Вы находитесь на странице: 1из 14

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/230787208

Optical and spectroscopic properties of rare


earth-doped (80−x)TeO2-xGeO2-10Nb2O5-
10K2O glasses

Article in Journal of Luminescence · August 2012


DOI: 10.1016/j.jlumin.2012.08.031

CITATIONS READS

10 181

4 authors:

Gonçalo Machado Monteiro Yigang Li


Technical University of Lisbon Technical University of Lisbon
4 PUBLICATIONS 38 CITATIONS 19 PUBLICATIONS 355 CITATIONS

SEE PROFILE SEE PROFILE

Luís F Santos Rui Almeida


University of Lisbon University of Lisbon
92 PUBLICATIONS 1,109 CITATIONS 164 PUBLICATIONS 3,680 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Mesoporous silicas doped with TiO2, PDT and Ancient Portuguese Ceramics View project

NOVOS FILMES FERROELÉCTRICOS LIVRES DE CHUMBO POR PLD PARA OPTOELECTRÓNICA (NOLEAD)
View project

All content following this page was uploaded by Gonçalo Machado Monteiro on 08 September 2017.

The user has requested enhancement of the downloaded file.


Journal of Luminescence ] (]]]]) ]]]–]]]

Contents lists available at SciVerse ScienceDirect

Journal of Luminescence
journal homepage: www.elsevier.com/locate/jlumin

Optical and spectroscopic properties of rare earth-doped


(80  x)TeO2–xGeO2–10Nb2O5–10K2O glasses
Gonc- alo Monteiro n, Yigang Li, Luı́s F. Santos, Rui M. Almeida
Departamento de Engenharia Quı́mica/ICEMS, Instituto Superior Técnico/ TULisbon, Av. Rovisco Pais, 1049-001 Lisboa, Portugal

a r t i c l e i n f o abstract

Article history: Erbium-doped germanotellurite glasses in the TeO2–GeO2–Nb2O5–K2O system have been prepared and
Received 16 May 2012 their absorption and photoluminescence properties were investigated. Judd–Ofelt theory was applied
Received in revised form to the different absorption intensities of the UV/vis Er3 þ transitions, to determine the intensity
2 August 2012
parameters, Ot (t ¼2, 4, 6), and to estimate the radiative lifetimes. However, for a single transition like
Accepted 17 August 2012
the Er3 þ emission to the ground state, at ca. 1.5 mm, the absorption cross-section method is shown to
be a simpler method to estimate the radiative lifetime of that particular transition. Experimental
Keywords: lifetimes were obtained from the photoluminescence decay curves of the 4I13/2 metastable state of Er3 þ
Erbium-doped glasses in different glasses and compared to the radiative lifetimes estimated by both methods. A better
Germanotellurite glasses
agreement was obtained by the absorption cross-section approach, yielding quantum efficiencies of
Photoluminescence
94% and 69% for 0.2 mol% and 2 mol% ErO1.5 doped samples, respectively. The up-conversion
Judd–Ofelt analysis
Absorption cross-section luminescence intensities of Er3 þ ions in these glasses were also analyzed in the 400–900 nm range
Quantum yield and compared with up-conversion in samples co-doped with Ytterbium and/or Thulium.
& 2012 Elsevier B.V. All rights reserved.

1. Introduction procedure. However, a simpler estimation method, the absorption


cross-section (ACS) method [12] can also be used, as long as the
Tellurite glasses doped with rare-earth ions have been widely terminal levels are ground states. This alternative method is a
researched in the past two decades as possible substitutes for convenient tool because it can also be applied for all PL processes
silica in applications like optical amplifiers, infrared and up- whose terminal levels are ground states and represents a much
conversion lasers, wavelength division multiplexing in fiberoptics quicker and simpler approach to calculate the Er3 þ emission rate
and waveguide laser systems [1,2]. These applications depend on from the 4I13/2 level to the 4I15/2 ground state.
the optical properties of rare-earth ions which, in turn, are In this study, a series of Er3 þ -doped germanotellurite glasses,
dependent on the host matrix and the ligand field around those from pure germanate to pure tellurite compositions, have been
ions. These two factors influence the optical absorption and the prepared and their absorption and photoluminescence properties
quantum efficiency of stimulated emissions [3,4]. were investigated. Experimental Er3 þ radiative lifetimes were
Tellurite glass matrices are known to present high rare-earth obtained and compared with the radiative lifetimes estimated by
ion solubility [5], extended infrared transmission [6] and low Judd–Ofelt analysis and also by the absorption cross-section
vibrational energies [7], which enhance the efficiency of Er3 þ method.
photoluminescence (PL) and up-conversion effects [8]; they also
possess good thermal and chemical stability and good fiber
drawing ability [9]. 2. Experimental
In Er-doped materials, the Er3 þ radiative lifetime at 1.5 mm is
an important parameter which is widely used to estimate the 2.1. Sample preparation
Er3 þ 1.5 mm photoluminescence quantum yield and to determine
the non-radiative relaxation rate of the Er: 4I13/2 level. A series of germanotellurite glasses with a nominal molar
Rare-earth ion radiative lifetimes are commonly estimated composition (80 x)TeO2–xGeO2–10Nb2O5–10K2O, where x ranges
by the Judd–Ofelt theory [10,11] based on their experimental from 0 to 80, hereafter named (8x)TxG, (x¼0,y, 8), were doped
absorption spectra, following a relatively complex calculation with 0.1 to 5% ErO1.5. Small amounts of YbO1,5 and TmO1,5 (up to
1.2 and 2 mol%, respectively) were also added in some cases. All
starting materials were reagent grade commercial powders with
n
Corresponding author. Tel.: þ351 218418137; fax: þ351 218418132. purity higher than 99.99%. The glasses were prepared by the melt-
E-mail address: goncalommmonteiro@ist.utl.pt (G. Monteiro). quenching technique, as previously described [7].

