearly seen in Fig. 3.5. The quantum efficiency of the fluorescent and 473 nm (6H15/2-4F9/2) are very weak due to the strong
level is defined as the ratio of the number of photons absorbed is absorption of the LCZSFBDy10 glass host, which also results the
estimated to be 57% for LCZSFBSm10 glass and comparison of disappearance of some of the absorption bands in UV region. The
lifetimes (τm ; τR ), energy transfer rate (W ET ) and quantum effi- intensities of the absorption bands are expressed in terms of
ciency (η) with other glass hosts [39–44] is shown in Table 3.5. oscillator strengths (f exp ) 6H15/2-6F11/2 transition centered at
7856 cm  1 shows higher intensity compared to the other transi-
tions for the same ion and is noted as hypersensitive transition
  in
4. Dy3 þ ions doped LCZSFB glass for simulation of white light Dy3 þ that obeys the selection rules: jΔSj¼0, jΔLj r 2 and ΔJ  r 2 is
very sensitive to the host environment [58]. The JO theory has been
Fig. 4.1 shows the absorption spectrum of LCZSFBDy10 glass, applied to evaluate the calculated oscillator strengths (f cal ) and JO
which has been recorded at room temperature in the wavelength intensity parameters Ωλ (λ ¼2, 4 and 6) by a least square fit analysis.
The values of experimental energies, experimental and calculated
Table 3 5 oscillator strengths of LCZSFBDy10 glass are presented in Table 4.1.
Comparison of decay properties of LCZSFBSm10 glass with different glass hosts.
The hypersensitive transition is normally associated with the larger
2
Glass τ m (ms) τ R (ms) η ð%Þ W NR (s  1) value of :U λ : matrix elements and therefore the hypersensitivity is
related to JO intensity parameters [59]. So the hypersensitive
LCZSFBSm10 0.94 1.66 57 462 transition 6H15/2-6F11/2 has the large values of matrix elements
PKFBASm10 [39] 2.40 4.29 56 184
and oscillator strengths. Reasonably small δrms of 7 0.5510  6
N5BS [40] 2.18 2.85 76 109
N4BS [40] 2.49 3.38 74 105 indicates a good fit between experimental and calculated oscillator
N3BS [40] 2.26 2.99 75 108 strengths. Table 4.2 shows a comparison of JO intensity parameters
PKMASm10 [41] 1.8 3.14 57 237 and root mean square deviation of various reported Dy3 þ ions
TZKCSm10 [42] 0.77 1.63 47 694 doped glasses [50–57]. The trend of JO intensity parameters in the
PKAZLFSm10 [43] 1.69 2.12 80 120
PKAlCaFSm10 [44] 1.56 3.18 49 327
present LCZSFBDy10 glass is Ω2 4Ω6 4Ω4 , which is similar to the
trends observed in other reported systems [51,52,54,55,57]. Among
the JO intensity parameters, Ω2 is more sensitive to the local
structure of the RE ion, and in turn depends strongly on the
hypersensitive transition and the higher value of Ω2 is due to the
relatively higher value of Ω2 is due to the relatively higher value of
the oscillator strength of the hypersensitive transition. As can be
seen from Table 4.2, the highest value of Ω2 in the present glass
system experiences relatively higher covalence and lower asymme-
try than in the other reported [50,51,54,56,57] glass systems except
in the reported glass [52,53,55] systems.
To analyze the luminescence properties of LCZSFBDy10 glass
the excitation spectrum was recorded in the spectral region 300–
500 nm by monitoring the emission at 570 nm and is shown in
Fig. 4.2. The excitation bands centered at 352, 365, 386, 430, 453
and 473 nm are attributed to the 6H15/2-6P7/2, 4P9/2, 4I13/2, 4G11/2,
4
I15/2 and 4F9/2 transitions respectively. It is well known fact that
the wavelength corresponding to the prominent excitation band
can give intense emission. In the present investigation, the
excitation band centered at 453 nm is found to be more intense.
To study the white light emission under UV light, the lumines-
cence spectra were carried out by exciting the sample at 386 nm.
Fig. 4.1. Absorption spectrum of LCZSFBDy10 glass Fig. 4.3 shows the luminescence spectra of LCZSFBDy10 glass in
574 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

Table 4.2
Comparison of J–O intensity parameters, root mean square deviation for LCZSFBDy10 glass with different glass hosts.

Glass system Ω2 (10  20 cm2) Ω4 (10  20 cm2) Ω6 (10  20 cm2) Trend δrms (10  6)

LCZSFBDy10 11.25 2.45 5.16 Ω2 4Ω6 4Ω4 7 0.55


PKAZLFDy10 [50] 7.05 1.43 1.15 Ω2 4Ω4 4Ω6 7 0.22
ZnAlBiBDy10 [51] 4.66 1.61 2.71 Ω2 4Ω6 4Ω4 7 0.39
PbPKANDy10 [52] 11.74 2.64 2.86 Ω2 4Ω6 4Ω4 7 0.68
PKAZFDy10 [53] 14.11 3.07 1.95 Ω2 4Ω4 4Ω6 7 0.58
PTBDy10 [54] 7.75 2.31 2.70 Ω2 4Ω6 4Ω4 7 0.34
BNLDy10 [55] 14.53 3.05 4.99 Ω2 4Ω6 4Ω4 7 0.23
CFBDy10 [56] 5.98 2.33 2.33 Ω2 4Ω4 ¼ Ω6 7 0.17
LBEDy10 [57] 6.45 2.02 3.87 Ω2 4Ω6 4Ω4 7 0.19

often dominant [52,54,56], when the Dy3 þ ion is located in the


low symmetry local site without inversion center and the blue
emission (4F9/2-6H15/2) is stronger than the yellow one
[50,51,53,55,57] when it is located at a high symmetry local site
with an inversion center [60,61]. To understand the Dy3 þ lumi-
nescence of glasses under study, the JO theory has been applied to
determine the radiative properties such as spontaneous emission
transition probability (AR ), radiative lifetime (τR ), radiative branch-
ing ratio (βR ) for the observed transitions are calculated and from
the emission spectra the laser properties such as effective band-
width (Δλp ), experimental branching ratio (βm ), stimulated emis-
sion cross section σ e , σ e  τR , σ e  Δλp are calculated and are
presented in Table 4.3. Table 4.4 shows a comparison of radiative
properties with other reported glass hosts [50–57]. In general the
luminescence branching ratio is a critical parameter to the laser
designer, because it characterizes the possibility of attaining
stimulated emission from any specific transition. For the 4F9/
6
2- H13/2 transition, the value of β R (0.66) is found to be Z 0.50
Fig. 4.2. Excitation spectrum of LCZSFBDy10 glass
is more potential for laser emission and is varying from βm (0.37).
The variation in between βR and βm is attributed to the non-
radiative contributions from the 4F9/2 level of LCZSFBDy10 glass. At
386 nm excitation, the 4I13/2 level of Dy3 þ ion initially gets
populated, but we do not observe intense luminescence from this
level as it loses its energy non-radiatively to its lower level and
populates 4F9/2 level. The energy separation between 4F9/2 state
and next lower lying 6F1/2 is about 6000 cm  1. The value of σ e has
been used to identify the potential laser transition of RE ion in a
host medium. A good laser transition can have a large stimulated
emission cross-section. The σ e of 42.610  22 cm2 obtained for the
4
F9/2-6H13/2 transition in LCZSFBDy10 glass is higher than other
reported glass systems [50,51,54–57] except in the reported
glasses[52,53]. The gain bandwidth σ e  Δλp and σ e  τR para-
meters are critical to predict the amplification of the medium in
which the RE ions are situated. A good optical amplifier should
have higher values of σ e  Δλp and σ e  τR . The relatively higher
values of σ e  Δλp and σ e  τR suggest that the LCZSFBDy10 glass is
a suitable candidate for optical amplifiers and yellow lasers.
Fig. 4.3. Emission spectrum of LCZSFBDy10 glass The strong blue and yellow luminescence  originating
 from the
4
F9/2 state are magnetic dipole (MD), (ΔJ  ¼ 7 1, but 020
 
