You are on page 1of 56

SECTION I

Dy3+ doped BaY2ZnO5 Nanophosphors


Over the last decade, rare earth based nanophosphors are at the forefront of
scientific research due to their unique electronic, optical and chemical properties
arising from 4f shell and their significant applications in different fields [1-6]. Apart
from possessing excellent magnetic, thermal, electronic and superconducting
properties, the ternary oxides, BaLn2ZnO5 (Ln= trivalent lanthanide ions) with stable
crystal structure have been emerged as a potential choice in the field of luminescent
materials when doped with different RE3+ ions [7-23]. BaLn2ZnO5 possess
orthorhombic structure belonging to space group Pbnm with Sm, Eu, Gd, Dy, Ho, Y,
Er or Tm ions and tetragonal structure having 14/mcm space group with La or Nd
ions. The orthorhombic BaY2ZnO5 structure with space group Pbnm comprises of
BaO11, YO7 and ZnO5 polyhedra units [24]. In this lattice, Y3+ atoms occupy two
different 7- fold oxygen coordinated sites inside monocapped triogonal prisms with
same crystal field symmetry (Cs) exhibiting slightly different Y-O distances. These
prisms share edges to form wave-like chain parallel to the b-axis and two such units
join to form basic Y2O11 structure. Ba atoms resides in distorted 11-fold coordination
cages while ZnO5 have distorted tetragonal pyramidal coordination [11-13].
Among RE3+ ions, trivalent dysprosium ions have been considered as promising
luminescent center for white light emission originated from blue (4F9/2 6H15/2) and
yellow (4F9/26H13/2) transitions [25-28]. From the prospective of efficient optoelectronic applications, nanophosphors having desired composition, reproducible size,
shape and structure are essential, which can be achieved using reliable synthesis
routes. To meet the requirement of white light emission, some conventional efforts
has been made to synthesize Dy3+ doped BaLn2ZnO5 (Ln= Gd or Y) phosphor at
higher temperature [11-12]. In comparison to earlier employed vibrating milled solid
state method for BaY2ZnO5: Dy3+ phosphor which requires 12h long sintering at
1250C, solution combustion synthesis (SCS) has been evolved as a simple, low-cost,
rapid, easy and relatively green technique to yield phosphor with defined size,
morphology, fewer defects and excellent luminescent properties [29]. Literature
reveals that no work has been carried out on BaY2ZnO5: Dy3+ phosphors using SCS
hence, the present work demonstrate the solution combustion synthesis of Dy3+ doped
BaY2ZnO5 nanophosphors. The phase formation and tuning of white emission color in
184

this single phased host lattice with the variation of dysprosium ions contents were also
investigated in details.
A. EXPERIMENTAL DETAILS
i) Powder Synthesis
BaY2(1-x)ZnO5: 2xDy3+ nanophosphors were synthesized by solution
combustion approach using high purity Ba(NO3)2, Y(NO3)3.6H2O, Zn(NO3)2,
Dy(NO3)3.6H2O and urea as starting materials. The chemical reaction was:
Ba(NO3)2 + 2(1-x)Y(NO3)3 + 2xDy(NO3)3 + Zn(NO3)2 + 8.32CH4N2O (urea)
BaY2(1-x)Dy2xZnO5(s) + gaseous products.
According to nominal composition of BaY2(1-x)Dy2xZnO5, where x = 0.005 to
0.07, a stoichiometric amount of metal nitrates were dissolved in minimum quantity
of deionized water and then urea was added as fuel. Its amount was calculated on the
basis of total oxidizing and reducing valencies of oxidizer and the fuel (urea)
according to the concept used in propellant chemistry [30]. This aqueous paste
containing calculated amount of metal nitrates and urea was then placed in a
preheated furnace maintained at 500oC. The mixture of metal nitrates (oxidizers) and
fuel (urea) undergo rapid and self-sustaining combustion process and the chemical
energy released during this exothermic redox reaction results in dehydration and
foaming followed by decomposition. Consequently, the large amounts of volatile
combustible gases generated alongwith flames, yields voluminous solid within 5-8
minutes. SCS utilizes the enthalpy of combustion for the formation of BaY2(1x)Dy2xZnO5

