Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Luminescence 211 (2019) 437–445

Contents lists available at ScienceDirect

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

Optical transition probabilities of white light emitting Sr2SiO4:Dy3+ T


nanophosphors for lighting applications using Judd−Ofelt analysis
C. Manjunatha,∗, M.S. Rudreshaa, B.M. Walshb,∗∗, R. Hari Krishnac, B.M. Nagabhushanac,
B.S. Panigrahid
a
Department of Physics, Sri Siddhartha Institute of Technology, Tumakuru, 572105, India
b
NASA Langley Research Center, Hampton, VA, 23681, USA
c
Department of Chemistry, M S Ramaiah Institute of Technology, Bengaluru, 560054, India
d
Radiological Safety Division, Indira Gandhi Centre for Atomic Research, Kalpakkam, 603102, India

ARTICLE INFO ABSTRACT

Keywords: Dysprosium doped strontium silicate (Sr2-xSiO4:xDy3+, x = 0–7 mol%) are prepared by solution combustion
Phosphors synthesis and their luminescent behaviour is investigated. X-ray diffraction (XRD) reveals the structure to be
Sr2SiO4 (strontium silicate) orthorhombic in nature and different structural parameters are calculated. The photoluminescence excitation
Photoluminescence and emission spectra are thoroughly discussed. The emission spectra excited at 350 nm showed the characteristic
Judd-Ofelt analysis
blue and yellow emission, the admixture of which gives probable white emission. Further, the absorption
spectrum is used to systematically study the spectral intensities of Sr2-xSiO4:xDy3+ (x = 0.008) nanophosphors
using Judd-Oflet (JO) theory. The JO parameters thus obtained followed the trend Ω2 > Ω6 > Ω4 that revealed
the field surrounding the ions lack inversion symmetry and is also confirmed by asymmetric ratios. JO para-
meters are also used to study different radiative properties. The magnitude of the calculated peak emission cross
section for the transition 4F9/2 → 6H13/2 is 1.02 × 10−20 cm2, a value suitable for laser applications. An energy
transfer quenching mechanism indicates that the mode of non-radiative transition is dominated by dipole-dipole
interactions. Color coordinates and color correlated temperature with respect to different Dy3+ concentrations
are studied in detail.

1. Introduction degradation rate of the phosphors used, resulting in poor chromatic


aberration and poor white light performance after long period of op-
Inorganic luminescent phosphors doped with rare earth ions (REI) eration [1]. Hence, developing a single phased phosphor that can emit
are suitable in numerous applications of scientific importance relating white light when exited by UV LED to replace YAG:Ce3+ yellow
to artificial production of light. They find application in phosphor phosphor is of prime importance [7]. In this context, triply ionized
converted white light emitting diodes (WLED's), display devices, active Dy3+ (Dysprosium) ions doped into nanophosphors plays a vital role in
laser materials, solid state lighting etc. [1–3]. Nowadays, WLED's have acquiring white light by tuning the yellow to blue intensity ratios (Y/B).
replaced conventional incandescent and fluorescent lights that com- Usually, Dy3+ are characterised by two emissions, blue (460–500 nm)
prise approximately 20–30% of all electrical energy consumption. and yellow (560–600 nm) corresponding to transitions 4F9/2 → 6H15/2
Moreover, the WLED's are more consistent, have high luminous effi- and 4F9/2 → 6H13/2, respectively. Therefore, color tuneable emissions
ciency and requires ten times less energy compared to conventional are expected to give the desired pure white light by adjusting the host
lighting, and hence leads to significant energy savings in solid state composition to change the Y/B intensity ratio [8]. In order to get better
lighting [4,5]. Basically two different techniques are available to gen- luminescent efficiency and chemical stability the role of choosing a
erate white light from LED's. The first is to use multiple LED's composed suitable host is important. In this perspective silicate based nanopho-
of complementary colors (red, blue and green), and the second is to sphors can be chosen as the appropriate host, as opposed to sulphide,
excite yellow phosphor (YAG: Ce3+) materials with blue LED chips [6]. phosphate, aluminate and oxide counterparts due to its high physical
However, the drawback of these techniques is the individual stability and water resistant property [9–11].


Corresponding author.
∗∗
Corresponding author.
E-mail addresses: manjunathc51085@gmail.com (C. Manjunath), brian.m.walsh@nasa.gov (B.M. Walsh).

https://doi.org/10.1016/j.jlumin.2019.03.054
Received 26 December 2018; Received in revised form 14 February 2019; Accepted 25 March 2019
Available online 01 April 2019
0022-2313/ © 2019 Elsevier B.V. All rights reserved.
C. Manjunath, et al. Journal of Luminescence 211 (2019) 437–445

Several reports are being published on silicate nanophosphor owing


to its favourable luminescent performances. Thus, this article is dedi-
cated to the study of luminescent properties of Dy3+ doped strontium
silicate (Sr2SiO4) nanophosphor. Even though there exists a consider-
able amount of work in the literature [12–22], so far no reports has
been published that is dedicated to the detailed investigation of spectral
intensities using absorption spectrum by means of Judd-Ofelt (JO)
theory. In the present work, Dy3+ doped Sr2SiO4 nanophosphor have
been prepared by solution combustion synthesis method and their
structural and luminescent properties were investigated. Further, dif-
ferent radiative properties were evaluated from JO analysis to check the
applicability of the phosphors in various technological applications.