forbidden) and electric dipole (ED), (ΔJ  ¼ 7 2) transitions respec-
the spectral region of 400–700 nm consists of four bands centered tively. Generally luminescence intensity ratio (ED/MD) has been
at around 456, 484, 576 and 664 nm and they are assigned to used to know the symmetry of the local environment of the RE
4
I13/2-6H13/2, 4F9/2-6H15/2, 6H13/2, and 6H11/2 transitions, respec- ions. Greater the intensity of ED transition, higher the asymmetry
tively. Among the three well-known bands, arising from 4F9/2 state, nature [62]. In case of Dy3 þ this ratio is generally termed as yellow
the emissions at 484 nm and 576 nm are observed to be strong, to blue luminescence intensity ratio (Y/B) due to 4F9/2-6H13/2
which correspond to the blue and yellow regions of the visible (yellow, ED) and 4F9/2-6H15/2 (blue, MD) transitions and is used to
spectrum respectively. In general, the optical properties of RE characterize the Dy–O bond covalence and the higher value of Y/B
doped systems are often influenced by the structure of glass and has lower covalence when compared with other reported
4
matrix synthesis technique [46].  The
 F9/2-6H13/2 is a hypersensi- systems [50–52,63–69] except in [70–75]. The Y/B intensity ratio
 
tive electric dipole transition ( ΔJ ¼ 7 2) which has been strongly of LCZSFBDy10 and other glass hosts is listed in Table 4.5. It can be
influenced by the crystal field environment. In the emission seen that there is a considerable variation in Y/B intensity ratio
spectrum of Dy3 þ ions, the yellow emission (4F9/2-6H13/2) is with composition of the host glass material. The chromaticity color
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 575

Table 4.3
Radiative properties of LCZSFBDy10 glass.

Transition 4F9/2- λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  22 cm2) σ e  Δλp (10  28 cm3) σ e  τ m (10  25 cm2 s)

6
H11/2 664 19.20 115.0 0.06 0.01 6.00 11.52 2.10
6
H13/2 576 16.30 1205.0 0.66 0.37 42.60 69.44 14.80
6
H15/2 484 18.92 383.8 0.21 0.60 5.83 11.03 2.03

Table 4.4
Comparison of radiative properties of 4F9/2-6H13/2 emission transition in LCZSFBDy10 glass with different glass hosts.

Glass λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  22 cm2) σ e  Δλp (10  28 cm3) σ e  τ m (10  25 cm2 s)

LCZSFBDy10 576 16.30 1205.0 0.66 0.37 42.60 69.44 14.80


PKAZLFDy10 [50] 577 13.04 594 0.77 0.43 28.2 36.7 17.43
ZnAlBiBDy10 [51] 577 15.21 875 0.59 0.47 26.2 39.85 12.34
PbPKANDy10 [52] 576 13.8 1188 0.67 0.72 54.5 75.21 25.67
PKAZFDy10 [53] 577 15.01 1095 0.69 0.43 45.7 68.59 27.88
PTBDy10 [54] 576 16 646 0.63 0.46 28.6 4.42 17.67
BNLDy10 [55] 574 15.88 834.7 0.66 0.68 35.84 56.9 14.52
CFBDy10 [56] 573 16.94 714 0.62 0.64 23.34 39.53 14.10
LBEDy10 [57] 575 15.00 639.57 0.66 0.61 26.01 39.02 –

Table 4.5
Comparison of yellow to blue intensity ratio (Y/B), chromaticity coordinates (x, y)
and correlated color temperature (CCT, K) for Dy3 þ systems.

System Y/B ratio Color coordinates CCT

x y

LCZSFBDy10 0.70 0.33 0.37 5595


PKAZLFDy10 [50] 0.77 0.32 0.37 5992
ZnAlBiBDy10 [51] 1.19 0.33 0.30 5532
PbPKANDy10 [52] 4.50 0.45 0.46 3200
PKAZFDy10 [53] – 0.31 0.34 6569
CFBDy10 [56] – 0.40 0.45 3990
NAPDy10 [63] 2.9 0.44 0.44 3223
PbFPDy10 [64] 0.63 0.31 0.34 6567
CaTPDy10[65] 1.23 0.33 0.34 5609
SrTPDy10[65] 1.38 0.33 0.34 5609
MgTPDy10[65] 1.20 0.32 0.35 6038
LTTDy10 [66] 1.02 0.34 0.40 5270
Silicate [67] 2.03 0.39 0.44 4142
SLBiBDy10 [68] 0.97 0.44 0.42 3080
LFBMDy10 [69] – 0.44 0.44 3080
Lead borate tungsten [70] 3.22 0.41 0.45 3803
Lead tellurofluoroborate [71] 0.88 0.32 0.36 6008
Fig. 4.4. CIE chromaticity diagram of LCZSFBDy10 glass. Phosphate [72] 1.99 0.39 0.44 4153
Fluorophosphate [72] 1.55 0.37 0.41 4462
Oxyfluoride [73] 1.05 0.34 0.39 5285
coordinates (x,y) of LCZSFBDy10 glass and other reported glass Oxyfluoride [74] 1.52 0.39 0.41 3982
Zinc borophosphate [75] 1.21 0.36 0.41 4722
hosts are listed in Table 4.5. The CIE chromaticity coordinates for
the present LCZSFBDy10 glass are found to be (0.33, 0.37) are
plotted in Fig. 4.4. From Fig. 4.4 and Table 4.5, it is observed that The decay profile of 4F9/2 emission level of Dy3 þ ion in LCZSFB
the chromaticity coordinates are very close to the standard equal glass was recorded at 386 nm excitation wavelength at 576 nm
energy white light illuminate. These results confirm that emission wavelength is shown in Fig. 4.6 The decay profile exhibit
LCZSFBDy10 glass is promising materials for white light emission. single exponential nature according to the equation It ¼ I0  exp
The values obtained for the LCZSFBDy10 glass is comparable with ( t/τ), where It is the actual luminescence intensity, I0 is the
other Dy3 þ doped [50–53,56,63–75] systems, which are very luminescence intensity at the start of the decay process, t is the
nearer to white light illuminate. The correlated color temperature time and τ is the decay time. The decay time of LCZSFBDy10 is found
(CCT) was calculated by using the color coordinates by the to be 348 ms. By applying the JO theory, the radiative decay time (τR )
McCamy’s approximate formula [76]: CCT ¼  49n3 þ3525n2  of the 4F9/2 emission level is determined as 550 ms. It is evident from
6823nþ 5520.33, where n ¼ ðx xe Þ=ðy  ye ) is the inverse slope the radiative and the experimental decay times that the emission
line and (xe ¼ 0.332, ye ¼ 0.186) is the epicenter. The CCT values from the 4F9/2 level contains both radiative and non-radiative
obtained for the LCZSFBDy10 glass along with other reported Dy3 þ contributions. In general, the non-radiative contribution might
systems are presented in Table 4.5. The CCT (5595 K) of takes place due to either energy transfer among the Dy3 þ ions or
LCZSFBDy10 glass is found to be higher than [50,53,64,65,72] multi phonon relaxation or both (Fig. 4.5). In the present case the
systems, and lower than [51,52,56,63,66–70,72–75] systems. The decay curve is single exponential and hence the non-radiative decay
CCT values for the LCZSFBDy10 glass is found to higher than rates (WNR) are equal to the multi phonon relaxation rates (WMPR)
fluorescent tube (3935 K) and day light (5500 K) [77]. only. The quantum efficiency, η¼τm /τR is a measure of the number
576 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

of photons emitted per excited ion. The quantum efficiency of 63% thermally populated 7F1 level to the excited states of Eu3 þ ion
of the LCZSFBDy10 glass suggests its utility as a potential material and similar to the other reported glasses [78–83]. It is clearly
for 576 nm laser applications. Table 4.6 presents τm , τR , η values for observed from the spectrum that, the bands in the UV–vis region
the present and several reported glasses [50,52–56]. correspond to 7F0-5D2 (464 nm), 7F0-5D1 (526 nm) and 7F1-5D1
(534 nm) transitions and the bands in the NIR region correspond
to 7F0-7F6 (2090 nm) and 7F1-7F6 (2210 nm) transitions respec-
5. Eu3 þ ions doped LCZSFB glass for visible red lasers and tively. The induced electric dipole 7F0-5D2 hypersensitive transi-
display devices tion is found to be more intense than the magnetic dipole allowed
7
F0-5D1 allowed transition. Since the absorption edge of the glass
The absorption spectrum of the LCZSFBEu10 glass recorded in matrix rises rapidly, transitions below 450 nm could not be
the UV–vis–NIR region is shown in Fig. 5.1 along with the band observed. By measuring the areas under the individual absorption
assignments. The observed bands are due to the in homogeneously bands, the experimental oscillator strengths (f exp ) are determined.
broadened f–f transitions from the 7F0 ground level and the The small energy gap of 260 cm  1 between 7F0 and 7F1 states at
room temperature indicates that some Eu3 þ ions also exist in the
first excited state 7F1. Hence the obtained oscillator strengths are
thermally corrected to get the exact experimental oscillator
strengths. Thermal correction enhances the oscillator strength of
absorption transitions and provides reliable JO intensity para-
meters. The experimental band energies and oscillator strengths
without thermal correction and with thermal correction are
presented in Table 5.1. The JO intensity parameters without

Table 4.6
Comparison of decay properties of LCZSFBDy10 glass with different glass hosts.