nanophosphor and the powders thus obtained were again sintered at

different temperature for 3h in order to evaluate the effect of sintering on the


structural and luminescent features of BaY2ZnO5: Dy3+ nanophosphors.
ii) Powder Characterization Techniques
The phase purity of BaY2(1-x)ZnO5: 2xDy3+ powders was evaluated by X-ray
diffraction (XRD) using Rigaku Ultima IV diffractometer with CuK radiation at 40
kV tube voltage and 40 mA tube current in the 2 range between 10-80o. The surface
morphology of the samples was examined using Jeol JSM-6510 scanning electron
microscope (SEM). The crystallite size and shape has been evaluated by transmission
electron microscopy (TEM) using H-7500, Hitachi, Japan. The photoluminescence
excitation and emission in the ultraviolet-visible region and decay curves under time
185

scan mode were carried out on Hitachi F-7000 fluorescence spectrophotometer


equipped with Xe-lamp at room temperature.
B. RESULTS AND DISCUSSION
i) X-ray Studies
The XRD profiles of BaY1.92Dy0.08ZnO5 nanophosphor as-synthesized and
sintered at 800C to 1100C with the standard JCPDS No.049-0516 are compared in
Fig.5.1.1a As-synthesized BaY1.92Dy0.08ZnO5 powder exhibits poor crystallinity as
evidenced by relatively weak and broad peaks corresponding to BaY2ZnO5 phase in
addition to many other peaks denoted as * due to undecomposed nitrates in the
sample. The powder sintered at 800C shows BaY2O4 (JCPDS No.082-2319) and
Y2O3 phase (JCPDS No.088-2162) ascribed as and , respectively alongwith the
BaY2ZnO5 phase. Even at 900oC, pure orthorhombic phase could not be achieved as
secondary BaY2O4 phase is still detected in the sample alongwith all characteristic
BaY2ZnO5 peaks. At 1100C, all the peaks in diffraction patterns are well indexed to
dominant orthorhombic BaY2ZnO5 phase (JCPDS No.049-0516, Pbnm space group
with lattice parameters a = 7.070 Ao, b = 12.33 Ao and c = 5.709 Ao). No traces of
other peaks due to additional phases apart from single orthorhombic BaY2ZnO5 phase
was observed at this temperature. It is clearly seen that on increasing the temperature,
the intensity of main peak at 29.61 enhanced with the decrease in line width,
concluding that the crystalline degree of dysprosium doped BaY2ZnO5 nanocrystals is
improved apparently with the rise in sintering temperature.
XRD patterns of BaY2(1-x)Dy2xZnO5 powders sintered at 1100C, doped with
different contents of dysprosium ions alongwith the standard JCPDS No.049-0516 are
illustrated in Fig. 5.1.1b. The closeness of ionic radii of Y3+ (0.96 Ao) and Dy3+ (0.97
Ao) make the substitution process easier at the same symmetry (Cs) sites in
BaY2ZnO5 host lattice [24]. It is quite evident that all the samples exhibit single
orthorhombic BaY2ZnO5 phase belonging to space group Pbnm, indicating small
quantity of doped Dy3+ ions had no influence on the crystallographic structure of
BaY2ZnO5.The crystallite size, D of BaY2(1-x)ZnO5: 2xDy3+ was estimated by
Scherrers formula, D 0.941/ cos, where is the wavelength of CuK radiation
(0.1548 nm), is the full width in radians at half-maximum (FWHM) and is the
Braggs angle of an observed X-ray diffraction peak.