2. Experimental

2.1. Synthesis of phosphor

Sr2-xSiO4:xDy3+ (x = 0.002, 0.004, 0.006, 0.008, 0.01, 0.03, 0.05


and 0.07) nanophosphors were prepared by well-known solution com-
bustion synthesis [23,24]. The materials used for the preparation were
Strontium nitrate (Sr(NO3)2, 99.9%, Merck Ltd.), fumed silica (SRL
chemicals), Dysprosium oxide (Dy2O3, 99.99%, Sigma Aldrich Ltd.),
glycine (C2H5NO2, 99.5%, SRL chemicals) and nitric acid (HNO3, Merck
Ltd.). Initially Dy2O3 was taken in crystallising dish and was treated
with 10 ml of 1:1 HNO3 and then heated on a sand bath and evaporated
to dryness to get Dy(NO3)2. Further stoichiometric amounts of Sr
(NO3)2, fumed silica and glycine were then transferred to crystallizing
dish and dissolved in 50 ml of deionised water. Measures were taken to
maintain the oxidizer to fuel ratio to be unity. The aqueous solution so
obtained was stirred well until a clear and transparent solution and is
kept in pre-heated furnace maintained at 500 ± 10 °C. The solution
was thermally dehydrated followed by sputtering with the liberation of
gases. The reaction process took around 10 min to complete and ag-
glomerated white powders were obtained. The resulting powders were Fig. 1. XRD patterns of Sr2-xSiO4:xDy3+ (x = 0 to 0.07) nanophosphors.
ground and were annealed at 900 °C for 3 h.
K
D=
2.2. Characterisation techniques cos (1)

where β is full width half maxima (FWHM), K is shape factor and is


The phase determination was done by X-ray diffraction (XRD) using the wavelength of X-ray.
Rigaku's x-ray diffractometer (Smartlabs, Japan) equipped with Cu-kα The estimation of crystallite size by considering stain (ε) broadening
radiation (1.5406 Å). The operating voltage and current were set at is done by Williamson-Hall (WH) relation and is shown in Fig. 2 [26];
30 kV and 40 mA, respectively. Transmission electron microscopy
imaging was recorded using a JOEL/JEM 2100 microscope operating at 0.9
cos = + 4 sin
200 kV. The luminescence spectra and decay lifetime measurements D (2)
were recorded using an Edinburgh instrument F-920 equipped with Equation (2) represents the straight line in which slope of the line
450 W Xenon lamp and microsecond flash lamp as excitation source. gives the strain present in the phosphor and reciprocal of the intercept
The absorption spectrum was recorded in the range 300 nm–1800 nm gives the average crystallite size.
by Varian, Cary-5000 spectrophotometer. The dislocation density (δ) and micro-strain (ε) can be calculated
using the equations [6];
3. Results and discussion 1
=
D2 (3)
3.1. Structural and morphological characterisation
cos
=
Fig. 1 shows the XRD patterns of Sr2-xSiO4:xDy3+ (x = 0.002, 0.004, 4 (4)
0.006, 0.008, 0.01, 0.03, 0.05 and 0.07) nanophosphors. The diffraction The lattice constants are obtained from the following equation [27].
patterns of all samples so formed coincide well with orthorhombic
phase of α-Sr2SiO4 (JCPDS#39-1256). This indicates that Dy3+ ions 4sin2 h2 k2 l2
= + 2 + 2
diffused well into the Sr2SiO4 and have occupied cationic (Sr2+) sites,
2 a2 b c (5)
indicating that Dy3+ ion substitution had no impact on the crystal The inter-planar spacing of the highest peak for orthorhombic
structure. structure is calculated by using the formula;
The structural parameters such as crystallite size (D), dislocation
density (δ), micro-strain (ε) and inter-planar spacing (dhkl) of Sr2- 1
d=
3+ h2 k2 l2
xSiO4:xDy (x = 0 to 0.07) nanophosphors are determined by XRD + +
a2 b2 c2 (6)
data. The average crystallite size (D) is determined from Debye-
Scherrer's (DS) equation [25]; All the calculated parameters are presented in Table 1. The values of

438
C. Manjunath, et al. Journal of Luminescence 211 (2019) 437–445

Dy3+ ions in the ground state are excited to 6P7/2 (28571 cm−1) but
relax non-radiatively and are populated at 4F9/2 (21144 cm−1). This
spectra is characterised by the emission bands at 478 nm, 572 nm,
664 nm and 754 nm corresponding to the transitions 4F9/2→6HJ
(J = 15/2, 13/2, 11/2 and 9/2), respectively [28]. Fig. 10 shows the
schematic representation of possible transitions between different en-
ergy levels. However, the Dy3+ is mainly characterised by blue emis-
sion (478 nm) ascribed to magnetic dipole (MD) transition which is
insensitive to local environment and the other is yellow emission
(572 nm) attributed to forced electric dipole (ED) transition and is al-
lowed at sites that lacks inversion symmetry [1]. It is evident from the
emission spectra that yellow emission at 572 nm is fairly stronger than
the blue emission at 478 nm indicating that Dy3+ ions prefers non
centro-symmetric sites in host matrix. The host matrix (Sr2SiO4) has
two sites Sr(I) and Sr(II) with co-ordination numbers assumed to be 10
and 9, respectively. The Sr(I) polyhedron has hexagonal pseudo-sym-
metry along the y-axis; it shares the face and vertex with two SiO4
tetrahedron which are vertically above and below it respectively, and
the three edges with three SiO4 tetrahedron, which is more symme-
trical. On the other hand, Sr(II) polyhedron is much less symmetrical
and shares neither edges nor faces with any SiO4tetrahedron. Thus,
from the results obtained, we can predict that less symmetrical Sr(II)
sites are more preferable than that of Sr(I) sites [29]. Further structure
diagram indicating the presence of Sr(I) and Sr(II) sites in the Sr2SiO4
host are shown in Fig. 4