Glass τ m (ms) τ R (ms) η ð%Þ W NR (s  1)

LCZSFBDy10 348 550 63 1055


PKAZLFDy10 [50] 618 1115 55 721
PbPKANDy10 [51] 471 597 84 448
PKAZFDy10 [52] 610 640 96 77
PTBDy10 [53] 618 979 63 597
BNLDy10 [54] 405 428 95 133
CFBDy10 [55] 604 873 69 510

Fig. 4.5. Energy level diagram of LCZSFBDy10 glass.

Fig. 5.1. Absorption spectrum of LCZSFBEu10 glass.

Table 5.1
Experimental and calculated oscillator strengths of LCZSFBEu10 glass.

Transition Energy f exp 10  6


(cm  1)
Without thermal With thermal
correction correction

7
F0-5D2 21,552 0.38 0.6
7
F0-5D1 19,012 0.04 0.06
7
F1-5D1 18,727 0.07 0.11
7
F0-7F6 4,785 1.48 2.35
7
F1-7F6 4,525 0.58 0.92
Fig. 4.6. Decay profile of 4F9/2 level of LCZSFBDy10 glass.
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 577

Table 5.2
Comparison of J–O intensity parameters (10  20 cm2) of LCZSFBEu10 glass with different glass hosts.

Glass system Absorption

Without thermal correction With thermal correction Emission

Ω2 Ω4 Ω6 Ω2 Ω4 Ω6 Ω2 Ω4 Ω6

LCZSFBEu10 14.27  1.38 22.6 – 2.18 3.62 1.43 –


1EBT [78] – – – – – – 5.30 0.31 –
Eu:BaLFB [79] – – – – – – 2.94 0.09 –
Eu:SrLFB [79] – – – – – – 2.85 0.27 –
Eu:BiLFB [79] – – – – – – 2.89 0.12 –
Eu:PbLFB [79] – – – – – – 2.73 0.16 –
Eu:CdLFB [79] – – – – – – 2.46 0.13 –
EuNKZLS [80] – – – – – – 7.44 0.72 –
CFBEu [81] 6.62 5.01 2.98 9.95 7.34 4.39 3.33 0.59 –
ZFPEu10 [82] 14.25 7.43 3.26 22.02 13.60 5.04 3.62 0.41 –
1EPbFB [83] – – – – – – 2.55 0.36 –

Table 5.3
Comparison of intensity ratio R ¼(5D0-7F2)/
(5D0-7F1) of LCZSFBEu10 glass with
different hosts.

Host matrix Intensity ratio (R)

LCZSFBEu10 2.38
1EBT [78] 4.12
Eu:BaLFB [79] 1.86
Eu:SrLFB [79] 1.84
Eu:BiLFB [79] 1.83
Eu:PbLFB [79] 1.77
Eu:CdLFB [79] 1.59
NKZLS [80] 4.90
ZFPEu10 [82] 2.48
1EPbFB [83] 1.79
1.0LBTAF [88] 1.76
Eu:LiLTB [89] 2.08
Eu:NaLTB [89] 2.15
Eu:KLTB [89] 2.25
PKFMAEu10 [90] 3.03 Fig. 5.3. Emission spectrum of LCZSFBEu10 glass.
PKFSAEu10 [90] 2.85
PKFBAEu10 [90] 3.75
determined from the emission spectra by evaluating the lumines-
cence intensity ratio of the 5D0-7FJ (J ¼2, 4 and 6) transitions to
the intensity of 5D0-7F1 magnetic dipole transition [84]. Table 5.2
presents the comparison of JO intensity parameters calculated
from the absorption and emission spectra with other glass hosts.
The current trend in the JO intensity parameters (Ω2 4 Ω4 ) has
also been observed for other glass matrices [78–83]. The higher
magnitude of Ω2 suggests that the ligands around the Eu3 þ ions
possesses higher asymmetry as well as high covalency.
The excitation spectrum of LCZSFBEu10 glass obtained by
monitoring the emission at 612 nm is shown in Fig. 5.2. From
the excitation spectrum, high energy transitions have been located
and assigned. The excitation bands are observed at 360, 365, 375,
380, 392 and 423 nm corresponding to the 7F0-5D4, 7F1-5D4,
7
F0-5G2, 7F1-5L7, 7F0-5L6 and 7F1-5D3 transitions, respectively.
Among these, the 7F0-5L6 (392 nm) transition is found to be more
prominent and it was used to record the emission spectra of
LCZSFBEu10 glass in the 550–750 nm spectral region. The emission
spectra shown in Fig. 5.3 recorded by exciting the glass with
392 nm exhibit five bands at 579, 591, 612, 655 and 702 nm
Fig. 5.2. Excitation spectrum of LCZSFBEu10 glass. corresponding to the 5D0-7FJ (J ¼0, 1, 2, 3 and 4) transitions,
respectively. The presence of the non-degenerate 5D0-7F0 transi-
thermal correction and with thermal correction using the experi- tion in the emission spectrum reveals that Eu3 þ ions are located at
mental oscillator strengths of 7F0-5D2 and 7F0-7F6 absorption the sites without inversion symmetry and in an environment of
2 2
bands because for these two transitions, the :U 2 : and :U 6 : low symmetry [85,86]. When the LCZSFBEu10 glass is excited at
4 2
respectively are the only non-zero matrix elements. As :U : 392 nm (7F0-5L6), 5L6 level is well populated. All the transitions
for all the observed transitions, the contribution of Ω4 parameter observed in the luminescence spectrum start from the 5D0 level,
becomes zero. The JO intensity parameters have also been which is populated by radiation less depopulation of the 5L6 level.
578 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

The observed narrow emission bands in the emission spectra is covalent bonding nature between Eu3 þ ion and the surrounding
due to the shielding effect of 4f6 electrons by 5s and 5p electrons ligand. The asymmetry ratio (R) of the LCZSFBEu10 glass is found
in outer shells in the Eu3 þ ion. The transitions 5D0-7FJ (J ¼5 and to be 2.38 and is smaller when compared with other glass hosts
6) are not observed as transition probabilities of these transitions [79,83,88,89] and higher when compared with [78,80,82,90]
are very weak and the 5D0-7FJ (J ¼2, 4 and 6) transitions are (Table 5.3 ).
electric dipole transition. The emission peak corresponding to the By using the JO intensity parameters and the refractive index,
hypersensitive 5D0-7F2 transition in the red region around various radiative properties such as radiative transition probability
612 nm is the most intense among all emission transitions and it (AR ), radiative lifetime (τR ) and branching ratio (βR ) for the

is very sensitive
 to the site symmetry of Eu ions following the 5
D0-7FJ (J¼ 1, 2) transitions of Eu3 þ ion have been evaluated
selection rule ΔJ  ¼ 2. Another transition 5D0-7F1 (591 nm) with and are presented in Table 5.4. The Ωλ parameters which are
 