186

Fig.5.1.1. XRD profiles of (a) BaY1.92Dy0.08ZnO5 powder sintered at different


temperatures; (b) BaY2(1-x)Dy2xZnO5 (x = 0.005 - 0.10) alongwith standard data of
BaY2ZnO5 (JCPDS No. 49-0516)
187

Table 5.1.1 summarizes the calculated average crystallite sizes (D)


corresponding to the observed phases for the different mol% of Dy3+ ions in
BaY2(1-x)ZnO5 nanophosphors, sintered at different temperatures. It was observed that
all the powders are in nano-regime range although an increase in crystallite size with
the rise in sintering temperature was noticed due to enhanced atomic mobility of
particles which leads to faster grain growth at higher temperature.
Table 5.1.1 Crystallites size (nm) of BaY2(1-x)ZnO5 : 2xDy3+ at different mol%
calculated using Scherrers equation

BaY2(1-x)Dy2xZnO5 phase and Crystallites size (nm) after sintering at;


500 oC
x = 0.005

x = 0.03

x = 0.04

x = 0.07

BaY2ZnO5

18.5

Ba(NO3)2

53.2

BaY2ZnO5

17.0

Ba(NO3)2

75.3

BaY2ZnO5

16.9

Ba(NO3)2

66.1

BaY2ZnO5

17.0

Ba(NO3)2

55.6

800 oC

900 oC

1100 oC

BaY2ZnO5
BaY2O4
Y 2O 3

34.6
35.9
36.5

BaY2ZnO5
BaY2O4

64.0
76.8

BaY2ZnO5

80.0

BaY2ZnO5
BaY2O4
Y 2O 3

36.3
43.9
35.5

BaY2ZnO5
BaY2O4

71.4
78.2

BaY2ZnO5

88.4

BaY2ZnO5
BaY2O4
Y 2O 3

34.8
51.2
37.9

BaY2ZnO5
BaY2O4

73.4
82.8

BaY2ZnO5

92.8

BaY2ZnO5
BaY2O4
Y 2O 3

36.9
37.3
32.9

BaY2ZnO5
BaY2O4

67.9
71.2

BaY2ZnO5

81.8

ii) Morphological Studies


The SEM image of BaY1.92Dy0.08ZnO5 powder sintered at 1100oC, exhibits
narrow size distribution of spherical shaped particles with a slight agglomerate
phenomenon

(Fig.5.1.2.)

Uniform

and

smooth

spherical

morphology

of

BaY1.92Dy0.08ZnO5 sample, sintered at 1100oC in diameter range 80-90 nm, is also


depicted by TEM micrograph (Fig.5.1.3). The high crystalline quality of the Dy3+
doped BaY2ZnO5 nanoparticles indicated by XRD results has been confirmed by
TEM as crystallite size estimation from both analyses are consistent. Morphology
studies reveal that the spherical shaped BaY1.92Dy0.08ZnO5 nanocrystalline particles
188

due to high packing density, lower scattering light and brighter luminescence
performance are found to be more desirable than irregularly shaped particles as good
crystallization affects the photoluminescence properties significantly [31].

Fig.5.1.2. SEM image of BaY1.92Dy0.08ZnO5 nanophosphors sintered at 1100C

Fig.5.1.3. TEM image of BaY1.92Dy0.08ZnO5 nanophosphors sintered at 1100C


iii) Luminescent Studies
The photoluminescence excitation (PLE) spectrum of BaY1.92Dy0.08ZnO5
nanophosphor sintered at 1100oC, measured with em = 580 nm corresponding to
4

F9/26H13/2 transition in 200-500 nm range is shown in Fig. 5.1.4. The PLE spectrum
189

consists of a broad band in the ultraviolet region (200-250 nm) with a maximum at
241 nm, which might be overlapping of charge transfer states (CTS) due to Dy3+ and
O2- interactions. The spectrum is composed of several sharp excitation lines originated
within the intra-configurational 4f transitions from the ground state 6H15/2 to excited
states of Dy3+ ions, peak centered at

355 nm (6H15/2 4M15/2 + 6P7/2) being the

prominent [13].