3.3. Judd-Ofelt analysis of Sr2-xSiO4:xDy3+ (x = 0.008)

Fig. 2. WH plots of Sr2-xSiO4:xDy3+ (x = 0 to 0.07) nanophosphors. The JO analysis is used for the detailed investigation of optical in-
tensities of rare earth ions in a given host matrix. Generally, optical
micro strain and crystallite size so obtained for all doping concentra- intensities are characterised by f-f transitions that contain contributions
tions vary slightly in a random manner and hence can be concluded that from magnetic dipole, electric dipole and quadruple transitions. For f-f
Dy3+ ions have negligible effect on the crystal structure. Fig. 3 shows transitions, the ED transitions are forbidden while MD transitions are
the TEM images of Sr2-xSiO4:xDy3+ (x = 0.008) nanophosphors cal- allowed whereas quadruple transitions are very small and hence can be
cined at 900 °C. The figure reveals that the particles are agglomerated neglected. The odd order terms of crystal field are responsible for ad-
and are in the nano regime. mixing states of opposite parity from higher lying configurations to
produce mixed parity states that allow for forced ED f-f transitions. The
even order terms of the crystal field are responsible for shifting and
3.2. Luminescence studies splitting of the energy levels [30–32]. Thus, the impact of both odd and
even order terms on the energy levels of the ions in the crystal field can
Fig. 5 shows the excitation spectrum of Dy3+ doped Sr2SiO4 nano- be described by intensity parameters called JO parameters
phosphor monitored for the emission of 571 nm. The spectrum showed ( 2, 4 and 6 ) and are sensitive to local environment. Among the
several bands in the wavelength range 250 nm–400 nm and the wave- parameters so obtained, 2 depends on the covalent bonding of Dy3+
length corresponding to 350 nm exhibited strong excitation. The band with the host, 4 is associated with electron density around the lan-
at 256 nm is associated with charge transfer band. The bands at 324 nm, thanide ion and 6 is related to rigidity of the host matrix in which the
350 nm and 366 nm are related to f-f transition of Dy3+ ions and cor- Dy3+ ions are situated [33].
respond to transition from 6H15/2 to 6P3/2, 6P7/2 and 6P5/2, respectively. The JO intensity parameters are obtained by conventional absorp-
The PL emission spectra of Sr2-xSiO4:xDy3+ (x = 0.002 to 0.07) nano- tion spectrum. Fig. 7 shows the room temperature absorption spectrum
phosphor excited at 350 nm is shown in Fig. 6. After excitation the of Sr2-xSiO4:xDy3+ (x = 0.008). The measured line strength (Smeas) is

Table 1
Comparison of different structural parameters of Sr2SiO4 doped Dy3+ nanophosphors.
Matrix Dy3+ (x) Lattice parameters (Å) crystallite size (nm) Micro Strain (ε) (x Dislocation Density (δ) m−2 Interplanar distance
concentration 10−3) (x 1015) (Å)
A b C Scherrer WH plot
formula

Sr2-xSiO4:xDy3+ JCPDS data 7.079 5.672 9.743 – – – – 2.870


0 7.083 5.712 9.719 25–55 41 0.954 59.4 2.869
0.002 7.080 5.725 9.712 28–50 43 0.824 54.66 2.867
0.004 7.066 5.654 9.775 26–52 42 0.891 56.68 2.869
0.006 7.081 5.689 9.748 28–54 52 0.879 37.99 2.861
0.008 7.069 5.641 9.708 26–60 59 0.816 29.11 2.856
0.01 7.071 5.652 9.738 27–52 47 0.818 45 2.865
0.03 7.061 5.632 9.745 21–51 35 0.937 82.47 2.862
0.05 7.063 5.656 9.783 21–41 36 0.983 76.73 2.870
0.07 7.066 5.674 9.763 24–51 31 0.999 10.33 2.869

439
C. Manjunath, et al. Journal of Luminescence 211 (2019) 437–445

Fig. 3. TEM images of Sr2-xSiO4:xDy3+ (x = 0.008) nanophosphors.

Fig. 4. Structure diagram of Sr2SiO4 nanophosphors.

Fig. 6. Emission spectra of Sr2-xSiO4:xDy3+ (x = 0.002 to 0.07) under the ex-


citation of 350 nm.

Fig. 5. Excitation spectrum of Sr2-xSiO4:xDy3+ (x = 0.01) monitored at


571 nm.

given by equation (7) [32];


Fig. 7. Absorption spectra of Sr2-xSiO4:xDy3+ (x = 0.008) nonophosphor.
2
3hc(2 J+ 1) 3
Smeas = n 2 ( )d
8 3e2 ¯ n +2 (7) 2
Scalc = f n [SL]J| U ( ) |f n [S L ]J
= 2, 4, 6 (8)
where h is Planck's constant, c is velocity of light, J is the total angular
momentum of initial state, e is the electron charge, ¯ is the mean wa- The terms in the brackets are doubly reduced matrix elements for
velength corresponding to different absorption bands, n is the refractive intermediate coupling which can be considered independent of the host
index of host matrix and ( ) d is integrated absorption cross sec- matrix. The values of reduced matrix elements for selected Dy3+ bands
tion. The absorption cross section is calculated using ( ) = t x N
2.303 x A
are taken from Carnall et al. [28]. Further, equation (7) and equation
where A is absorbance, t = 1 mm is thickness of the sample and (8) are least square fitted to obtain three phenomenological parameters
N= 7.142 x 10 20 is Dy3+ ion concentration/cm3 [34,35]. for Dy3+ ions in Sr2SiO4 host,
The calculated line strength (Scalc) with the help of JO parameters is
given by equation (8) [32]; 2 = (1.51 ± 0.027) x 10 20 cm2

440
C. Manjunath, et al. Journal of Luminescence 211 (2019) 437–445

Table 2
RME, measured and calculated line strength for the transitions observed in absorption spectrum.
Transition (From 6H5/2) U (2)
2
U (4)
2
U (6)
2 ¯ (nm) Smeas 10 20cm2 Scalc 10 20cm2 S 10 20cm2

6
F3/2 0 0 0.0610 776 0.05 0.02 0.03
6
F5/2 0 0 0.3452 824 0.12 0.11 0.01
6
F7/2 0 0.1360 0.7146 890 0.22 0.23 0.01
6
F9/2 0 0.5736 0.7213 1076 0.26 0.26 0
6
F11/2 0.9387 0.8292 0.2048 1269 1.52 1.52 0