ΔJ  ¼ 1 could be found as a magnetic dipole transition and is determined by applying thermal correction to 7F0-5D2 and 7F6
independent of the crystal field strength around Eu3 þ ion [87]. oscillator strengths have found maximum value of the sponta-
When a Eu3 þ ion occupies an inversion symmetry site in the host neous emission probability rates (AR ). The radiative transition
matrix, the orange emission will be the dominated emission. The probability rate is found to be considerably larger in magnitude
emission mechanism (partial energy level diagram) indicating the for the 5D0-7F2 transition with a magnitude of 534 s  1. In order
possible emission transitions of Eu3 þ ions in the LCZSFBEu10 at to know the probable lasing transition in the emission spectrum of
392 nm excitation wavelength shown in Fig. 5.4. Due to the Eu3 þ ion, the effective bandwidth (ΔλP ), experimental branching
presence of high energy photons in the glasses, the emission ratio (βm ) and peak stimulated emission cross-section σ e for the
starting from the excited levels 5DJ (J¼ 1, 2, and 3) are suppressed i. 5
D0-7FJ (J¼ 1, 2) transitions have been determined and are
e., there is a fast non-radiative relaxation takes place at 5D0 level collected in Table 5.4. The large stimulated emission cross-
when the Eu3 þ ions are excited to any level above the 5D0 level. section is one attractive feature for the design and development
The ratio of the emission intensity of the 5D0-7F2 transition to of low threshold and high gain laser applications. The value of σ e is
that of the 5D0-7F1 transition defined as asymmetry ratio (R) can found to be higher for the hypersensitive transition 5D0-7F2. To
be used to investigate the relative strength of covalent/ionic obtain highly efficient and stable laser active materials, the optical
bonding between Eu3 þ ions and the polarization of the surround- gain parameter (σ e  τR ) and optical gain bandwidth (σ e  ΔλP ) are
ing of the Eu3 þ ions by short range effects and to establish the very important and their values are found to be 43.1110  25 cm2 s
degree of asymmetry in the vicinity of Eu3 þ ions. Higher the value and 395.710  25 cm3. The luminescence branching ratio is
of R, lower the symmetry around the Eu3 þ ions and higher the Eu– another important parameter that characterizes the lasing power
O covalence and vice versa. Hence, Ω2 and the asymmetry ratio R of a transition and it is well established that an emission transition
strongly reveal similar physical significance of the asymmetry and having βR greater than 50% is considered to be more potential for
laser emission. The experimental and radiative branching ratios
are found to be 58% and 90% respectively. Tables 5.5 presents a
comparison of λP , ΔλP , AR , βR , βm , σ e of LCZSFBEu10 glass with other
glass hosts [78,79,81–83]. Moreover, from Table 5.6, it is concluded
that 5D0-7F2 transition of LCZSFBEu10 glass has the highest value
of σ e besides βR 450% of any other glass [78,79,81–83] which
suggest that it may be useful for the development of visible red
laser as well as optical display devices at around 612 nm.
The fluorescence decay curve of LCZSFBEu10 for the 5D0-7F2
transition at room temperature are measured by exciting the
samples at 392 nm and monitoring the emission at 612 nm is
shown in Fig. 5.5. The decay curve is found to be almost single
exponential nature with lifetime (τm ) 1.37 ms obtained by taking
the first e-folding times of the emission intensities. The radiative
lifetime (τR ) obtained by JO theory is 1.87 ms. The variation
between τm and τR may be due to non-radiative process. The
non-radiative relaxation rates (WNR) found to be 142 s  1 and such
a low value indicates the negligible non-radiative process. The
quantum efficiency (η) was found to be 80%. Table 5.6 presents the
comparison of τm , τR , η, and WNR of 5D0-7F2 transition of
LCZSFBEu10 glass with different glass hosts [78,79,81–83]. The
quantum efficiency of the present glass host is higher than those
other reported glass hosts [78,79,81,83] except in the glass host
[82]. In the present investigation, the value of η for the 5D0 level is
found to be nearly 80% due to the absence of major non-radiative
channels such as cross-relaxation, multi phonon relaxation (MPR)
etc., from the 5D0 level and hence LCZSFBEu10 glass is suitable for
Fig. 5.4. Energy level diagram of LCZSFBEu10 glass. potential laser material.

Table 5.4
Radiative properties of LCZSFBEu10 glass.

Transition 5D0- λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  22 cm2) σ e  Δλp (10  28 cm3) σ e  τ m (10  25 cm2 s)

5
D0-7F1 591 12.3 55 0.10 0.25 2.9 3.52 3.97
5
D0-7F2 613 12.2 534 0.90 0.58 32.5 39.57 44.53
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 579

Table 5.5
Comparison of radiative properties of 5D0-7F2 emission transition in LCZSFBEu10 glass with different glass hosts.

Glass λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  22 cm2) σ e  Δλp (10  28 cm3) σ e  τ m (10  25 cm2 s)

LCZSFBEu10 613 12.2 534 0.90 0.58 32.5 39.57 44.53


2EBT [78] 616 6.33 250 0.78 0.78 27.69 17.5 37.93
Eu:BaLFB [79] 616 11.31 109 0.62 0.65 7.2 8.14 17.57
Eu:BiLFB [79] 616 10.97 98 0.61 0.63 7.6 8.33 22.8
Eu:CdLFB [79] 616 10.4 88 0.59 0.60 6.5 6.76 19.24
Eu:PbLFB [79] 616 10.25 97 0.61 0.63 7.4 7.58 20.57
Eu:SrLFB [79] 616 10.83 101 0.63 0.64 7.5 8.12 22.8
Eu:ZnLFB [79] 616 9.76 117 0.65 0.67 9.2 8.98 27.97
CFBEu10 [81] 612 11.38 142 0.52 0.64 9.0 10.2 20.88
ZFBEu10 [82] 611 10 203 0.58 0.66 15.8 15.8 39.97
1EPbFB [83] 615 8.11 212 0.62 0.59 13.5 10.94 22.27

Table 5.6
Comparison of decay properties of LCZSFBEu10 glass with different glass hosts.

Glass τ m (ms) τ R (ms) η ð%Þ W NR (s  1)

LCZSFBEu10 1.37 1.7 80 142


2EBT [78] 1.54 3.65 42 375
Eu:BaLFB [79] 2.44 6.0 41 242
Eu:SrLFB [79] 3.04 6.37 48 172
Eu:BiLFB [79] 3.0 6.41 47 180
Eu:PbLFB [79] 2.78 6.44 42 204
Eu:CdLFB [79] 2.96 6.9 44 193
Eu:ZnLFB [79] 3.04 5.74 53 154
CFBEu10 [81] 2.32 3.65 63 157
ZFBEu10 [82] 2.53 2.83 89 42
1EPbFB [83] 1.65 2.81 58 252

Fig. 6.1. Absorption spectrum of LCZSFBTb10 glass.

Table 6.1
Experimental and calculated oscillator strengths of LCZSFBTb10 glass.

Transition (7F6-) Energy (cm  1) f exp 10  6 f cal 10  6

5
D4 20,619 0.05 0.05
7
F1 5,280 1.19 0.97
7
F2 5,120 0.97 1.18
7
F3 4,533 0.98 0.94
δrms ¼ 7 0.07610  6

Ω2 4Ω6 4Ω4. The higher value of Ω2 suggested that in LCZSFBTb10


glass has higher covalency and asymmetry between Tb–O bonds.
Fig. 5.5. Decay profile of 5D0 level of LCZSFBEu10 glass. When compared with the other hosts [91,93,95,97], the sites
occupied by Tb3 þ ions in LCZSFBTb10 glass posses lower symme-
6. Tb3 þ ions doped LCZSFB glass for green fiber lasers try and greater degree of covalence for Tb–O bond.
Fig. 6.2 presents the excitation spectrum of LCZSFBTb10 glass
Fig. 6.1 represents the room temperature absorption spectra of recorded in the region 300–500 nm at 547 nm emission wavelength
LCZSFBTb10 glass in the spectral region 450–2500 nm. A total of corresponding to 5D4-7F5 emission level. It consists of five excita-
four absorption bands 7F6-5D4 (485 nm), 7F1 (1894 nm), 7F2 tion transitions 7F6-5D1, 5L9, 5L10, 5G6 and 5D4 at 318, 352, 370, 378
(1953 nm) and 7F3 (2206 nm) are observed. All these transitions and 485 nm, respectively. The emission spectra of LCZSFBTb10 glass
are similar to the other reported glasses [91–97] and are due to was measured by exciting at 378 nm and is shown in Fig. 6.3. In the
electric dipole interaction following the selection rules, ΔS ¼0, | case of Tb3 þ ions doped glasses the emissions below 480 nm
ΔL| r 6, |ΔJ| r 6. The experimental and theoretical oscillator originate from the 5D3 (blue) levels and the emissions above
strengths (f exp; f cal ) of present LCZSFBTb10 glass is presented in 480 nm originate from 5D4 (green) levels. The emission spectra in
Table 6.1 and the relatively small root mean square deviation (δrms ) the region 400–650 nm originating from 5D4 state contain four
70.07610  6 between f exp and f cal confirm the validity and peaks at 492, 547, 588 and 623 nm assigned to 5D4-7F6, 7F5, 7F4
applicability of JO theory for the LCZSFBTb10 glass system. The and 7F3 transitions. Among the observed transitions 5D4-7F5
obtained values of the JO intensity parameters along with the (green, 547 nm) transition is the most intense one. The emission
intensity parameters of other glass hosts are presented in from the excited 5D3 state has been reported by several researchers
Table 6.2. The JO intensity parameters are found in the order [92,93,95]. Three transitions are identified at 418, 442 and 458 nm
580 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