Fig.5.1.4. Photoluminescence excitation (PLE) spectrum of BaY1.92Dy0.08ZnO5


nanophosphors sintered at 1100oC, monitored with em = 580 nm
Table 5.1.2 listed the identification of all sharp excitation peaks observed in
the range 250 -500 nm for BaY1.92Dy0.08ZnO5 nanophosphor, sintered at 1100oC,
monitored with 580 nm as emission wavelength.
Table 5.1.2. Various excitation transitions in BaY1.92Dy0.08ZnO5 nanophosphors,
sintered at 1100oC, monitored with em = 580 nm.
Transitions
6

Wavelength (nm)

H15/24K13/2 + 4H13/2
6
H15/2 4K15/2
6
H15/24I9/2 + 4G9/2
6
H15/2 4M15/2 + 6P7/2
6
H15/2 4I11/2

300.5 nm
327.2 nm
339.2 nm
354.5 nm
367.8 nm

Transitions
6

Wavelength (nm)

H15/24I13/2
H15/2 M21/2 4I13/2 + 4K17/2 +
4
F7/2 6H15/24G11/2
6
H15/24I15/2
6
H15/24F9/2

190

388.5 nm
394.5 nm
427.8 nm
453.2 nm
467.8 nm

The photoluminescence (PL) spectra of BaY1.92Dy0.08ZnO5 nanophosphors assynthesized and sintered at different temperatures, measured with 355 nm as
excitation wavelength in 400-650 nm range is depicted in Fig. 5.1.5. The PL spectra
comprise of two main groups of lines in the blue region (450-525 nm) and yellow
region (530-650 nm). The blue emission peak at 490 nm and yellow emission peak at
580 nm are assigned to 4F9/26H15/2 and 4F9/26H13/2 transitions of Dy3+ ions,
respectively [27].

Fig.5.1.5. Photoluminescence (PL) spectra of BaY1.92Dy0.08ZnO5 nanophosphors


as-synthesized and sintered at different temperatures, monitored at ex = 355 nm
It is clearly apparent that the PL intensity of as-synthesized BaY1.92Dy0.08ZnO5
nanophosphor corresponding to both blue and yellow transitions increases rapidly on
increasing the sintering temperature from 800 to 1100oC due to reduction in nonradiative recombination effects, quenching sites and surface defects in the crystal
structure at higher temperature, leading to better crystallinity [2]. In all
BaY1.92Dy0.08ZnO5 samples, the relative PL intensity of yellow emission
(4F9/26H13/2) is found to be greater than that of blue emission (4F9/26H15/2) as Dy3+
ions are located at non-inversion symmetry sites (Cs) in BaY2ZnO5 lattice. This
phenomenon occurs because the blue emission belonging to magnetic dipole
transition is hardly influenced by crystal field symmetry of Dy3+ ions while yellow
emission corresponding to hypersensitive forced electric transition with the selection
191

rule, J = 2, get strongly affected by the outside surroundings. The crystal field
splitting of 4F9/26Hj can provide information regarding the site symmetry of Dy3+
ions with Kramers doublets (2J+1)/2 for each lattice site, where J is the angular
momentum of the electrons [32]. In a host lattice, when Dy3+ ions occupies noninversion symmetry site, the yellow emission is stronger while blue transition
dominates the emission spectrum on occupying a high symmetry local site (with
inversion center). The photoluminescence features of this nanophosphor confirms that
Dy3+ ions are located at 7-folded Y3+ low symmetry sites (Cs) in BaY2ZnO5 lattice as
suggested by XRD studies.
The PL spectra of BaY2(1-x)Dy2xZnO5 sintered at 1100oC as a function of
dysprosium doping concentration, measured at 355 nm as excitation wavelength are
illustrated in Fig. 5.1.6. No distinct emission spectral features of BaY2(1-x)ZnO5 :
2xDy3+ were observed on varying the x value from 0.5 to 7 mol% as the shape &
position remained the same. The emission spectra of all BaY2(1-x)Dy2xZnO5 samples
exhibits characteristic both blue (4F9/26H15/2) and yellow (4F9/26H13/2) emission of
Dy3+ ions, hypersensitive transition (4F9/26H13/2) being the stronger.