4 = (0.21 ± 0.031) x 10 20 cm2 nanophosphor given in terms radiative transition (A) for 4F9/2 state is
20
[36];
6 = (0.61 ± 0.015) x 10 cm2
4
The obtained JO parameters are further used to calculate the line (J J, ˜ ) =
p
A(J J) g( ˜ )
8 cn2 (15)
strength. Table 2 shows the values of reduced matrix elements (RME),
along with calculated and measured line strengths. where p is wavelength of a particular transition at peak intensity,
The accuracy of the fit between calculated and measured line g( ˜) =
I( ˜)
is line shape function. I( ˜) is intensity at ˜ . Fig. 8 shows
strengths can be calculated using following relation [36]; I( ˜) d ˜
PL spectrum expressed in terms of wavenumber.
(Smeas Scalc)2
1/2
The emission cross section can be used to determine important
Srms = optical parameters such as gain bandwidth ( e ×
(q p) 1 (9) eff ) and optical gain
( e × ) have also been calculated. Table 4 shows the comparison of
where q is the number of manifolds analysed in absorption spectrum calculated and measured branching ratio, emission cross section, gain
and p = 3 are the number of parameters sought. Thus, the rms devia- bandwidth and optical gain of different transitions.
tion is found to be around 0.047 × 10−20 cm2. Asymmetric ratio is defined as the ratio of integrated intensities of
The JO parameters further are used to calculate total radiative electric dipole transition to magnetic dipole transition. The importance
transition rates A(J J) between the initial state J and final states J is of asymmetric ratio is that it gives the extent to which the crystal filed is
given by Ref. [37]; distorted around Dy3+ ions. Table 5 shows the variation of asymmetric
2 ratios for different Dy3+ concentration in the host.
64 4e2 n2 + 2
A(J J) = n SED (J J) + n3SMD (J J)
3h(2 J+ 1) 3 (10) IED
Asymmetric ratio =
IMD (3.14)
where n is the refractive index, SED (J J) and SMD (J J) are electric and
magnetic dipole line strengths, respectively. Equation (8) is used to where I1 d and I2 d are the area under the curve corresponding to
calculate ED line strength and MD line strength can be written as; transitions 4F9/2 → 6H13/2 and 4F9/2 → 6H15/2, respectively.
2
SMD (J J) = µ 2B f n [S L ]J L+ 2S ?f n [SL]J (11)
3.3.3. Nephelauxtic effect – bonding parameter
( )
where µ B = 2mc . Calculation of magnetic dipole matrix elements is This effect arises due to the expansion of partially filled f-shell and is
due to charge transfer from ligand to the core of the central ion [38].
fairly straightforward and can be done using the relation derived from
angular momentum consideration [37]. Further, the bonding parameter (˜ ) is given by ˜ =
1 ¯
( )
× 100 . As the
¯

name itself indicates, ˜ gives the nature of bonding between Dy3+ ions
3.3.1. Branching ratios and radiative lifetime and the host matrix. For instance, if the bonding parameter is positive
The branching ratios calculated using radiative transition prob- then the bonding is of covalent nature, otherwise it is ionic in nature.
abilities between any two manifolds, which is useful for predicting the The value of ˜ so obtained is found to be 1.875, suggesting that covalent
relative emission intensities from given excited state to all possible low bonding exists for Dy3+− ligands [39].
lying states, is given by Ref. [36];

A(J J) 3.3.4. JO parameters


=
calc
A(J J) (12) Aforementioned, the JO parameters ( 2 , 4 and 6) are de-
termined from conventional absorption spectrum. Of these parameters
The branching ratio measured from fluorescence spectrum is [37];
2 is very sensitive to the crystal field of the host. Also it is well known

IJ J ( )d fact that 2 is closely associated with hypersensitive transitions, such


meas = that hypersensitive transitions increase as the 2 parameter increases,
IJ J ( )d (13)
J
and vice versa. In the case of Dy3+ ions the transitions corresponding to
6
where IJ J ( )d is the area under the curve corresponding to the H5/2 → 6F11/2 and 4F9/2 → 6H13/2 are hypersensitive for which the
given transition under consideration. values of U (2) are 0.9378 and 0.0492 respectively. Further, it is
The radiative lifetime ( rad) is the reciprocal of summation of all experimentally proved that 4 and 6 are less sensitive to the host
probable radiative transitions (A) originating from particular excited environment and are structural dependent parameters [40].
dstate to all low lying manifolds is [32]; The JO parameters in our article followed the trend 2 > 6 > 4 .
1 The greater value of 2 indicates that the field is non-centrosymmetric
rad = in nature and lacks inversion symmetry. This suggests that strong
A(J J) (14)
covalent bond exists between Dy−O. The bonding parameter being
The values of line strengths, transition probabilities, calculated (JO) positive proposes that the covalent nature exists between rare earth and
branching ratio and radiative lifetime obtained are shown in Table 3. ligand anions. Furthermore, the obtained values of asymmetric ratios
are greater than 1 show that the field is asymmetric in nature for all
3.3.2. Emission cross section and asymmetric ratios concentrations. Moreover, the value for 0.008 mol concentration is
The emission cross section of Sr2-xSiO4:xDy3+ (x = 0.008) maximum (4.98) illustrates that crystal field around Dy3+ ions is most

441
C. Manjunath, et al. Journal of Luminescence 211 (2019) 437–445

Table 3
Predicted transition probabilities and branching ratios for all possible transitions from 4F9/2 in Sr2-xSiO4:xDy3+ (x = 0.008) nanophosphors.
Transitions U (2)
2
U (4)
2
U (6)
2 λ (nm) SED 10 20cm2 SMD 10 20cm2 A(S 1) β r (µs)

4
F9/2 → 6F1/2 0 0.0003095 0 1350 0.000065 0 0.093 < 0.0001 241
4
F9/2 → 6F3/2 0 0.0000001 0.0002578 1258 0.000159 0 0.281 < 0.0001
4
F9/2 → 6F5/2 0.0060318 0.0009278 0.0004103 1147 0.009553 0 22.29 0.0053
4
F9/2 → 6F7/2 0.0001994 0.0043310 0.0031234 989 0.003115 0.018659 79.26 0.0027
4
F9/2 → 6H5/2 0.0000029 0.0035694 0.0011671 916 0.001465 0 6.710 0.0016
4
F9/2 → 6H7/2 0.0008258 0.0078220 0.0068200 836 0.007049 0.006335 80.66 0.0102
4
F9/2 → 6F9/2 0.0008188 0.0055662 0.0004444 830 0.002676 0.010954 134.92 0.0162
4
F9/2 → 6F11/2 0.0033784 0.0032310 0.0024795 746 0.007292 0.073359 690.05 0.0163
4
F9/2 → 6H9/2 0.0022314 0.0022888 0.0031636 745 0.006154 0.004208 88.25 0.0126
4
F9/2 → 6H11/2 0.0090912 0.0016778 0.0036066 655 0.016280 0.011614 291.1 0.0351
4
F9/2 → 6H13/2 0.0492126 0.0164855 0.0557581 569 0.111785 0 2137 0.515
4
F9/2 → 6H15/2 0 0.0047065 0.0293262 476 0.018877 0 616.3 0.148