0022-2313/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jlumin.2012.08.031

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
2 G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]]

2.2. Optical and spectroscopic measurements where fed (aJ; bJ0 ) represents the oscillator strength from aJ to bJ0
states by the electric dipole mechanism; J and J0 are the total angular
The fundamental optical constants of the bulk glass samples momenta of the two states, respectively; a and b denote all other
were determined with a Horiba–Jobin Yvon spectroscopic quantum numbers; h is Planck’s constant; wed is the local field
ellipsometer (model UVISEL), at an incidence angle of 701, in correction factor for the electric dipole transition, which is usually
the 4.7–0.75 eV (265–1650 nm) spectral range, with a 0.005 eV approximated by wed ¼n(n2 þ2)2/9; Sed (aJ; bJ0 ) is a sum of the
increment. products of three phenomenological parameters Ot(2,4,6) and their
The Arquimedes principle was used to measure the densities corresponding matrix elements of unit tensor operators U(t) [13]:
of the glasses, with toluene (r ¼0.87 g/cm3) as the liquid medium. X
Sed ðaJ; bJ 0 Þ ¼ e2 Ot of N ½aSLJJU ðtÞ Jf N ½a0 S0 L0 J 0 4 2 ð3Þ
Bulk polished samples were analyzed by absorption spectro-
t ¼ 2,4,6
scopy, in the wavelength range between 200 and 1100 nm,
with an UV/Vis Thermo Helios a spectrophotometer, with 1 nm where 9f [aSL]J4 and 9fN[a0 S0 L0 ]J0 o are intermediate coupled states
N

resolution. of the rare earth ion. They are linear combinations of basic Russell–
For the near infrared (NIR) absorption range (770–3700 nm), a Sounders coupled states (or L–S coupled states).
Nicolet 5700 FT-IR equipment from Thermo Electron Corporation The values of the three host dependent parameters, Ot, are
was used, with a quartz beam splitter and a DTGS-TEC detector. usually estimated by solving a set of linear equations, for every
The samples, finely ground to avoid self-absorption effects, equation written for each absorption band:
were also irradiated with the 514.5 nm line of an Argon ion laser Z X
4p2 awed n  
(Spectra-Physics, model 2016) to excite the 4I13/2-4I15/2 transi- sðnÞdn ¼ Ot o f N ½aSLJJU ðtÞ Jf N a0 S0 L0 J 0 4 2
3n2 2J þ 1 t ¼ 2,4,6
tion of Er3 þ , in the range of  1450–1650 nm. The photolumines-
cence was measured with an InGaAs detector (NIR spectrograph ð4Þ
model CDINIR/256L, Control Development, USA) with 6 nm where a is the fine structure constant
resolution.
The fluorescence decay was also measured, by including a e2 1
a¼ ffi ð5Þ
rotating chopper in the setup, at a frequency of 30 Hz and an _c 137:036
InGaAs photodetector (GermaniumPower Devices Corp., model and s is the ACS of the transition between the ground state and
GAP1000), the signal being collected with a FLUKE PM 3370B the Er: 4I13/2 level.
oscilloscope.
ln½I0 ðnÞ=IðnÞ
The up-conversion effect in the co-doped samples was mea- sðnÞ ¼ ð6Þ
Nx
sured by irradiating the samples with a 975 nm laser (Lumics
High Power Module, model IPS-1820D), with an output power of where I0 and I are incident and transmitted light intensities,
0.44 W. The emitted signal was captured by an optical fiber, sent respectively, N is the ion concentration and x is the thickness of
to a 0.85 m double monochromator (Spex 1403) and finally to a the sample. If at least three absorption bands are observed,
photomultiplier tube detector (PMT, Hamamatsu, R928) being resulting from transitions in which only electric dipole mechan-
recorded between 400 and 900 nm. ism is involved, then the set of equations (Eq. (4)) is solvable.
For the magnetic dipole transition, the oscillator strength is
given by [14]:
3. Theory
8p2 mn
f md ðaJ; bJ 0 Þ ¼ w S ðaJ; bJ0 Þ ð7Þ
3hð2J þ 1Þe2 n2 md md
According to the published literature, researchers usually
estimate the radiative lifetime, t0, through the Judd–Ofelt theory where wmd is the local field correction factor for the magnetic
[10,11]. However, t0 can also be estimated by the absorption dipole transition, which is usually approximated by wmd ¼n3 [14].
cross-section (ACS) method, so the basics of both methods will be The sum of matrix elements of the magnetic dipole operator is
concisely reviewed here. N ! ! N
Smd ðaJ; bJ 0 Þ ¼ mB 2 o f ½aSLJJ L þ 2 S Jf ½a0 S0 L0 J0 4 2 ð8Þ
3.1. Judd–Ofelt theory where mB is the Bohr magnetron.
In the case of Er3 þ , the 4I13/2 to 4I15/2 transition involves
For an electronic transition of a rare earth ion, the reciprocal of both electric and magnetic dipole mechanisms. Therefore, in
the radiative lifetime is the spontaneous radiative emission rate, order to determine the radiative lifetime t0, of this transition by
A, which is given by the Judd–Ofelt theory, the absorption spectrum must be measured
first. Er3 þ absorption phenomena are due to electronic transitions
8p2 e2 n2 n2 from the Er3 þ ground level 4I15/2 to higher excited levels and, apart
A¼ f ð1Þ
mc3 from the 4I15/2 to 4I13/2 and to 2K15/2 transitions, the other low
energy absorption transitions only involve the electric dipole
where e is the electronic charge, n is the material refractive index, mechanism. Therefore, in principle, one can use the 11 absorption
n is the mean frequency of the transition, m is the electron rest bands in the range of 350–1100 nm to obtain 11 equations.
mass, c is the velocity of light in vacuum and f is the oscillator However, because of the UV absorption edge of the glass matrix,
strength of the transition. the interference from other co-doping ions or overlapping between
In resonance theory, the mechanisms of a transition include Er3 þ absorption bands, the number of obtainable equations is
electric dipole, magnetic dipole, electric quadrupole, etc. [13], lower. The three Ot parameters can only be determined if the
although the transition probabilities contributed by electric number of equations is at least 3. Then the value of Sed for the
4
quadrupole mechanism and above can be neglected [13]. For an I13/2-4I15/2 transition is calculated as
electric dipole transition:
Sed ð4 I13=2 ; 4 I15=2 Þ ¼ e2 ð0:0188O2 þ0:1176O4 þ 1:4617O6 Þ ð9Þ
2
8p mn in which the values of matrix elements of unit tensor operators are
f ed ðaJ; bJ 0 Þ ¼ w S ðaJ; bJ0 Þ ð2Þ
3hð2J þ 1Þe2 n2 ed ed given by Weber [13]. Then fed in Eq. (2) can also be calculated and