Fig.5.1.6. Photoluminescence (PL) spectra of BaY2(1-x)Dy2xZnO5 sintered at


1100oC, doped with different dysprosium contents, monitored at ex = 355 nm
192

The PL intensities of 4F9/26HJ (J = 15/2 and 13/2) transitions of the


nanophosphor enhances with increase of doping concentration and reaches maximum
at 4 mol% of Dy3+ ions contents in BaY2(1-x)ZnO5 nanophosphor. For 5 mol% of
dopant concentration, the relative emission intensity was decreased by nominal value
only, indicating concentration quenching at higher dysprosium contents. The cross
relaxation phenomenon among neighboring Dy3+ ions at higher concentration is the
most plausible explanation for such quenching behavior, which depopulates the 4F9/2
state of Dy3+ ions via non-radiative energy transfer processes i.e. Dy3+ (4F9/2) + Dy3+
(6H15/2) Dy3+ (4F9/2/6H7/2) + Dy3+ (6F3/2) [33]. Furthermore, the yellow to blue
emission ratio (Y/B) keeps an almost constant value of 1.05

0.02 over the range of

Dy3+ contents in BaY2ZnO5 nanophosphors. This leads to assume that the


hypersensitive electric forced transition (4F9/26H13/2) senses no considerable changes
in the crystal field symmetry of Dy3+ sites in BaY2ZnO5 lattice at different doping
concentrations.
The luminescence decay curves corresponding to 4F9/26H13/2 emission line at
580 nm in BaY2(1-x)ZnO5 nanophosphors, where x = 0.005 to 0.07, recorded under
excitation 355 nm (6H15/2 4M15/2 + 6P7/2) are shown in Fig. 5.1.7. All the decay
curves were found to obey single exponential behavior, represented by the equation I
=I0 exp (-t/), where is the radiative decay time, I and I0 are the luminescence
intensities at time t and 0, respectively. The calculated average lifetimes for 0.5, 1, 3,
4, 5 and 7 mol% of Dy3+ ions in BaY2(1-x)ZnO5 nanophosphors are 2.84, 2.78, 2.59,
2.38, 2.29 and 2.09 ms, respectively. The lifetime of BaY2(1-x)Dy2xZnO5
nanophosphors decreases with the increasing dopant concentration due to nonradiative energy transfer between optically active ions at higher contents, owing to
well known concentration quenching mechanism. The Commission International De
IEclairage (CIE) chromaticity coordinates for BaY2(1-x)ZnO5 nanophosphors doped
with 0.5 to 7 mol% dysprosium contents under 355 nm excitation were calculated and
represented in Fig. 5.1.8. BaY2(1-x)Dy2xZnO5 powders show excellent color (x, y)
coordinates of (0.279, 0.328), (0.303, 0.365), (0.311, 0.379), (0.323, 0.399), (0.339,
0.396) and (0.345,0.392) for 0.5, 1, 3, 4, 5 and 7 mol%, respectively in white region,
comparable to other standard color systems such as NTSC (0.3101, 0.3162),
PAL/SECAM/HDTV (0.3127, 0.329), ProPhoto/Color Match (0.3457, 0.3585) and
CIE white light point (0.33, 0.33) [25]. It is quite visible that BaY2(1-x)Dy2xZnO5
nanophosphor at low concentration exhibits cool white color, tending slightly towards
193

Fig.5.1.7. Luminescence decay curves of BaY2(1-x)Dy2xZnO5 nanophosphors,


sintered at 1100oC doped with different dysprosium contents, on monitoring
with ex = 355 nm and em = 580 nm