Table 6.2
Comparison of J–O intensity parameters, root mean square deviation for LCZSFBTb10 glass with different glass hosts.

Glass system Ω2 (10  20 cm2) Ω4 (10  20 cm2) Ω6 (10  20 cm2) Trend δrms (10  6)

LCZSFBTb10 14.54 0.17 4.09 Ω2 4Ω6 4Ω4 7 0.076


LBTAFTb10 [91] 13.23 1.23 2.96 Ω2 4Ω6 4Ω4 7 0.09
PTBTb10 [93] 11.98 2.87 10.51 Ω2 4Ω6 4Ω4 7 0.41
SFBTb10 [95] 0.45 5.53 1.39 Ω4 4Ω6 4Ω2 7 0.08
PbBAlTb10:A [97] 9.05 2.52 4.57 Ω2 4Ω6 4Ω4 7 0.011
PbBAlTb10:B [97] 8.78 2.39 4.02 Ω2 4Ω6 4Ω4 7 0.014
PbBAlTb10:C [97] 8.62 2.27 3.71 Ω2 4Ω6 4Ω4 7 0.046
PbBAlTb10:D [97] 8.20 1.95 3.31 Ω2 4Ω6 4Ω4 7 0.051
PbBAlTb10:E [97] 8.54 2.20 3.50 Ω2 4Ω6 4Ω4 7 0.013
PbBAlTb10:F [97] 7.92 2.03 3.32 Ω2 4Ω6 4Ω4 7 0.034
PbBAlTb10:G [97] 7.59 2.01 3.27 Ω2 4Ω6 4Ω4 7 0.029
PbBAlTb10:H [97] 7.47 2.15 3.14 Ω2 4Ω6 4Ω4 7 0.048

Fig. 6.4. Partial energy level diagram of LCZSFBTb10 glass.

diagram drawn for Tb3 þ ions in LCZSFBTb10 glass is shown in


Fig. 6.2. Excitation spectrum of LCZSFBTb10 glass. Fig. 6.4. The radiative transition rates (AR ), effective emission
bandwidth (Δλp ), branching ratio (βR ), measured branching ratio
(βm ) and stimulated emission cross-section (σ e ) have been deter-
corresponding to 5D3-7F5, 7F4 and 7F3 transitions respectively. mined for the 5D4 level of LCZSFBTb10 glass and are presented in
When the Tb3 þ ions are incorporated in pairs or clusters, no blue Table 6.3. The fluorescent properties such as measured branching
emission is observed. The existence of blue emission indicates the ratio (βm ), lifetime (τm ), stimulated emission cross-section (σ e ),
homogeneous distribution of Tb3 þ ions in the LCZSFBTb10 glass optical gain parameter (σ e  τm ) and quantum efficiency (η) are
[98]. Based on the energies of different transitions obtained from of important to select a good host material. The transition 5D4-7F5
the absorption, excitation and emission spectra, the energy level has highest value of AR , βR , βm , σ e when compared with the other
reported glass hosts [91,95,97] (Table 6.4) and hence are very useful
for the applications of LCZSFBTb10 glass as gain medium of the
green fiber luminescence.
The fluorescence decay profile of 5D4 emission level of
LCZSFBTb10 glass is recorded by monitoring the emission and
excitation wavelengths at 547 and 378 nm, respectively and is
shown in Fig. 6.5. The decay curve is suitably fit to single
exponential function indicating the absence of multi phonon decay
from 5D4 excited states to its lower lying levels. The experimental
lifetime (τm ) of 5D4 level determined by taking the first e-folding
times of the decay curve is 1.39 ms. The τm value is significantly
smaller than the predicted value (τR ¼ 2.0 ms) obtained using the JO
theory. This change may be due to resonant energy transfer but not
due to multi phonon relaxation because the energy gap between
the emitting level (5D4) and the next lower level (7F0) is very high
about (14,000 cm  1). The magnitude of resonant energy transfer
(WRT) calculated is 219 s  1. For photonic applications like fiber
lasers and planar waveguides the fluorescence quantum efficiency
(η) is an important factor controlling the performance of the
Fig. 6.3. Emission spectrum of LCZSFBTb10 glass. (For interpretation of the material. For LCZSFBTb10 glass the η is found to be 70%, which is
references to color in this figure legend, the reader is referred to the web version very high and hence 5D4-7F5 transition is utilized for green
of this article.) luminescence for fiber laser. Table 6.5 presents a comparison of
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 581

Table 6.3
Radiative properties of LCZSFBTb10 glass.

Transition 5D4- λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  22 cm2) σ e  Δλp (10  28 cm3) σ e  τ m (10  25 cm2 s)

7
F6 492 13.88 47 0.094 0.398 1.04 1.44 1.45
7
F5 547 12.00 395 0.790 0.547 15.46 18.55 21.49
7
F4 588 10.49 14 0.028 0.043 0.84 0.88 1.17
7
F3 623 7.11 44 0.086 0.012 4.80 3.41 6.66

Table 6.4
Comparison of radiative properties of 5D4-7F5 emission transition in LCZSFBTb10 glass with different glass hosts.

Glass λP (nm) Δλp (nm) βR βm σ e (10  22 cm2) σ e  Δλp (10  28 cm3) σ e  τ m (10  25 cm2 s)

LCZSFBTb10 547 12.00 395 0.79 0.547 15.46 18.55 21.49


LBTAFTb10 [91] 544 11.6 274 0.76 0.52 12.78 14.85 22.5
SFBTb10 [95] 544 12.36 34.4 0.31 0.51 1.37 1.69 4.4
PbBAlTb10:A [97] 543 11.88 244.84 0.739 – 9.94 11.8 28.7
PbBAlTb10:B [97] 543 11.2 260.03 0.743 – 10.53 11.8 27.77
PbBAlTb10:C [97] 543 10.19 265.64 0.746 – 11.51 11.72 28.97
PbBAlTb10:D [97] 543 12.06 266.75 0.751 – 9.41 11.35 22.4
PbBAlTb10:E [97] 543 12.40 288.83 0.749 – 9.70 12.03 22.19
PbBAlTb10:F [97] 543 12.15 276.86 0.748 – 9.31 11.31 20.83
PbBAlTb10:G [97] 543 12.26 261.68 0.746 – 8.8 10.79 19.91
PbBAlTb10:H [97] 543 11.88 254.11 0.745 – 8.9 10.57 26.09

Table 6.5
Comparison of decay properties of LCZSFBTb10 glass with different glass hosts.

Glass τm (ms) τR (ms) η ð%Þ W NR (s  1)

LCZSFBTb10 1.39 2.00 70 219


LBTAFTb10 [91] 1.76 2.76 46 425
PTBTb10 [93] 0.78 2.75 28 918
SFBTb10 [95] 3.21 8.92 36 199
PbBAlTb10:A [97] 2.887 3.019 95.63 15
PbBAlTb10:B [97] 2.637 2.859 92.24 29
PbBAlTb10:C [97] 2.517 2.807 89.67 41
PbBAlTb10:D [97] 2.382 2.816 84.59 65
PbBAlTb10:E [97] 2.288 2.592 88.27 52
PbBAlTb10:F [97] 2.237 2.702 82.79 77
PbBAlTb10:G [97] 2.263 2.851 79.38 91
PbBAlTb10:H [97] 2.245 2.931 76.60 104

Fig. 7.1. Absorption spectrum (400–650 nm) of LCZSFBNd10 glass.