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]] 3

Eq. (1) is written as Ot values (since this involves several Er3 þ energy levels, the
method presents global accuracy, but it introduces errors for each
8p2 e2 n2 n2 8p2 e2 n2 n2 8p2 e2 n2 n2 1
A¼ f¼ ðf ed þf md Þ ¼ f ed þ Amd ¼ specific transition); the second error source is that all the
mc3 mc3 mc3 t0
intermediate coupled states of Er3 þ ions used in Eq. (3) and
ð10Þ
Eq. (8) are from LaF3:Er3 þ and not from the material under study.
where Amd is the spontaneous emission probability by magnetic- It is assumed that this will not introduce too much error because
dipole mechanism. the 4 fN electrons are shielded by 5d orbitals and thus are not
Amd can be deduced from the probabilities for spontaneous affected by the surrounding ions, but no one knows exactly the
emission of electric- and magnetic-dipole radiation in LaF3:Er3 þ , error introduced here. It is only a convenient way because
predicted by Weber (Table IV of Ref. [13]), where Amd for the intermediate coupled states come from very complicated fittings
4
I13/2-4I15/2 transition is 37.6 s  1y According to Eqs. (1), (7) and (8): according to high-resolution absorption spectra. The third advan-
tage of using Eq. (13) is that it can be used even when the three
Amd f md wmd n3
¼ ¼ ¼ ð11Þ Ot cannot be obtained. For instance, Yb3 þ ions only have two
Amd,LaF 3 f md,LaF 3 wmd,LaF 3 n3LaF 3
energy levels so that only one absorption band exists, which is not
Therefore: enough to solve for the Ot values. Using the ACS method,
 3  n 3 however, this absorption band is enough to estimate the radiative
n lifetime of the reverse transition
Amd ¼ Amd,LaF 3 ¼ 37:6 s1 ð12Þ
nLaF 3 1:60
where n is the refractive index of the material studied and 1.60 is the 3.3. Experimental lifetime
ordinary refractive index of LaF3 crystal, obtained from the dispersion
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2 To obtain the experimental lifetime (tm), it is necessary to
formula n0 ¼ ð1:5376l =l 0:0881Þ þ1, (0.35–0.70 mm) which
analyze the PL decay curve from the metastable state 4I13/2. The
was extrapolated for 1.54 mm [15,16]. The procedure outlined above decay curves can be approximately expressed by an exponential
is a route to estimate the Er3 þ 1.5 mm radiative lifetime t0 for an function as:
Er3 þ -doped material using the Judd–Ofelt theory.
It ¼ I0 expðt=tm Þ ð14Þ

3.2. Absorption cross-section method After fitting the decay curves, expressed by It(t) (signal
strength versus time), where tm is the time for which the
The Judd–Ofelt theory is a useful tool to predict various lumines- intensity has decreased to 1/e of the initial value I0, it is possible
cence properties originating from f–f electronic transitions. How- to extract the experimental lifetime.
ever, if one just wants to estimate the radiative lifetime t0 of a single Finally, the ratio between the experimental and theoretical
transition such as the Er3 þ 1.5 mm emission (Er3 þ : 4I13/2-4I15/2), lifetimes is the quantum efficiency (Q.E.) of the transition.
in which the terminal energy level is the ground state, a simpler way
to estimate t0 from the reverse absorption process can be used [12]:
Z 4. Results and discussion
1 8pn2 g 1
¼ 2
n2 sðnÞdn ð13Þ
t0 c g2 Almost all compositions yielded transparent glasses, but different
where g1 and g2 are the degeneracies of the ground and excited compositions showed different Er3 þ solubility. For instance, low
levels, respectively, and the other parameters are the same as tellurium content glasses doped with 2% ErO1,5 (3T5G, 2T6G, 1T7G)
previously defined. Eq. (13) is derived from the interrelationships presented a translucent or even opaque pink appearance, whereas
between the three Einstein coefficients for atomic spectral clear, transparent samples of compositions 6T2G and 8T0G were
phenomena [17]. The details can be found in Appendix A. One easily obtained with up to 5 and 4 mol% ErO1.5, respectively. The
can verify that Eq. (13) is consistent with Eqs. (1)–(8), which is co-doped samples were all transparent.
expected because Einstein coefficients are the basis of almost all
atomic absorption, luminescence and laser phenomena. The major 4.1. Physical properties of glasses
limitation of Eq. (13) is that it can only be used for the transitions
whose terminal energy level is the ground state and their (reverse) Fig. 1 shows the density and the refractive indices (at 514.5 nm,
absorption cross-sections must be measurable. Therefore, for those since this was the value considered in the Judd–Ofelt calculations)
transitions between any other two excited levels of rare-earth ions, of erbium-doped germanotellurite glasses with 0.2 and 2 mol%
the Judd–Ofelt theory is the only choice. In the case of the 1.5 mm ErO1,5. The presence of Er increased the density of the glasses as
transition of Er3 þ , (Appendix A—Derivation of Eq. (13)) the observed in Table 2. On the other hand, both the density and
integration in Eq. (13), where g1 ¼8 and g2 ¼7, must cover the refractive index decreased by 25% and 20% from pure tellurite to
whole Er3 þ absorption band peaking near 1.5 mm. pure germanate glasses, respectively. These results are similar, yet
There are several advantages in using this method to estimate more pronounced than those obtained in a previous study of a
the radiative lifetime for the transitions whose terminal energy similar germanotellurite series of undoped compositions [7].
level is the ground state. Firstly, it is easier, since one just needs to
measure the ACS for the reverse (absorption) transitions, while 4.2. Absorption spectroscopy
Judd–Ofelt theory requires a full analysis of all the f–f transitions
inside 4fN or 5fN configurations. The second is that, in principle, The absorption spectra recorded for the 2 mol% ErO1,5-doped
the result obtained from Eq. (13) should be more accurate, since it 0T8G, 4T4G and 8T0G glasses are shown in Fig. 2, in the visible
is based on a general relationship between spontaneous emission and NIR regions and depict the Er3 þ transitions that can be
rate and absorption intensity for any two energy levels. So, for the distinguished in the 350–3500 nm range for these samples.
single question of Er3 þ 1.5 mm radiative lifetime, the Judd–Ofelt According to Fig. 2 the bandgap energy increases with the
approach has at least two likely error sources: one is the non- germanium content of the glasses. This allows us to distinguish
unique way to obtain the three Ot parameters (Eq. (4)), where the 4G9/2 and 4G11/2 transitions for higher germanate content
three or more absorption bands are needed to fit the three compositions and, consequently, it makes possible for the rare

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
4 G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]]

Fig. 1. Density and refractive index of the germanotellurite series of samples doped with 0.2 and 2 mol% ErO1.5.