Fig.5.1.8. CIE color (x, y) coordinates for (a) 0.5 mol%, (b) 1 mol%, (c) 3 mol%,
(d) 4 mol%, and (e) 7 mol%, of Dy3+ ions in BaY2ZnO5 nanophosphors sintered
at 1100oC after excitation at 355 nm
194

yellowish white light at higher concentration. The precise emission color tuning on
CIE color coordinates makes BaY2(1-x)Dy2xZnO5 nanophosphors, as an ideal optical
material for UV-white LED applications.
SECTION II
Dy3+ doped LaSrAl3O7 Nanophosphors
Luminescent materials are extremely attractive for both fundamental research
and various technological applications including solid state lighting, ionizing
radiations, photonics, imaging, and biological assays [34-36]. In rare earth ions (RE3+)
based luminescent nanomaterials, doping is significantly important as they exhibit
high color purity flexible emission lines with different activators due to their intra 4f
transitions [37-38]. The optical characteristics of these materials generally depend on
the symmetry of local environment of RE3+, spin-orbit coupling of 4f electrons and
crystal field effects in the host lattice [6, 39]. In rare earth family, Dy3+ ions show
intense luminescence in blue and yellow spectral region due to transition from 4F9/2
energy level to 6H15/2 and 6H13/2 level, respectively. By suitably adjusting the relative
yellow to blue intensities ratios, Dy3+ doped nanophosphors may be potentially used
for development of white light emission [27-28]. In the last decade, as RE3+ doped
luminescent host materials, melilite oxides, ABC3O7 (A= Ca, Sr, Ba; B= Y, La, Gd; C
= Al, Ga) which can accept high dopant contents and show efficient emission spectra
suitable for high resolution optical display systems have been considerably envisioned
[1, 40-47]. These oxides possess tetragonal crystal structure with the space group P421m; made up of tetrahedral layered CO45- units and eight coordinated A2+ and B3+
ions having Cs symmetry, randomly distributed in between the layers [48-50].
For the nanocrystalline phosphors with optimized brightness, the prime area
concerned are stoichiometry, composition and surface chemistry as lattice defects that
acts as non-radiative relaxation channels and quenching sites may degrades the
luminescence efficiency [51]. Solution combustion synthesis (SCS) has opened new
vistas for the synthesis of highly pure powders having homogeneous and fine particles
with large surface area at low temperature in a very short duration as compared to
other conventional methods where impurity phase co-exists in the products of same
compositions even after long sintering time due to insufficient mixing and low
reactivity of raw materials [43-45]. LaSrAl3O7 is one of the low cost melilite

195

phosphor that could be easily synthesized like other family members using simple and
rapid solution combustion approach. This melilite oxide seems to be an ideal host for
doping of RE3+ ions due to its stable structural features, indicated by recent
investigations on luminescent properties of well crystallized Eu/Tb doped LaSrAl3O7
using SCS and citrate assisted sol-gel method [1, 42]. In the present work, herein we
proposed the doping of dysprosium ions in melilite oxide, LaSrAl3O7 via low
temperature solution combustion synthesis and studied its structural, morphological
and luminescent features by means of XRD, SEM, TEM, FT-IR and luminescent
spectra along with decay curves.
A. EXPERIMENTAL DETAILS
i) Powder synthesis
La1-xSrAl3O7: xDy3+ nanopowders, where x = 0.01 to 0.15 were synthesized by
urea

assisted

solution

combustion

process

using

high

purity

Sr(NO3)2,

La(NO3)3.6H2O, Al(NO3)3, Dy(NO3)3.6H2O and urea as starting materials. The


chemical reaction was:
1-xLa(NO3)3 + xDy(NO3)3 + Sr(NO3)2 + 3Al(NO3)3 + 11.66CH4N2O (Urea)

La1-xDyxSrAl3O7(s) + gaseous products.