Table 5
Asymmetric ratios of Sr2SiO4:xDy3+ nanophosphors.
Dy3+ mol %0 0.2 0.4 0.6 0.8 1 3 5 7
Asymmetric ratio 3.35 2.99 3.01 4.98 2.67 2.97 3.52 3.85

Fig. 8. PL spectrum of Sr2-xSiO4:xDy3+ (x = 0.008) nanophosphor in terms of


wavenumber.

distorted, and is also supported by the values of JO parameter ( 2 ) and


bonding parameter. The value of parameter ( 6 ) is bit higher and is
related to the density of electrons in 6s shell and suggests that the host
Fig. 9. Photoluminescence decay curves of Sr2-xSiO4:xDy3+ (x = 0.008).
is fairly rigid.
The phenomenological JO parameters are further used to calculate
the branching ratio that plays a vital role in determining the probability may be suitable for laser application.
of stimulated emission from a particular transition, which is the main
requisite for laser action. In the case of Dy3+ the transition from 4F9/2 3.4. Photoluminescence lifetime decay measurements
level is considered. Table 4 shows the values of branching ratio for the
following transitions 4F9/2→6HJ (J = 15/2, 13/2, 11/2 and 9/2). Of all Fig. 9 shows the room temperature lifetime decay curves of Sr2-
3+
the transitions, the measured branching ratio corresponding to the xSiO4:xDy (x = 0.008) nanophosphor monitored for the emission of
transition 4F9/2 → 6H13/2 is 71%, and is the largest compared to other 571 nm under the excitation of 350 nm. The decay curves thus obtained
three transitions. It is beneficial good laser emission to have a large exhibited bi-exponential behaviour given by equation (17) [44,45].
stimulated emission cross section, which is associated with the gain of
I(t) = A1e( ) + A2e( t2 )
t
the laser material [38,41,42]. The importance of emission cross section 1 (17)
is that it has a significant role to play in low threshold, high gain laser where A1 and A2 are the constants, 1 and 2 are decay times respec-
application for continues wave (CW) laser action, for example [39,43]. tively. The bi-exponential fitting confirms that the host matric is asso-
The emission cross section for the transition 4F9/2 → 6H13/2 is ciated with two sites Sr(I) and Sr(II) at which Dy3+ ions can be sub-
1.02 × 10−20 cm2 and the values of optical gain and gain bandwidth stituted. Further equation (18) is used to calculated average lifetime
are shown in Table 4. The result obtained suggests that the phosphors ( av )

Table 4
Measured emission cross section, gain bandwidth and optical gain for different transitions in Sr2-xSiO4:xDy3+ (x = 0.008) nanophosphor.
Transition λ (nm) g(¯ ) cm meas calc (J,J ; ¯ ) cm2 Gain bandwidth cm3 Optical gain cm2s

4
F9/2 → 6H15/2 478 0.0034 0.143 0.148 0.21 × 10−20 0.21 × 10−26 5.11 × 10−25
4
F9/2 → 6H13/2 573 0.0024 0.712 0.515 1.02 × 10−20 1.77 × 10−26 24.5 × 10−25
4
F9/2 → 6H11/2 664 0.0019 0.048 0.0351 0.07 × 10−20 0.13 × 10−26 1.69 × 10−25
4
F9/2 → 6H9/2 754 0.0023 0.096 0.0126 0.23 × 10−20 0.37 × 10−26 5.54 × 10−25

442
C. Manjunath, et al. Journal of Luminescence 211 (2019) 437–445

Fig. 10. Energy level diagram and schematic representation of cross relaxation of Dy3+ ions.

A1 12 + A2 22 mechanisms, specifically the exchange interaction, radiation reabsorp-


=
av
A1 1 + A2 2 (18) tion and multipole-multipole interaction.
According to Blasse et al. [47], the critical transfer distance (R C ) is
The average lifetime thus calculated is found to be 0.728 ms.
an important parameter that plays a vital role in the exchange inter-
action and is represented by equation (19)
3.5. Cross relaxation and concentration quenching
1
3
Fig. 10 shows the energy level diagram of Dy3+ ions. Considering 3V
RC =
the resonance conditions, there are three probable cross relaxations CR- 4 x CN (19)
1, CR-2 and CR-3 that are responsible for the de-population of Dy3+
ions from 4F9/2 level. where x C is critical concentration of Dy3+ ions and is 0.008 in our case,
4
V and N are volume of the unit cell and number of cationic sites
F 9/2 + 6 H15/2 6
H9/2/ 6 F11/2 + 6F3/2 available. Thus, the critical transfer distance calculated is found to be
4 11.3 Å. In principle, exchange interaction occurs only if the critical
F 9/2 + 6 H15/2 6
H7/2/ 6 F9/2 + 6 F5/2
distance is in the range 3 − 4 Å, which in our case is not satisfied.
4
F 9/2 + 6 H15/2 6
F3/2/ 6 H9/2 + 6 F11/2 Hence, the process of quenching by this mechanism is ruled out. Fur-
ther, the possibility of radiation reabsorption is considered only if the
The possible explanation behind the process is as follows. The Dy3+ excitation and emission spectra overlap with each other, but it is clear
ions (CR-1) in excited state (4F9/2 = 21144 cm−1) make a transition to from Figure (5 and 6) that the probability of contributions due to
the lower state (6H9/2/6F11/2 = 7692/7730 cm−1) emitting a photon of quenching by this method can be discarded. The multipole-multipole
suitable energy (13452/13414 cm−1). This emitted photon is taken by interaction is said to be dominant if the non-radiative transfer of energy
the Dy3+ ions in the ground state (6H15/2 = 40 cm−1) and is excited to is proportional to x /3 where θ = 6, 8 or 10 corresponds to dipole-di-
the higher state (6F3/2 = 13212 cm−1). Similarly CR-2 and CR-3 pro- pole, dipole-quadrupole or quadrupole-quadrupole interaction respec-
cesses occur in an analogous way. Finally, the Dy3+ ions participating tively [48]. The energy transfer mechanism in this case is determined
in all three processes will make transition to ground state and hence are by correlating emission intensity and Dy3+ ion concentration by
responsible for quenching of luminescence related to 4F9/2 level [46]. equation (20) [8],
It is quite obvious from Fig. 6 that the peak positions did not vary
with Dy3+ ion concentration. However, the emission intensities steadily I 1
=k (x) 3
increased with increase in Dy3+ doping concentration and at con- x (20)
centration x = 0.008 the intensity reached a maximum. Subsequently,
the intensity gradually decreased with increase in Dy3+ concentration, ()
where x is emission intensity per ion concentration, k and β are the
I