Fig. 2. Absorption coefficient data for 0T8G, 4T4G and 8T0G samples doped with 2 mol% ErO1,5, obtained through absorption measurements with an UV–vis-NIR
equipment (a) and IR absorption measurements with a FT-IR equipment in the NIR-MIR range (b).

earth ions to be pumped in UV region. Fig. 2 also shows that the Fig. 4 shows the absorption spectra of the low tellurium
width of the transmission window of the glasses is approximately content samples, doped with 2 mol% ErO1.5, which resulted in
3000 nm. non-transparent samples and the pure germanate composition.
The visible range of the absorption spectra (normalized at the The 1T7G composition presents low absorption peaks, while the
2
H11/2 transition—521.6 nm) of the 2 mol% ErO1.5 doped compo- 2T6G composition shows an increase in area for all the typical
sitions (except for 10 and 20 mol% TeO2) is presented in Fig. 3. It is transitions and at the same time a splitting of each transition
shown that the Er transitions 4F7/2 and 4F9/2 become more intense in sub-level transitions which results in an inhomogeneous
with increasing tellurite content, when compared with the 2H11/2 broadening of the peaks. This might be an indication that, in this
transition. sample, the erbium ions are in a more ordered, if not crystalline,

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]] 5

environment, however it was not possible to confirm this [9,18–22] O2 is strongly dependent on the asymmetry of the
assumption from XRD, since no peaks could be observed in the glass hosts and rare-earth’s environment, increasing with cova-
diffraction pattern (Fig. 5a). On the other hand, after a 10 h heat lent bond strength. In fact it was observed that this parameter
treatment, at the crystallization temperature of the sample increased with the increase of GeO2 content in tellurite glasses, in
(800 1C), the X-ray diffraction analysis revealed incipient crystal- agreement with the literature [23]. On the other hand, the
lization of an unknown crystalline phase (Fig. 5b). expected decrease of the O6 parameter [23], which is another
indication that the covalence of the rare earth–ligand bond
4.3. Judd–Ofelt parameters increased could not be confirmed in the present work.
Considering the calculated Judd–Ofelt parameters in Fig. 6, a
The experimental absorption spectra were used to calculate general increase of the O2 parameter can be observed with the
the phenomenological Judd–Ofelt parameters Ot, and the absorp- increase of the germanium content, presenting a maximum for
tion cross-sections. As referred previously by many authors the 2T6G sample. This result also occurs in the 2 mol% Er doped

Fig. 3. Absorption coefficient spectra, normalized to the intensity of the peak 2H11/2, for samples with 2 mol% ErO1,5 (excluding samples 1T7G and 2T6G) in the visible
range (475–700 nm). The inset graphic represents the evolution of the ratio between the areas of each absorption peak with respect to the peak 2H11/2 (the most intense
transition, common to all spectra) as a function of the tellurium content.

Fig. 4. Absorption coefficient of samples 1T7G and 2T6G, doped with 2 mol% ErO1.5, compared to the 0T8G þ 2 mol% ErO1.5 sample. (The baseline was subtracted).

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
6 G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]]

Fig. 5. X-ray diffraction patterns of the sample of composition 2T6Gþ2 mol% ErO1,5 before (a) and after (b) a heat treatment at 800 1C.

Fig. 6. Judd–Ofelt parameters Ot (t¼ 2,4,6) calculated for ErO1.5 doped samples as a function of the Erbium concentration (for samples of composition 6T2G) and as a
function of the TeO2 content (for 0.2 mol% Er content).

sample, where the 2T6G also presents the higher O2 value of the The values obtained for this parameter, for the 2T6G composi-
series (Table 1), reflecting a higher asymmetry of the glass matrix. tion, are significantly higher than for other germanate or tellurite
Therefore, this composition seems to be a good host for erbium compositions, as observed in Table 2.
ions since it favors a crystalline environment, perhaps due to a
possible phase separation. In fact, in a previous work [7] it was 4.4. Absorption cross-section
observed that germanate and tellurite coordination spheres do
not mix and that the Raman signature of the TeO3 units are Considering the NIR absorption spectra of the 2 mol% ErO1.5-
present even for concentrations of TeO2 as low as 10 mol %, doped compositions in terms of absorbance or optical density
suggesting the occurrence of phase separation. (OD), the concentration of dopant (N3)(Erþ ) and the thickness of the