Fig. 6.5. Decay profile of 5D4 level of LCZSFBTb10 glass. Fig. 7.2. Absorption spectrum (650–950 nm) of LCZSFBNd10 glass.

decay properties for LCZSFBTb10 glass with other glass hosts 7. Nd3 þ ions doped LCZSFB glass for 1.06 lm laser applications
[91,93,95]. The quantum efficiency of the present LCZSFBTb10 glass
is high then the reported glasses [91,93,95] except in the reported The absorption spectra of the LCZSFBNd10 glass recorded at
glass [97]. room temperature in the wavelength region 450–1000 nm is shown
582 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

Table 7.1 values of LCZSFBNd10 glass with other glass hosts [99–104]. In
Experimental and calculated oscillator strengths of LCZSFBNd10 glass. general, it is assumed that the higher the site symmetry, the lower
the intensity i.e., when the lanthanide ion is at a center of
Transition (4I9/2-) Energy (cm  1) f exp 10  6 f cal 10  6
symmetry, the intensity of hypersensitive transition is zero. The
4
F3/2 11,403 5.439 5.020 intensity of hypersensitive glasses in the glasses is enhanced by a
4
F5/2 þ 2H9/2 12,422 13.299 14.080 high asymmetry of the Nd3 þ ion site. In the present case the lesser
4
F7/2 þ 4S3/2 13,369 14.181 13.67 the value of Ω2, the more centro symmetrical the ion site and the
4
F9/2 14,641 1.955 1.105
2 more ionic its chemical bond with the ligands. The small value of Ω2
H11/2 15,924 0.100 0.125
4
G5/2 þ 2G7/2 17,094 33.379 33.422 in LCZSFBNd10 glass indicates that the host exhibit weaker cova-
4
G7/2 19,011 8.185 7.204 lence and lower symmetry for Nd–O bond when compared with
4
G9/2 þ 2K13/2 19,531 4.733 4.100 other glass hosts [99–104] that presented in Table 7.2.
2
G9/2 20,921 0.248 0.926 The NIR emission spectra of LCZSFBNd10 glass in the region
2
D3/2 þ 2P3/2 21,097 0.063 0.630
4
G11/2 þ 2K15/2 21,645 0.134 0.828
800–1500 nm, excited with 808 nm laser radiation are shown in
2
P1/2 23,256 0.076 1.343 Fig. 7.3. The 808 nm radiation excites the Nd3 þ ion to 4F5/2 level.
These excited Nd3 þ ions reach 4F3/2 metastable state through fast
δrms ¼ 7 0.510  6 non-radiative relaxation as indicated in the partial energy level
diagram (Fig. 7.4). The emission spectra from this 4F3/2 metastable
state shows three emission bands and are assigned to 4F3/2-4I9/2
in Figs. 7.1 and 7.2. This spectrum is similar to those found for other (912 nm), 4F3/2-4I11/2 (1060 nm) and 4F3/2-4I13/2 (1336 nm) tran-
Nd3 þ doped glass [99–104], except for some small changes in the sitions. From the JO theory and from emission spectra different
band positions and relative intensities. The absorption spectrum radiative parameters and certain laser characteristic parameters
reveals a total of 12 absorption bands originated from the 4I9/2 are calculated for the LCZSFBNd10 glass and are presented in
ground state to various excited states. The wavelengths of the Table 7.3. The total radiative transition probability (AT ) is found to
absorption bands along with their assignments and the values of be 6060 s  1 and the radiative lifetime (τR ) is 165 ms. The stimu-
experimental (f exp ) and calculated (f cal ) oscillator strengths are lated emission cross-section (σ e ) together with effective band-
listed in Table 7.1. The intensity parameters follow the trend as width (Δλp ) and experimental decay lifetime (τm ) are used to
Ω6 4Ω4 4Ω2 and are found higher than those reported for other determine gain parameters such as optical gain (σ e  τm ) and gain
Nd3 þ doped glass hosts [99–104]. In the present investigation, the bandwidth (σ e  Δλp ), which play a significant role in predicting
4
I9/2-4G5/2 transition centered at 585 nm is identified as a hyper- the amplification of the medium in which the Nd3 þ ions are
sensitive transition following the selection rule ΔS¼0, |ΔL|r 2, | doped. A higher value of βR signifies a higher stimulated emission
ΔJ|r 2. Table 7.2 compares the JO intensity parameters and δrms cross-section. 4F3/2-4I11/2 emission band possess relatively higher

Table 7.2
Comparison of J–O intensity parameters, root mean square deviation for LCZSFBNd10 glass with different glass hosts.

Glass system Ω2 (10  20 cm2) Ω4 (10  20 cm2) Ω6 (10  20 cm2) Trend δrms (10  6)

LCZSFBNd10 6.11 9.29 10.14 Ω6 4Ω4 4Ω2 7 0.5


ZnAlBiBNd10 [99] 5.31 3.84 4.61 Ω2 4Ω6 4Ω4 7 0.41
LaNaZnTeNd [100] 4.49 5.04 4.31 Ω4 4Ω2 4Ω6 –
LBTAFNd05 [101] 5.82 1.88 4.74 Ω2 4Ω6 4Ω4 7 0.38
BNaNd5 [102] 7.79 3.03 2.86 Ω2 4Ω4 4Ω6 7 0.47
BNaNd4 [102] 4.99 5.34 4.72 Ω4 4Ω2 4Ω6 7 0.37
BNaNd3 [102] 5.92 4.68 2.46 Ω2 4Ω4 4Ω6 7 0.49
TZNLNNd10 [103] 2.13 3.29 3.83 Ω6 4Ω4 4Ω2 7 0.39
CFBNd10 [104] 4.37 4.50 6.48 Ω6 4Ω4 4Ω2 7 0.43

Fig. 7.3. Emission spectrum (650–950 nm) of LCZSFBNd10 glass. Fig. 7.4. Partial energy level diagram of LCZSFBNd10 glass.
C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584 583

Table 7.3
Radiative properties of LCZSFBNd10 glass.

Transition 4F3/2- λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  20 cm2) σ e  Δλp (10  25 cm3) σ e  τ m (10  24 cm2 s)

4
I9/2 912 43.70 2330 0.38 0.14 1.94 0.85 2.62
4
I11/2 1060 26.29 2988 0.49 0.69 7.53 1.98 10.17
4
I13/2 1336 64.59 742 0.12 0.17 1.92 1.24 2.59

Table 7.4
Comparison of radiative properties of 4F3/2-4I11/2 emission transition in LCZSFBNd10 glass with different glass hosts.

Glass λP (nm) Δλp (nm) AR (s  1) βR βm σ e (10  20 cm2) σ e  Δλp (10  25 cm3) σ e  τ m (10  24 cm2 s)

LCZSFBNd10 1060 26.29 2988 0.49 0.69 7.53 1.98 10.17


ZnAlBiBNd10 [99] 1060 30 2136 0.50 0.72 3.67 1.10 –
LBTAFNd05 [101] 1070 34 1184 0.56 0.62 2.60 0.88 5.98
BNaNd5 [102] 1067 18.83 – 0.47 0.44 4.24 0.80 6.48
BNaNd4 [102] 1064 22.23 – 0.54 0.52 4.95 1.10 7.57
BNaNd3 [102] 1064 20.89 – 0.53 0.51 4.69 0.98 7.08
TZNLNNd10 [103] 1061 28.35 – 0.49 0.77 4.27 1.21 5.81

shown in Fig. 7.5. The experimental decay time (τm ¼135 ms) has
been determined by taking the first e folding time of intensity of
the decay curves. The predicted radiative lifetime (τR ) of the 4F3/2
emission state calculated from JO theory is 165 ms. The τm is found
shorter than τR . This may be due to energy transfer through cross-
relaxation at higher concentration. The non-radiative decay rate
(WNR) of LCZSFBNd10 glass is 1347 s  1. Table 7.5 presents a
comparison of decay properties for LCZSFBNd10 glass with other
glass hosts [101–104]. The quantum efficiency (η) of LCZSFBNd10
glass is 82% is found to be higher than those reported glasses
[101,102] except in the reported glasses [103,104].