which may be attributed to concentration quenching. The quenching is constants for each interaction.
mainly due to non radiative transfer of energy among Dy3+ ions. On further simplifying equation (20) we get
Quenching mechanisms can be characterised by three possible

443
C. Manjunath, et al. Journal of Luminescence 211 (2019) 437–445

nanophosphors have a favourable position close to standard white light


emission (0.33, 0.33) [49].
Further, correlated color temperature (CCT) is adopted to study the
performance of phosphors and is given by Ref. [50].
CCT = 437n3 + 3601n2 6861 n+ 5514.3 (22)
The significance of equation (22) is that it gives the quality of white
x x
light emission. n= y ye is the reciprocal of the slope line, and
e
(x e ye) = (0.3320, 0.1858) are the coordinates of epicentre. The CCT
values of synthesised nanophosphors are in the range 4976 K–8055 K.
The Sr2-xSiO4:xDy3+ (x = 0.008) nanophosphors with CIE coordinates
close to ideal white light has CCT values of 5347 K. From the com-
mercial point of view the CCT values should be greater than 5000 K to
be ascertained as cold white light. However, the obtained CCT (5347 K)
value has a good agreement with the standard CIE daylight series of
illuminants D55 and corresponds to sunny daylight around noon [50].
Further, the CIE coordinate (0.337, 0.388) for Sr2-xSiO4:xDy3+
Fig. 11. Variation of log(I/x) w.r.t. log(x) in Sr2SiO4 doped Dy3+ nanopho- (x = 0.008) is close to standard illuminate D55 with coordinates (0.333,
sphors. 0.348) [51].

4. Conclusion

Self-propagating solution combustion method is adopted for the


synthesis of Sr2-xSiO4:xDy3+. The crystallinity of the nanophosphors is
studied by XRD along with calculations of relevant structural para-
meters. The crystallite size obtained by scherrer and WH plots suggests
the nanophosphors to be in nano scale and is also confirmed by TEM.
The emission spectra of Sr2-xSiO4:xDy3+ excited at 350 nm showed that
the optimised concentration is x = 0.008 for which intense emission
was observed and decreases with further increase of Dy3+. The JO
parameters obtained from JO analysis showed the behaviour
Ω2 > Ω6 > Ω4 indicating the surrounding field to be non-centrosym-
metric in nature and is further validated by the calculated asymmetric
ratios. The JO parameters obtained were used to estimate different
radiative properties. The peak emission cross section, determining the
strength of the emission, is 1.02 × 10−20 cm2 for the 4F9/2 → 6H13/2
transition. The corresponding branching ratio is around 72%, sug-
gesting that the nanophosphors are suitable for laser application. The
decrease in emission intensity after optimum concentration is attributed
to dipole-dipole interaction and is the main cause for non-radiative
energy transfer. CIE coordinates varied randomly based on Y/B ratio
and the coordinates corresponding to optimised concentration is
Fig. 12. CIE diagram of Sr2-xSiO4:xDy3+ nanophosphors excited at 350 nm.
(0.337, 0.388) are close to ideal white light emission with coordinates
(0.33, 0.33). A CCT value varies between 4976 K and 8055 K for dif-
ferent concentration with 0.8 mol% concentration having CCT of
I
log = k log(x) 5347 K and is in good agreement with standard illuminant D55.
x 3 (21)
Acknowledgements
where ( )3
= 2.35 ± 0.2521 is the slope of the line obtained after
linear fitting the graph of log(I/x) v/s log(x) and is shown in Fig. 11.
C. Manjunath is grateful to the Management, Principal and HOD
Hence, the value is found to be = 7 and can be taken as approximately
Physics of Sri Siddhartha Institute of Technology, Tumkur, for their
equal to 6. The value thus suggests that the dipole-dipole interaction
constant support and encouragement. R. Hari Krishna is grateful to the
dominates the phenomenon of quenching in this case.
Management and HOD Chemistry of M.S. Ramaiah Institute of
Technology, Bangalore, for their constant support and encouragement.
3.6. CIE coordinates and CCT values
References
The Commission International del’Eclairage (CIE) chromaticity co-
ordinates of Sr2-xSiO4:xDy3+ (x = 0.2–11 mol%) nanophosphors ex- [1] V.R. Bandi, Y.T. Nien, I.G. Chen, Enhancement of white light emission from novel
Ca3Y2Si3O12:Dy3+ phosphors with Ce3+ ion co-doping, J. Appl. Phys. 108 (2010).
cited at 350 nm are shown in Fig. 12.
[2] S.D. Meetei, M.D. Singh, D.S. Singh, Facile synthesis, structural characterization,
The CIE coordinates plays a vital role in categorizing the sig- and photoluminescence mechanism of Dy3+ doped YVO4 and Ca2+ co-doped
nificance of the phosphors in lighting and display applications. The YVO4 :Dy3+ nano-lattices, J. Appl. Phys. 115 (2014) 204910.
coordinates thus obtained for 0.002, 0.004, 0.006, 0.008, 0.01, 0.03, [3] G. Blasse, B.C. Grabmaier, Luminescent Materials, Springer, Berlin, 1994.
[4] S. Nakamura, InGaN/AlGaN blue light emitting diodes, J. Vac. Sci. Technol. A
0.05 and 0.07 Dy3+ concentration in Sr2SiO4 host are (0.304, 0.300), Vacuum, Surfaces, Film. 13 (1995) 705–710.
(0.314, 0.407), (0.304, 0.367), (0.337, 0.388), (0.350, 0.406), (0.344, [5] G.S.R. Raju, E. Pavitra, J.S. Yu, Pechini synthesis of lanthanide (Eu3+/Tb3+or
0.408), (0.282, 0.344) and (0.281, 0.365) respectively. It is clear from Dy3+) ions activated BaGd2O4 nanostructured phosphors: an approach for tunable
emissions, Phys. Chem. Chem. Phys. 16 (2014) 18124–18140.
the coordinates in CIE diagram that Sr2-xSiO4:xDy3+ (x = 0.008)