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]] 7

samples (l), the respective absorption cross-sections of 4I13/2 level emission decay and, consequently, the radiative efficiency of the
(sa) can be obtained: erbium ions.
Fig. 9(a) also shows that the lifetime of the PL emission
ODðlÞ
sa ¼ ð15Þ increases for increasing germanium content. When in glass
N ðEr3 þ Þ l
matrices, the small differences between each erbium environ-
It can be seen that 1T7G and 2T6G compositions have a ment cause a shift in energy in each one of the 56 possible
broader and more intense ACS spectra (Fig. 7), resulting in lower transitions between all sub-levels of the 4I13/2-4I15/2 transition.
estimated values for the correspondent transition lifetimes. As for This means that, for more amorphous matrices, there is less
the other compositions we can see that the ACS area decreases overlapping of each of the sub-level transitions, which results in
from pure tellurite to pure germanate composition and is inver- an inhomogeneous broadening of the emission peak. Fig. 9(a)
sely related to the respective lifetimes, as seen in Table 1. indicates that with increasing germanium content the erbium
environment becomes less ordered inducing an increase in the
FWHM of the emission peak.
4.5. Photoluminescence
Table 1 contains all the measured and calculated lifetimes, as
well as the respective quantum yields.
Fig. 8 presents the normalized spectra for the 4I13/2-4I15/2
The experimental lifetimes were obtained by fitting a single
emission for the 2 mol% ErO1.5 compositions. It can be seen that
exponential to the decay curves, like the ones presented in Fig. 10.
the main peak, located at  1529 nm, remains practically in the
For these measurements, powdered samples were used to avoid
same position while the shoulders at  1485 and at  1550 nm,
self-absorption effects, which can influence the experimental
deviate from each other with increasing germanium content. On
lifetime. In case of bulk samples the re-absorption of early
the other hand, a second shoulder at 1540 nm becomes more
emitted photons will in fact lengthen the emission process.
visible. It is also perceptible that the ratio between the main peak
As observed in Fig. 10, the PL lifetime decreases with increas-
and the shoulders decreases with increasing germanium content.
ing erbium concentration. According to Stokowski et al. [24,25]
These results lead to higher spectral bandwidth for the higher
this evolution can be described by the following expression:
GeO2 content compositions, as confirmed from the FWHM data
plotted in Fig. 9(a). t*
Fig. 9(b) illustrates, for the 8T0G composition, the variation of texp ¼ ð16Þ
1 þ ðr=Q Þp
the FWHM with the erbium content. As expected, the more rare-
earth ions are incorporated in the matrix, the broader the FWHM Where texp is the lifetime obtained experimentally, t* is the
of the 1.54 mm PL signal. On the other hand, the increase of the lifetime when the concentration tends to zero, r is the erbium
erbium concentration naturally tends to bring closer neighboring concentration, p is a phenomenological parameter characterizing
erbium ions. Therefore, the shorter the distance between these how fast the lifetime decreases with the erbium concentration
ions, the more intense become the electric dipole interactions, as and Q is the quenching concentration. Fig. 11 presents the
it depends on the sixth power of the distance. On these condi- experimental lifetime as a function of erbium concentration for
tions, phenomena like energy migration and cooperative up- the 6T2G and 8T0G compositions. From the inset tables it is
conversion can occur, lowering the experimental lifetime of the observed that the ideal lifetime is lower for the pure tellurite

Table 1
Glass sample labels, full width at half maximum (7 0.5 nm) of the Er3 þ : 4I13/2-4I15/2 PL emission band, concentration of Er3 þ ions per cubic centimeter of glass, Judd–
Ofelt intensity parameters, experimental lifetimes ( 7 0.05 ms) of the PL emission, radiative lifetimes according to Judd–Ofelt theory and absorption cross-section
approach, and quantum efficiencies according to both approaches.

Samples FWHM-emission N3Erþ Judd–Ofelt Parameters Lifetime (ms) Quantum efficiency


(nm) (  1020 cm  3) (  10  20 cm2)

X2 X4 X6 Radiative ACS method Experimental Radiative ACS method


(J–O theory) (J–O theory)

6T2G þ0.1Er 51 0.189 6.4 1.6 0.6 3.2 2.5 2.9 92 118
0T8G þ 0.2Er 53 0.388 8.7 1.2 0.8 6.1 4.9 3.5 58 72
1T7G þ0.2Er 51 0.387 10.7 1.5 0.9 5.1 4.7 3.6 72 78
2T6G þ0.2Er 50 0.387 12.4 1.9 1.0 4.5 4.2 3.7 83 90
3T5G þ0.2Er 50 0.384 8.2 1.8 0.9 4.6 3.8 3.5 75 92
4T4G þ0.2Er 50 0.381 7.6 1.5 0.9 4.0 3.3 2.9 73 88
5T3G þ0.2Er 48 0.379 6.8 1.5 0.6 4.6 3.3 3.1 68 93
6T2G þ0.2Er 50 0.376 6.9 1.3 0.9 3.7 3.4 2.9 80 85
7T1G þ0.2Er 49 0.371 8.0 1.6 1.0 3.7 3.4 2.7 73 80
8T0G þ 0.2Er 47 0.367 5.9 1.1 0.9 3.1 2.8 2.6 84 94
6T2G þ1Er 55 1.869 6.2 1.3 0.9 2.8 2.7 2.5 90 93
8T0G þ 1Er 48 1.484 5.9 1.4 0.9 3.2 2.3 2.5 80 110
0T8G þ 2Er 55 3.847 8.1 1.8 1.1 4.9 4.7 1.9 39 41
1T7G þ2Er 59 3.809 0.9 0.1 0.1 14.8 0.9 2.8 19 309
2T6G þ2Er 58 3.801 9.3 4.6 4.1 1.5 2.2 2.4 158 109
3T5G þ2Er 58 3.784 7.6 1.4 1.1 4.1 3.7 2.1 51 57
4T4G þ2Er 48 3.749 2.6 1.7 1.0 3.8 3.3 1.3 34 40
5T3G þ2Er 54 3.727 1.6 1.0 0.6 5.2 3.5 1.6 30 45
6T2G þ2Er 54 3.707 6.1 1.2 0.9 3.5 3.2 2.0 56 63
7T1G þ2Er 56 3.650 7.4 1.6 1.1 3.0 3.0 2.0 68 69
8T0G þ 2Er 52 3.606 6.5 1.4 1.0 3.2 2.9 2.0 61 68
6T2G þ5Er – 8.988 – – – – 6.24 1.22 – 20
8T0G þ 4Er 74 7.097 – – – – 4.30 1.52 – 35

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
8 G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]]

sample and that these compositions have an erbium ion solubility

[23]
[18]

[19]
[33]

[34]
[35]
[20]
Ref.

[9]
of around 4 and 5 mol%, corresponding to a quenching concentra-
tion of 6.7  1020 and 8.4  1020 erbium ions/cm3, for the 6T2G

o78%
se*(  10  21 cm2) Z (J–O

and 8T0G compositions, respectively.

99.4%
35.9–
anls.)

Table 1 reveals that, from the absorption cross-section method,







the obtained radiative lifetime results correlate better among each
other and also to most experimental results, leading to higher and
more homogeneous quantum yield values, which for the 0.2 mol%

10.1–12.2
7.2–10.7

ErO1.5 series varied between 72% and 94%, when compared with

7.6–7.8
7.1–9.0
those calculated from Judd–Ofelt theory.