8. Conclusions

In conclusion, the RE doped materials are of crucial importance


to energy saving lighting devices, displays, optical fibers and are
widely displayed in fiber amplifiers and solid state lasers. However
Fig. 7.5. Decay profile of 4F3/2 level of LCZSFBNd10 glass. the technological interest in RE luminescence reaches beyond
telecommunications, displays, laser materials, data storage, radia-
tion detection and medical applications are all areas in which
luminescent rare earths are playing an increasingly important role.
Table 7.5
Not only in RE luminescence in itself technologically important,
Comparison of decay properties of LCZSFBNd10 glass with different glass hosts.
but it can also be used as a valuable diagnostic tool to probe the
Glass τ m (ms) τ R (ms) η ð%Þ W NR (s  1) physical and optical properties of a range of optoelectronic
materials. There is a great deal of interesting physics to be
LCZSFBNd10 135 165 82 1347 investigated, particularly in the area of optoelectronic materials,
LBTAFNd05 [101] 190 470 50 3135
and the needs of technology are constantly rushing forward
BNaNd5 [102] 153 470 32 4408
BNaNd4 [102] 153 402 56 4047 research into new ways of integrating light emitting materials
BNaNd3 [102] 151 380 40 3990 with existing semiconductor and fiber technology. The authors
TZNLNNd10 [103] 136 154 88 859 believe that the results of the investigated LCZSFB glasses doped
CFBNd10 [104] 256 284 90 373
with RE ions may reach the useful applications in the diverse
scientific fields of present needs.

stimulated emission cross section and experimental branching


Acknowledgements
ratio (βm ) suggesting the LCZSFBNd10 glass is a suitable glass host
for 1.06 mm laser applications. Table 7.4 represents the comparison
One of the authors (BDPR) thankful to DAE-BRNS for providing
of radiative parameters of LCZSFBNd10 glass with other reported
the financial assistance to carry out this work vide Reference No.
glass systems [99,101–103]. The stimulated emission cross-section
34/14/06/2014-BRNS/0124 Dated: 21-04-2014.
(σ e ) is of the present LCZSFBNd10 glass is high compared to all the
reported glass systems [99–103] and the experimental branching
References
ratio (βm ) of LCZSFBNd10 glass is high in the glasses [101–103]
except in the glass [99].
[1] Naftaly M, Jha A. J Appl Phys 2000;87:2098.
The decay curves of 4F3/2 excited state recorded for the [2] Machewirth DP, Wei K, Krasteva V, Datta R, Snitzer E, Sigel Jr. GH. J Non-Cryst
LCZSFBNd10 glass is well fitted to a single exponential function Solids 1997;213:295.
584 C. Madhukar Reddy et al. / Renewable and Sustainable Energy Reviews 51 (2015) 566–584