444
C. Manjunath, et al. Journal of Luminescence 211 (2019) 437–445

[6] S. Dutta, S. Som, S.K. Sharma, Luminescence and photometric characterization of [29] M. Catti, G. Gazzoni, G. Ivaldi, Structure of twinned β-Sr2SiO4 and α-Sr1.9Ba0.1SiO4,
K+ compensated CaMoO4:Dy3+ nanophosphors, Dalton Trans. 42 (2013) 9654. Acta Crystallogr. C39 (1983) 29–34.
[7] P. Haritha, I.R. Martín, K. Linganna, V. Monteseguro, P. Babu, S.F. León-Luis, [30] B.R. Judd, Optical absorption intensities of rare-earth ions, Phys. Rev. 127 (1962)
C.K. Jayasankar, U.R. Rodríguez-Mendoza, V. Lavín, V. Venkatramu, Optimizing 750–761.
white light luminescence in Dy3+-doped Lu3Ga5O12nano-garnets, J. Appl. Phys. [31] G.S. Ofelt, Intensities of crystal spectra of RareEarth ions, J. Chem. Phys. 37 (1962)
116 (2014). 511–520.
[8] P. Du, L.K. Bharat, X.Y. Guan, J.S. Yu, Synthesis and luminescence properties of [32] B.M. Walsh, Judd-Ofelt theory: principles and practices, in: B. Di Bartolo, O. Forte
color-tunable Dy3+-activated CaWO4 phosphors, J. Appl. Phys. 117 (2015) 3–9. (Eds.), Advances in Spectroscopy for Lasers and Sensing, Springer, Netherlands,
[9] K. Szczodrowski, J. Barzowska, N. Gorecka, M. Grinberg, Stabilization of Eu3+ 2006, pp. 403–433.
under a reductive atmosphere by the Al3+ co-doping of Sr2SiO4:Eu2+/Eu3+, RSC [33] C.K. Jorgensen, R. Reisfeld, Judd-ofelt parameters and chemical bonding, J. Less
Adv. 6 (2016) 48001–48008. Common. Met. 93 (1983) 107–112.
[10] H. Terraschke, C. Wickleder, UV , blue , green , yellow , red , and Small : newest [34] W.F. Krupke, Induced-emission cross sections in neodymium laser glasses, IEEE J.
developments on Eu2+ -doped nanophosphors, Chem. Rev. 115 (2015) Quantum Electron. 10 (1974) 450–457.
11352–11379. [35] W.F. Krupke, Radiative transition probabilities within the 4f3 ground configuration
[11] I.P. Sahu, D.P. Bisen, N. Brahme, Dysprosium doped di-strontium magnesium di- of Nd:Yag, IEEE J. Quantum Electron. 7 (1971) 153–159.
silicate white light emitting phosphor by solid state reaction method, Displays 35 [36] D.K. Sardar, D.M. Dee, K.L. Nash, R.M. Yow, J.B. Gruber, Optical absorption in-
(2014) 279–286. tensity analysis and emission cross sections for the intermanifold and the inter-Stark
[12] L. Zhang, P. Han, K. Wang, Z. Lu, L. Wang, Y. Zhu, Q. Zhang, Enhanced lumines- transitions of Nd3+ (4f3) in polycrystalline ceramic Y2O3, J. Appl. Phys. 100 (2006)
cence of Sr2SiO4:Dy3+ by sensitization (Ce3+/Bi3+) and its composition-induced 123106–123111–7.
phase transition, J. Alloys Compd. 541 (2012) 54–59. [37] B.M. Walsh, N.P. Barnes, B. Di Bartolo, Branching ratios, cross sections, and ra-
[13] S.K. Gupta, M. Kumar, V. Natarajan, S.V. Godbole, Optical properties of sol-gel diative lifetimes of rare earth ions in solids: application to Tm3+ and Ho3+ ions in
derived Sr2SiO4:Dy3+- Photo and thermally stimulated luminescence, Opt. Mater. LiYF4, J. Appl. Phys. 83 (1998) 2772–2787.
35 (2013) 2320–2328. [38] A. Agarwal, I. Pal, S. Sanghi, M.P. Aggarwal, Judd-Ofelt parameters and radiative
[14] D.V. Sunitha, H. Nagabhushana, S.C. Sharma, B.M. Nagabhushana, B. Daruka properties of Sm3+ ions doped zinc bismuth borate glasses, Opt. Mater. 32 (2009)
Prasad, R.P.S. Chakradhar, Study on low temperature solution combustion syn- 339–344.
thesized Sr2SiO4:Dy3+ nano phosphor for white LED, Spectrochim. Acta Part A Mol. [39] P. Van Do, V.P. Tuyen, V.X. Quang, N.T. Thanh, V.T.T. Ha, N.M. Khaidukov,
Biomol. Spectrosc. 127 (2014) 381–387. Y.I. Lee, B.T. Huy, Judd–Ofelt analysis of spectroscopic properties of Sm3+ ions in
[15] D. Dutczak, A. Milbrat, A. Katelnikovas, A. Meijerink, C. Ronda, T. Jüstel, Yellow K2YF5 crystal, J. Alloys Compd. 520 (2012) 262–265.
persistent luminescence of Sr2SiO4:Eu2+,Dy3+, J. Lumin. 132 (2012) 2398–2403. [40] S. Tanabe, T. Ohyagi, N. Soga, T. Hanada, Compositional dependence of Judd-Ofelt
[16] L. Zhang, Z. Lu, H. Yang, P. Han, N. Xu, Q. Zhang, Preparation of Dy3+-activated parameters of Er3+ ions in alkali-metal borate glasses, Phys. Rev. B 46 (1992)
strontium orthosilicate (Sr2SiO4:Dy3+) phosphors and its photoluminescent prop- 3305–3310.