(  10  20 cm2) (  10  20 cm2) (  10  20 cm2)

4.6. Up-conversion

The up-conversion of Er3 þ -doped and Yb3 þ /Tm3 þ co-doped


1.3–1.4
0.8–1.4
1.0–1.2

0.9–1.2

samples was also studied. From the up-conversion signals of


1–2
1

2.3
O6


erbium-doped samples, presented in Fig. 12, one observes that


Physical, optical and spectroscopic properties of doped-glass compositions with one or more of the constituents used in the glass compositions analyzed in this paper.ss

the emission from 4F7/2 to the ground state (  490 nm), logically
expected from the frequency doubled transition 4I15/2-4I11/2-
 1.5–2
1.5–2.6
1.5–1.6

1.4–2.5

0.8–2.0

4
F7/2 (a ground state absorption – GSA followed by an excited
2–2.3
2.8

state absorption – ESA), does not occur and that a non-radiative


O4


decay immediately happens from 4F7/2 levels to 2H11/2, 4S3/2 and


4
F9/2 levels. The green and red emissions (530/550 nm and
660 nm wavelengths, respectively) seen in the up-conversion
 4.5–7
3.5–5.8

5.8–7.3

5.8–6.3
6.4–9.0

6–9.2

spectrum, result from the radiative decay from these three levels
7.4
O2

to the ground state [20]. It is also observed, for glasses 6T2G, 7T1G

and 8T0G with 2 mol% Er, that the intensity of the transition
emission of Er3 þ:4I13/

2
H11/2-4I15/2 increased with increasing tellurite content, with
respect to the other transitions. For the co-doped samples,
ytterbium and thulium seem to play an important role in the
up-conversion luminescence process: thulium broadens the
total emission bandwidth by adding emissions in the blue and
NIR regions of the spectrum without affecting to much the green
and red emissions from erbium; while Yb3 þ is known to act as a
obs (ms)

2- I15/2

2.2–4.1
3.2–6.0

0.1–0.5

sensitizer of both erbium and thulium ions as shown in the


4

3.4
1.0

diagram on Fig. 13. For the present study we have considered a




t

Yb3 þ :Er3 þ ratio of 6, because higher ratios have been demon-


strated to quench the energy transfer from Yb3 þ to Er3 þ dopants,
FWHM

38–52

67–85
53–55
65–72
(nm)

but at the same time a ratio of 6 is high enough to maintain


68
69

an equilibrium between the population of the Er3 þ :4I11/2 and



Yb3 þ :2F5/2 levels that ensures a high efficiency of energy transfer


(g cm  3) 633 nm)

between the two [26,27]. The Tm3 þ :Er3 þ ratio was 10 because it
1.8–2.1

1.5–1.7
1.5–2.0
1.9–2
n (at

has been shown that higher concentration ratios of Tm3 þ :Er3 þ


2.1
1.7

1.7
1.6

lead to a broader range of the up-conversion luminescence in the


4.8–5.4
2.8–4.8

3.1–4.3

3.8–4.1

red spectral region ( 650–725 nm) [28].


4.0–5.0

The presence of Yb3 þ ions, with a 6:1 Yb3 þ to Er3 þ ratio,


4.4
3.8
r


results in strong absorption of the 975 nm pump laser beam,


70(TeO2  GeO2)  5 K2O  15(Na2O  Nb2O5)  10ZnOþ 0.4  0.6%ErO1.5

mostly through the GSA absorption 2F7/2-2F5/2 of Yb3 þ , as shown


in Fig. 13, followed by the energy transfer step from Yb3 þ to Er3 þ
(Yb3 þ :2F5/2 þEr3 þ :4I15/2-Yb3 þ :2F7/2 þEr3 þ :4I11/2); but also from
99.5(15Ga2O3_75GeO2_10R2O) _1%ErO1.5 (R ¼ Li, Na, K)

the Er3 þ 4I15/2-4F9/2 GSA that already occurred in erbium-doped


samples.
It is seen in Fig. 12, for 6T2G composition co-doped with
erbium and ytterbium, that the red emission is intensified by the
80(TeO2.NaPO3)  10ZnO  10Na2Oþ 1%ErO1.5

80(GeO2  B2O3)  10ZnO  10BaO þ 1%ErO1.5

presence of Yb3 þ ions. The Er3 þ :4F9/2 level now becomes popu-
70TeO2  30(ZnO  Nb2O5  BaO) þ2%ErO1.5
60TeO2  20ZnO  20ZnCl2 þ1  10%ErCl3

lated through two other mechanisms: a two photon ESA mechan-


20TeO2  74.5B2O3  5Li2Oþ 1%ErO1.5

ism followed by a non-radiative decay and a two photon energy


20TeO2  71B2O3  5Li2Oþ 8%ErO1.5

transfer process from Yb3 þ to Er3 þ ions, with an intermediate


20TeO2  80GeO2 þ0.5%ErO1.5

non-radiative decay (4F11/2-4F13/2), as demonstrated by dashed


90TeO2  10GeO2 þ1%ErO1.5

lines in Fig. 13.


For the sample co-doped with erbium and thulium, a decrease
in the intensity of the red emission plus an increase of the green
emission (530 nm) is observed, together with the appearance of a
NIR emission (  825 nm), when compared to the erbium-doped
Samples

6T2G sample. The mechanisms behind the up-conversion emis-


Table 2

sions for this sample could tentatively be explained by a cross


energy transfer (ET): the green and red emissions are still assured

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]] 9

Fig. 7. Absorption cross-section of the NIR 4I13/2-4I15/2 transition for the 2 mol% ErO1.5-doped glass compositions.

Fig. 8. Photoluminescence emission spectra of the 2 mol% ErO1,5-doped samples (in powder form) irradiated with 514.5 nm laser line.

by the erbium ions from the two photon ESA mechanism, this stage the Tm3 þ ions are re-excited to the Tm3 þ :1G4 level via
followed by the respective non-radiative decay, but two other ESA and then radiatively decay to the Tm3 þ :3H5 level [29] or
cooperative processes take place. Now, the energy of part of the undergo another ET to the Er3 þ ions (4F7/2 level). These two
electrons populating the Er3 þ :4F9/2 level is transferred to the cooperative processes may respectively explain the appearance
thulium ions through a transition to Tm3 þ :3F3 level, but then an of the 825 nm low intensity broad band and the increase in
immediate non-radiative decay to the Tm3 þ :3H4 level occurs. intensity of the 530 nm band with respect to the green band at
This justifies the decrease in intensity of the red emission. From 550 nm.