[3] Nogami M, Abe Y. J Non-Cryst Solids 1996;197:73. [55] Arul Rayappan I, Marimuthu K, Surendra Babu S, Sivaraman M. J Lumin
[4] Yaru N, Chunhua L, Yan Z, Qitu Z, Zhongzi X. J Rare Earths 2007;25:94. 2010;130 2407-212.
[5] Koechner W. Solid state laser engineering. 3rd ed.Springer; 2006. [56] Suresh Kumar J, Pavani K, Mohan Babu A, Giri Neeraj Kumar, Rai SB, Moorthy
[6] Varshnaya AH. Fundamentals of inorganic glasses. San Diego: Academic L Rama. J Lumin 2010;130:1916–23.
Press; 1994. [57] Karunakaran RT, Marimuthu K, Surendra Babu S, Arumugam S. Physica B
[7] Nii H, Ozaki K, Herren M, Morita M. J Lumin 1998;76–77:116–9. 2009;404:3995–4000.
[8] Mori A, Ohishi Y, Sudo S. Electron Lett 1997;33:863–4. [58] Jorgensen CK, Judd BR. Mol Phys 1964;8:281.
[9] Kityk IV, Wasylak J, Benet S, Dorosz D, Kucharski J, Krasowski J, Sahraoui B. [59] Peacock RD. Struct Bond 1975;22:83–122.
J Appl Phys 2002;92:2260–8. [60] Ratnam BV, Jayasimhadri M, Jang K, lee HS, Yi SS, Jeong JH. J Am Ceram Soc
[10] Pisarski WA, Goryczka T, Wodecka-Dus B, Plonska M, Pisarska J. Mater Sci 2010;93:3857–61.
Eng B 2005;122:94. [61] Krishna KM, Anoop G, Jayaraj MK. J Electrochem Soc 2007;154:1310–3.
[11] Speghini A, Peruffo M, Casarin M, Ajo D, Bettinelli M. J Alloys Compd [62] Tanabe S, Kang J, Hanada T, Soga N. J Non-Cryst Solids 1998;239:170–5.
2000;300–301:174. [63] Amarnath Reddy A, Chandra Sekhar M, Pradeesh K, Surendra Babu S, Prakash
[12] Fermi F, Ingletto G, Aschieri C, Brttinelli M. Inorg Chim Acta 1989;163:123. G Vijaya. J Mater Sci 2011;46:2018–23.
[13] Saisudha MB, Ramakrishna J. Phys Rev B 1996;53:6186. [64] Kesavulu CR, Jayasankar CK. Mater Chem Phys 2011;130:1078–85.
[14] Shojiya M, Takahashi M, Kanno R, Kawamoto Y, Kadono K. J Appl Phys [65] Moorthy DVR, Jamalaiah BC, Mohan Babu A, Sasikala T, Moorthy L Rama. Opt
1997;82:6259–66. Mater 2011;32:1112–6.
[15] Colaizzi J, Matthewson MJ. J Lightwave Technol 1994;12:1317. [66] Mohan Babu A, Jamalaiah BC, Suresh Kumar J, Sasikala T, Rama Moorthy L.
[16] Yamasaki H, Minato K, Nezaki D, Okamoto T, Kawamoto A, Takata M. Solid J Alloys Compd 2011;509:457–62.
State Ion 2004;172:349–52. [67] Sun X, Huang S, Gong X, Gao Q, Ye Z, Cao C. J Non-Cryst Solids
[17] Alivov YI, Clook D, Ataev BM, Chukichev MV, Mamedov VV, Zineko VI, 2010;356:98–101.
Agafonov YA, Pustovit AN. Solid State Electron 2004;48:2343–6. [68] Rajesh D, Ratnakaram YC, Seshadri M, Balakrishna A, Krishna T Sathya.
[18] Saito N, Haned H, Skiguchi T, Ohashi N, Sakaguchi I, Koumoto K. Adv Mater J Lumin 2012;132:841–9.
2002;14:418–21. [69] Balakrishna A, Rajesh D, Ratnakaram YC. J Lumin 2012;132:2984–91.
[19] Ferber J, Luther J. Phys Chem B 2001;105(21):4895–903. [70] Jamalaiah BC, Rama Moorthy L, Seo HJ. J Non-Cryst Solids 2012;358:204–9.
[20] Zhu BL, Xie CS, Zeng DW, Song WL, Wang AH. Mater Chem Phys 2005;89 [71] Saleem SA, Jamalaiah BC, Jayasimhadri M, Srinivasa Rao A, Jang K, Rama
(1):148–53. Moorthy L. J Quant Spectrosc Radiat Transfer 2011;112:78–84.
[21] Judd BR. Phys Rev 1962;127:750–61. [72] Kumar K Upendra, Srinivasa Rao Ch, Jayasankar CK, Surendra Babu S, Lucio JL,
[22] Ofelt GS. J Chem Phys 1962;37:511–9. Valleo H, Miguel A, Martinez Gamez Ma Alejandriana. Phys Procedia
[23] Gorller-Walrand C, Binnemans K. In: Gschneidner Jr. KA, Eyring L, editors. 2011;13:70–3.
Handbook on the physics and chemistry of rare earths, vol. 25. Amsterdam: [73] Babu P, Jang KH, Srinivasa Rao Ch, Shi L, Jayasankar CK, Lavin V, Seo HJ. Opt
North-Holland; 1998. p. 101–264 Chapter 167. Express 2011;19:1836–41.
[24] Tanabe S, Hirao K, Soga N. J Non-Cryst Solids 1989;113:178. [74] Babu P, Jang KH, Kim ES, Shi L, Seo HJ, Rivera-Lopez F, Rodriguez-Mendoza
[25] Tanabe S, Hirao K, Soga N. J Non-Cryst Solids 1992;142:148. UR, Lavin V, Vijaya R, jayasankar CK, Rama Moorthy L. J Appl Phys
[26] Ebendorff-Heidepriem H, Ehrt D. Glastech Ber Glass Technol 2009;105:013516–22.
[75] Jayasimhadri M, Jang K, Lee HS, Chen B, Yi SS, Jeong JH. J Appl Phys
1998;71:289–99.
2009;106:013105–9.
[27] Laser spectroscopy of solids. In: Yen MW, Selzer PM, editors. Berlin:
[76] McCamy CS. Color Res Appl 1992;17:142–4.
Springer; 1981.
[77] Fuchs EC, Sommer C, Wenzl FP, Bitschnau B, Paulitsch AH, Munianger A,
[28] Optical spectroscopy of glasses. In: Zschokke J, editor. Dordrecht: Reidel;
Gatterer K. Mater Sci Eng B 2009;156:73–8.
1986.
̈ der W. Laser spectroscopy, basic concepts and instrumentation. [78] Maheshvaran K, Marimuthu K. J Lumin 2012;132:2259–67.
[29] Demtro
[79] Arunkumar S, Marimuthu K. J Lumin 2013;139:6–15.
Berlin: Springer; 1996.
[80] Ramachari D, Rama Moorthy L, Jayasankar CK. J Lumin 2013;143:674–9.
[30] di Bartolo B. Optical interaction in solids. New York: John Wiley & Sons Inc.;
[81] Pavani K, Suresh Kumar J, Chengaiah T, Moorthy L Rama. J Mol Struct
1967.
2012;1028:170–5.
[31] Menzel R. Photonics linear and nonlinear interactions of laser light and
[82] Vijaya N, Jayasankar CK. J Mol Struct 2013;1036:42–50.
matter. Berlin: Springer; 2001.
[83] Arunkumar S, Venkata Krishnaiah K, Marimuthu K. Physica B
[32] Rodriguez VD, Martin IR, Alcala R, Cases R. J Lumin 1992;54:231–6.
2013;416:88–100.
[33] Jayasankar CK, Rukmini E. Opt Mater 1997;8:193–205.
[84] Babu P, Jayasankar CK. Physica B 2000;279:262.
[34] Potter jr BG, Sinclair MB. J Electroceram. 1998;2-4:295–308.
[85] Reisfeld R, Zigansky E, Gaft M. Mol Phys 2004;102:1319.
[35] Kenyon AJ. Prog Quantum Electron 2002;26:225–84.
[86] Binnemans K, Van Herck K, Gorller-Walrand C. Chem Phys Lett
[36] Auzel F. Chem Rev 2004;104:139–73.
1997;266:297.
[37] Tanabe S, Ohyagi T, Soga N, Hanada T. Phys Rev B 1992;46:3305–10.
[87] Blasse G, Bril A, Niewpoort WC. J Phys Chem Solids 1966;27:1587.
[38] Venkatramu V, Babu P, Jayasankar CK, Troster Th, Sievers W, Wortmann G.
[88] B.C. Jamalaiah, J. Suresh Kumar, A. Mohan Babu, K. Pavani, L. Rama Moorthy, J
Opt Mater 2007;29:1429–39. Alloys Compd 492 (2019). 63.
[39] Suhasini T, Suresh Kumar J, Sasikala T, Jang Kiwan, Sueb Lee Ho, Jayasimhadri [89] Saleem SA, Jamalaiah BC, Mohan Babu A, Pavani K, Rama Moorthy L. J Rare
M, Hyun Jeong Jung, Soo Yi Soung, Rama Moorthy L. Opt Mater Earths 2010;28(2):189.
2009;31:1167–72. [90] Upendra Kumar K, Babu S Surendra, Srinivasa Rao Ch, Jayasankar CK. Opt
[40] Shanmuga Sundari S, Marimuthu K, Sivaraman M, Babu S Surendra. J Lumin Commun 2011;284:2909.
2010;130:1313–9. [91] Jamalaiah BC, Suresh Kumar J, Mohan Babu A, Sasikala T, Moorthy L Rama.
[41] Srinivasa Rao Ch, Jayasankar CK. Opt Commun 2013;286:204–10. Physica B 2009;404:2020–4.
[42] Sasikala T, Rama Moorthy L, Babu A Mohan. Spectrochim Acta A [92] Sontakke Atul D, Biswas Kaushik, Annapurna K. J Lumin 2009;129:1347–55.
2013;104:445–50. [93] Jamalaiah BC, Vijaya Kumar MV, Gopal K Rama. Physica B 2011;406:2871–5.
[43] Ki-Soo Lim N Vijaya, Kesavulu CR, Jayasankar CK. Opt Mater [94] Zur L, Pisarska J, Pisarski WA. J Rare Earths 2012;29:1198–200.
2013;35:1557–63. [95] Umamaheswari D, Jamalaiah BC, Sasikala T, Chengaiah T, Il-Gon Kim L,
[44] Nayab Rasool Sk, Rama Moorthhy L, Jayasankar CK. Opt Commun Moorthy Rama. J Lumin 2012;132:1166–70.
2013;311:156–62. [96] Liaolin Zhang, Peng Mingying, Dong Gouping, Qiu Jianrong. Opt Mater
[45] Jorgensen CK, Reisfeld R. J Less-Common Met 1983;93:107. 2012;34:1202–7.
[46] Tanabe S, Ohyagi T, Todoroki S, Handa T, Soga N. J Appl Phys 1993;73:8451–4. [97] Abdul Azeem P, Kalidasan M, Reddy RR, Ramagopal K. Opt Commun
[47] Adam JL, Sibley WA. J Non-Cryst Solids 1985;76:267–79. 2012;285:3787–91.
[48] Jayasankar CK, Venkatramu V, Babu P, Troster Th, Sievers W, Wortmann G, [98] Lammers MJJ, Blasse G. Chem Phys Lett 1986;126:405.
Holzapfel WB. J Appl Phys 2005;97:093523. [99] Mahamuda Sk, Swapna K, Srinivasa Rao A, Jayasimhadri M, Sasikala T, Pavani
[49] Jayasankar CK, Babu P, Troster Th, Holzapfel WB. J Lumin 2000;91:33. K, Rama Moorthy L. J Phys Chem Solids 2013;74:1308–15.
[50] Vijaya N, Upendra Kumar K, Jayasankar CK. Spectrochim Acta A [100] Sobczyk Marcin. J Quant Spectrosc Radiat Transfer 2013;119:128–36.
2013;113:145–53. [101] Jamalaiah BC, Suhasini T, Moorthy L Rama, Kim Il-Gon, Yoo Dong-Sun, Jang
[51] Swapna K, Mahamuda Sk, Srinivasa Rao A, Jayasimhadri M, Sasikala T, Kiwan. J Lumin 2013;132:1144–9.
Moorthy L Rama. J Lumin 2013;139:119–24. [102] Karunakaran RT, Marimuthu K, Arumugam S, Surendra Babu S, Leon-Luis SF,
[52] Linganna K, Srinivasa Rao Ch, Jayasankar CK, Quant. J. Spect Radiat Trans Jayasankar CK. Opt Mater 2010;32:1035–41.
2013;118:40–8. [103] Surendra Babu S, Rajeswari R, Jang Kiwan, Eun Jin Cho, Jang KyoungnHyuk,
[53] Sreedhar VB, Ramachari D, Jayasankar CK. Physica B 2013;408:158–63. Jin Seo Hyo, Jayasankar CK. J Lumin 2010;130:1021–5.
[54] Vijaya Kumar MV, Jamalaiah BC, Rama Gopal K, Reddy RR. J Lumin [104] Suresh Kumar J, Mohan Babu A, Sasikala T, Moorthy L Rama. Chem Phys Lett
2012;132:86–90. 2010;484:207–13.

Você também pode gostar