erties, J. Alloys Compd. 512 (2012) 5–11. [41] R. Rajaramakrishna, B. Knorr, V. Dierolf, R.V. Anavekar, H. Jain, Spectroscopic
[17] M.P. Saradhi, N. Lakshminarasimhan, S. Boudin, K.V.K. Gupta, U.V. Varadaraju, properties of Sm3+-doped lanthanum borogermanate glass, J. Lumin. 156 (2014)
B. Raveau, Enhanced luminescence of Sr2SiO4:Dy3+ by sensitization (Ce3+/Eu2+) 192–198.
and fabrication of white light-emitting-diodes, Mater. Lett. 117 (2014) 302–304. [42] M.C. Pujol, J. Massons, M. Aguilo, F. Diaz, M. Rico, C. Zaldo, Emission cross sections
[18] M.A. Tshabalala, F.B. Dejene, S.S. Pitale, H.C. Swart, O.M. Ntwaeaborwa, and spectroscopy of Ho3+ laser channels in KGd(WO4)2 single crystal, IEEE J.
Generation of white-light from Dy3+ doped Sr2SiO4 phosphor, Phys. B Condens. Quantum Electron. 38 (2002) 93–100.
Matter 439 (2014) 126–129. [43] F. Ahmadi, R. Hussin, S.K. Ghoshal, Judd-Ofelt intensity parameters of samarium-
[19] N. Lakshminarasimhan, U.V. Varadaraju, Luminescence and afterglow in doped magnesium zinc sulfophosphate glass, J. Non-Cryst. Solids 448 (2016)
Sr2SiO4:Eu2+, RE3+ [RE = Ce, Nd, Sm and Dy] phosphors-Role of co-dopants in 43–51.
search for afterglow, Mater. Res. Bull. 43 (2008) 2946–2953. [44] S. Som, S. Das, S. Dutta, H.G. Visser, M.K. Pandey, P. Kumar, R.K. Dubey,
[20] B. Zhang, X. Yu, T. Wang, S. Cheng, J. Qiu, X. Xu, Photostimulated and long per- S.K. Sharma, Synthesis of strong red emitting Y2O3:Eu3+ phosphor by potential
sistent luminescence properties from different crystallographic sites of β-Sr2SiO4: chemical routes : comparative investigations on the structural evolutions , photo-
Eu2+ , R3+ (R = Tm, Gd), J. Am. Ceram. Soc. 98 (2015) 171–177. metric properties and Judd – ofelt analysis, RSC Adv. 5 (2015) 70887–70898.
[21] X. Fan, W. Chen, S. Xin, Z. Liu, M. Zhou, X. Yu, D. Zhou, X. Xu, J. Qiu, Achieving [45] B.P. Singh, A.K. Parchur, R.S. Ningthoujam, A.A. Ansari, P. Singh, S.B. Rai,
long-term zero-thermal-quenching with the assistance of carriers from deep traps, Enhanced photoluminescence in CaMoO4:Eu3+ by Gd3+ co-doping, Dalton Trans.
J. Mater. Chem. C. 6 (2018) 2978–2982. 43 (2014) 4779–4789.
[22] Z. Zhe, X. Xu-Hui, Q. Jain-Bie, Z. Xiu, Y. Xue, Photo-stimulated and long persistent [46] L. Cheng, X. Li, J. Sun, H. Zhong, Y. Tian, J. Wan, W. Lu, Y. Zheng, T. Yu, L. Huang,
luminescence properties of Sr3SiO5∶Eu2+,RE3+ (RE = Nd3+,Ho3+,La3+), H. Yu, B. Chen, Investigation of the luminescence properties of Dy3+doped α-
Spectrosc. Spectr. Anal. 34 (2014) 1486–1491. Gd2(MoO4)3 phosphors, Phys. B Condens. Matter. 405 (2010) 4457–4461.
[23] K.C. Patil, M.S. Hegde, T. Rattan, S.T. Aruna, Nanocrystalline Oxide Materials, [47] G. Blasse, Energy transfer in oxidic phosphors, Phys. Lett. 28 (1968) 444–445.
World Scientific, Singapore, 2008. [48] L.G.V. Uitert, Characterization of energy transfer interactions between rare earth
[24] J.J. Kingsley, K.C. Patil, A novel combustion process for the synthesis of fine par- ions, J. Electrochem. Soc. 114 (1967) 1048.
ticels α-Alumina and related oxide materials, Matrials Lett 6 (1988) 427–432. [49] A. Kumar, S.J. Dhoble, D.R. Peshwe, J. Bhatt, J.J. Terblans, H.C. Swart, Crystal
[25] P. Klug, L.E. Alexander, X-ray Diffraction Procedure, Wiley, New York, 1954. structure, energy transfer mechanism and tunable luminescence in Ce3+/Dy3+ co-
[26] G.K. Williamson, W.H. Hall, X-ray line broadening from filed aluminium and wol- activated Ca20Mg3Al26Si3O68 nanophosphors, Ceram. Int. 42 (2016) 10854–10865.
fram, Acta 1 (1953) 22–31. [50] C.S. McCamy, Correlated color temperature as an explicit function of chromaticity
[27] J.W. Edington, Electron Diffraction in the Electron Microscope, Palgrave, London, coordinates, Color Res. Appl. 17 (1992) 142–144.
1975. [51] A. Balakrishna, O.M. Ntwaeaborwa, Study of luminescent behavior and crystal
[28] W.T. Carnall, P.R. Fields, K. Rajnak, Spectral intensities of the trivalent lanthanides defects of different MNa[PO4]-Dy3+ phosphors (M = Mg, Ca, Sr and Ba), Sensor.
and actinides in solution. II. Pm3+, Sm3+, Eu3+, Gd3+, Tb3+, Dy3+, and Ho3+, Actuator. B Chem. 242 (2017) 305–317.
J. Chem. Phys. 49 (1968) 4412–4423.

445

You might also like