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
10 G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]]

Fig. 9. Full width at half maximum (FWHM) of the Er: 4I13/2 normalized PL spectra of the glasses excited at 514.5 nm and corresponding lifetimes: (a) as a function
of tellurium content for samples with 0.2 mol% ErO1.5; (b) as a function of the erbium concentration for samples of composition 8T0G.

Fig. 10. PL decay curves for the emission from the 4I13/2 metastable state of Er3 þ ions in samples 6T2G and 8T0G upon 514.5 nm excitation, detected at 1528 nm.

The 790 nm very intense band, observed for the 6T2G sample Er3 þ :4F7/2 level must not be neglected, since a significant increase
co-doped with Er, Yb and Tm, is due to the Tm3 þ :3H4-3H6 of the 530 nm emission (2H11/2-4I15/2) is observed.
transition (which in fact has a slightly higher energy than the
Tm3 þ :1G4-3H5). Ytterbium plays here a cooperative role with
Tm3 þ ions by populating the 3F2 level in a two photon absorption 5. Conclusion
process with intermediate non-radiative decays, resulting in the
low intensity 700 nm band, which can be justified by the Niobium germanotellurite glasses doped with 0.1–5 mol% of
transition 3F2-3H6 [30]. A third ET from Yb3 þ excites the Tm3 þ ErO1.5 have been prepared and characterized. Their density and
ions from the 3H4 to the 1G4 level and, from this, a radiative refractive indices increased by 25% and 20%, respectively, when
transition to the ground level takes place, giving rise to the blue the TeO2 content increased from 0 to 80 mol%. The absorption
band at  480 nm [31]. Despite this, the ET from Tm3 þ :1G4 to the spectra showed that the areas of the 4F7/2 and 4F9/2 absorption

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]] 11

Fig. 11. Experimental lifetime as a function of the erbium concentration for samples 6T2G and 8T0G, and fitting results from of the empirical Eq. (16). y refers to texp,
A1 to tn, r to x and x0 to Q from Eq. (16).

Fig. 12. IR-to-visible up-conversion luminescence spectra of Er3 þ , Yb3 þ and/or Tm3 þ co-doped glasses under 980 nm laser excitation.

peaks increased with increasing tellurite content and the inten- The quenching concentration was estimated at 6.7  1020 and
sity of the former peak was also influenced by the erbium 8.4  1020 erbium ions/cm3, for the 6T2G and 8T0G compositions,
concentration. The emission spectra revealed an increase of the respectively, corresponding to 4 and 5 mol% ErO1.5 respectively.
spectral bandwidth with increasing GeO2 content, translated into Based on the ACS approach, the estimated lifetimes correlated
a full width at half maximum increase from 47 to 53 nm, for better among each other and also with most experimental results
0.2 mol% ErO1.5; on the other hand, a FWHM of 74 nm was than those calculated from Judd–Ofelt theory. The resultant
registered for the composition 8T0G doped with 5 mol% ErO1.5. quantum yields varied between 72–94% for the glass samples
The measured lifetimes of the glasses decreased with increas- doped with 0.2 mol% Er.
ing erbium and tellurium oxide contents, ranging from 2.7 ms to The up-conversion luminescence of Er3 þ -doped glasses and
1.5 ms for the 8T0G glass doped with 0.2–5 mol% Er, respectively, glasses co-doped with Yb3 þ and/or Tm3 þ , was studied. It was
and from 3.7 to 2.6 ms when the TeO2 content increased from 0 to found that the green luminescence of the 2 mol% Er3 þ -doped
80 mol%, respectively, for glasses doped with 0.2 mol% ErO1.5. glasses increased with their Te content. For the co-doped samples,

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
12 G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]]

the luminescence profile, which obeys the normalization condi-


R
tion f (n)d n ¼1; r(n) is the radiation density, which has a
relationship:
n
rðnÞ ¼ In ðnÞ ðA  3Þ
c
Inserting Eq. (A-1) into Eq. (A-2) and rearranging the equation
as the right term of Eq. (A-1) one will obtain:
1 dIn ðnÞ hnn
sðnÞ ¼  ¼ B12 fðnÞ ðA  4Þ
NIn ðnÞ dx c
According to the interrelationship between Einstein coeffi-
cients:
A21 8phn3 n3
¼ ðA  5Þ
B21 c3
and
B21 g
¼ 1 ðA  6Þ
B12 g2
where A21 is the spontaneous emission rate, also called the
Einstein coefficient for spontaneous emissions. Inserting Eq. (A-3)
and Eq. (A-4) into Eq. (A-2), one will have:
8pn2 n2 g 1
A21 fðnÞ ¼ sðnÞ ðA  7Þ
c2 g 2

Fig. 13. Simplified diagram showing the most important transitions involved Finally, upon integration of Eq. (A-5) in the whole range of the
between erbium, ytterbium and thulium ions (adapted from [30,32]). absorption:
Z Z
1 8pn2 g 1
¼ A21 ¼ A21 fðnÞdn ¼ 2
n2 sðnÞdn ðA  8Þ
on the other hand, Yb3 þ acted as a sensitizer for Er3 þ emission, t0 c g2
causing an increase of the red up-conversion, while Tm3 þ i.e., Eq. (13) is obtained.
increased the intensity of the 530 nm green emission band
relative to the 550 nm band. Combining the three rare-earth
References
elements two additional bands: one small peak in the blue region
(475 nm) and one intense NIR band near 800 nm.
[1] A.P. Caricato, M. Fernandez, M. Ferrari, G. Leggieri, M. Martino, M. Mattarelli,
M. Montagna, V. Resta, L. Zampredi, R.M. Almeida, M.C. Concalves, L. Fortes,
L.F. Santos, Mat. Sci. Eng. B-Solid 105 (2003) 65.
Acknowledgments [2] A. Mori, J. Ceram. Soc. Jpn. 116 (2008) 1040.
[3] A. Terrasi, F. Priolo, G. Franzo, S. Coffa, F. D’Acapito, S. Mobilio, J. Lumin.
80 (1998) 363.
This research was developed under projects PTDC/CTM/64235/ [4] M.E. Fermann, I. Hartl, IEEE J. Sel. Top. Quantum Electron. 15 (2009) 191.
2006 and PTDC/CTM/101154/2008, both funded by the Fundac- a~ o [5] T. Xu, X. Zhang, G. Li, S. Dai, Q. Nie, X. Shen, X. Zhang, Spectrochim. Acta Part A:
para a Ciência e a Tecnologia (FCT) and by FEDER, under the Mol. Biomol. Spectrosc. 67 (2007) 559.
[6] G.N. Wang, S.Q. Xu, S.X. Dai, J.J. Zhang, L.L. Hu, Z.H. Jiang, Mater. Lett.
Program POCI 2010. One of the authors, Y.G. Li, would also like to 59 (2005) 366.
thank FCT for support under an Assistant Researcher contract at [7] G. Monteiro, L.F. Santos, J.C.G. Pereira, R.M. Almeida, J. Non-Cryst. Solids 357
IST (Instituto Superior Técnico). (2011) 2695.
[8] Z.D. Pan, S.H. Morgan, K. Dyer, A. Ueda, H.M. Liu, J. Appl. Phys. 79 (1996)
8906.
[9] M. Mattarelli, A. Chiappini, M. Montagna, A. Martucci, A. Ribaudo,
Appendix A M. Guglielmi, M. Ferrari, A. Chiasera, J. Non-Cryst. Solids 351 (2005) 1759.
[10] B.R. Judd, Phys. Rev. 127 (1962) 750.
A-derivation of Eq. (13) [11] G.S. Ofelt, J. Chem. Phys. 37 (1962) 511.
[12] Y.G. Li, H. Yang, Z. He, L.Y. Liu, W.C. Wang, F.Y. Li, L. Xu, J. Mater. Res.
A differential form of Eq. (6) can be written as 20 (2005) 2940.
[13] M.J. Weber, Phys. Rev. 157 (1967) 262.
1 dIn ðnÞ
sðnÞ ¼  ðA  1Þ [14] M.D. Shinn, W.A. Sibley, M.G. Drexhage, R.N. Brown, Phys. Rev. B 27 (1983)
NIn ðnÞ dx 6635.
[15] P.D. Marvin J. Weber, Handbook of Optical Materials, CRC Press, 2003.
where In(n)¼dI(n)/dn, is the light intensity in a unit frequency [16] R. Laiho, M. Lakkisto, Philos. Mag. B 48 (1983) 203.
interval. The number of photons in the unit frequency interval [17] G. Liu, B. Jacquier, Spectroscopic Properties of Rare Earths in Optical
that are absorbed by a two-level system when light propagates a Materials, Springer, 2005.
[18] L.M. Fortes, L.F. Santos, M.C. Goncalves, R.M. Almeida, M. Mattarelli,
small distance dx is M. Montagna, A. Chiasera, M. Ferrari, A. Monteil, S. Chaussedent, G.C. Righini,
Opt. Mater. 29 (2007) 503.
dIn ðnÞ
 ¼ ðB12 N1 B21 N2 ÞfðnÞrðnÞdx  B12 N fðnÞrðnÞdx ðA  2Þ [19] Y.M. Yang, Z.P. Yang, B.J. Chen, P.L. Li, X. Li, Q.L. Guo, J. Alloys Compd. 479
hn (2009) 883.
[20] Y.M. Yang, Z.P. Yang, P.L. Li, X. Li, Q.L. Guo, B.J. Chen, Opt. Mater. 32 (2009)
where B12 is the Einstein coefficient for photo absorption and B21
133.
is the Einstein coefficient for stimulated emission; N1 is the ion [21] P.C. Becker, N.A. Olsson, J.R. Simpson, Erbium-doped Fiber Amplifiers:
concentration in the ground level and N2 is the ion concentration Fundamentals and Technology, Academic Press, San Diego, 1999.
in the excited level; here it is assumed that N1 EN and N2 E0, [22] S. Tanabe, J. Non-Cryst. Solids 259 (1999) 1.
[23] C. Zhao, G.F. Yang, Q.Y. Zhang, Z.H. Jiang, J. Alloys Compd. 461 (2008) 617.
which is sufficiently accurate in most cases, according to the [24] S.E. Stokowski, L. Cook, H. Mueller, M.J. Weber, J. Lumin. 31–32 (2) (1984)
Boltzmann distribution; f(n) is a normalized function describing 823.

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031
G. Monteiro et al. / Journal of Luminescence ] (]]]]) ]]]–]]] 13

[25] S.E. Stokowski, R.A. Saroyan, M.J. Weber, Nd-doped Laser Glass Spectroscopic [30] H. Hu, Y. Bai, J. Alloys Compd. 527 (2012) 25.
and Physical Properties, Lawrence Livermore National Laboratory, Livermore, [31] S.Q. Xu, H.P. Ma, D.W. Fang, Z.X. Zhang, Z.H. Jiang, Mater. Lett. 59 (2005)
California, 1981. 3066.
[26] V.K. Tikhomirov, V.D. Rodrı́guez, J. Méndez-Ramos, J. del-Castillo, [32] S. Liang, Y. Liu, Y. Tang, Y. Xie, H.Z. Sun, H. Zhang, B. Yang, J. Nanomater.
D. Kirilenko, G. Van Tendeloo, V.V. Moshchalkov, Solar Energy Mater. Solar (2011).
Cells 100 (2012) 209. [33] J. Coelho, J. Azevedo, G. Hungerford, N.S. Hussain, Opt. Mater. 33 (2011) 1167.
[27] R. Wang, Y.S. Zhang, J.C. Sun, L. Liu, Y.L. Xu, J. Rare Earth 29 (2011) 826. [34] D.M. Shi, Y.G. Zhao, X.F. Wang, G.H. Liao, C. Zhao, M.Y. Peng, Q.Y. Zhang, Phys.
[28] D.C. Zhou, Z.G. Song, G.W. Chi, J.B. Qiu, J. Alloys Compd. 481 (2009) 881. B 406 (2011) 628.
[29] Y. Gandhi, M.V.R. Rao, C.S. Rao, I.V. Kityk, N. Veeraiah, J. Lumin. 131 (2011) [35] D.D. Chen, Y.H. Liu, Q.Y. Zhang, Z.D. Deng, Z.H. Jiang, Mater. Chem. Phys.
1443. 90 (2005) 78.

Please cite this article as: G. Monteiro, et al., J. Lumin. (2012), http://dx.doi.org/10.1016/j.jlumin.2012.08.031

View publication stats

Вам также может понравиться