According to nominal composition of La1-xDyxSrAl3O7, a stoichiometric
amount of metal nitrates were dissolved in minimum quantity of deionized water and
then urea was added as fuel. The amount of urea was calculated using total oxidizing
and reducing valencies according to the concept used in propellant chemistry [30].
Finally the beaker containing the aqueous paste was placed in a preheated furnace
maintained at 500oC. The homogenous solution of different oxidizers (metal nitrates)
and fuel (urea) undergo rapid and self-sustaining combustion process and the
chemical energy released during this exothermic redox reaction results in dehydration
and foaming followed by decomposition. Consequently, the large amounts of volatile
combustible gases generated alongwith flames, yields voluminous solid within 5-8
minutes. The product thus obtained was again sintered at 550oC for 1h in order to
eliminate unreacted nitrates, resulting in pure phased LaSrAl3O7: Dy3+ nanophosphor.
ii) Powder characterization techniques
X-ray diffraction (XRD) was performed on the Rigaku Ultima-IV X-ray
powder diffractometer using CuK radiation in the 2 range of 15-70o. The
196

accelerating voltage and current were at 40 kV and 40 mA. Fourier transform infrared (FT-IR) spectroscopy was carried out on Perkin-Elmer spectrometer using KBr
pellet technique in the spectral range 4000-400 cm-1. The morphology and particle
size were inspected using scanning electron microscope, Jeol JSM-6510 and
transmission electron microscope, Hitachi F-7500. The photoluminescence excitation
and emission spectra in the ultraviolet-visible region and decay curves under time
scan-mode were obtained on Hitachi F-7000 spectrofluorimeter equipped with Xelamp as the excitation source.
B. RESULTS AND DISCUSSION
i) X-ray Studies
LaSrAl3O7 belongs to the melilite oxide family and composed of AlO45tetrahedral layers with randomly distributed Sr2+ and La3+ ions having Cs symmetry
sites in between the layers. This melilite host comprises of tetragonal crystals with the
space group P-421m and lattice parameter a = 7.890Ao and c = 5.228Ao. The XRD
patterns of as-synthesized (500oC) and sintered (550oC, 1h) sample of
La0.90Dy0.10SrAl3O7 nanophosphor along with standard data of LaSrAl3O7 (JCPDS
No. 50-1815) are depicted in Fig.5.2.1a. The sharp diffraction peaks of as-synthesized
La0.90Dy0.10SrAl3O7 sample are well indexed to tetragonal melilite phase belonging to
JCPDS No. 50-1815, indicating that highly crystalline powders has been formed at
500oC. It implies that SCS provides the conditions of crystallization through
molecular level mixing of the precursors due to self sustained combustion of oxidizers

and urea at low temperature (500oC). Minor peak at 19.8 indicates the presence
nitrates residues in the as-synthesized La0.90Dy0.10SrAl3O7 sample. On sintering the
powders at 550oC for 1h, crystalline degree of the products is improved apparently as
peaks pertaining to unreacted nitrate phase disappeared completely although the
tetragonal melilite phase is formed even without sintering.
The XRD profiles of La1-xDyxSrAl3O7 (x = 1 to 15 mol%) nanophosphors
doped with different Dy3+ ions concentration, sintered at 550oC alongwith standard
reference data (JCPDS No. 50-1815) are presented in Fig. 5.2.1b. It can been clearly
seen that all the samples are well consistent with standard data of tetragonal
LaSrAl3O7 phase having space group P-421m belonging to JCPDS No. 50-1815. The
well defined diffraction peaks of all samples confirmed that La1-xDyxSrAl3O7 powders
maintain the pure tetragonal structure with little variation in dysprosium contents. As
197

Fig.5.2.1. XRD patterns of (a) La0.90Dy0.10SrAl3O7 nanophosphor as-synthesized


(500oC) and sintered (550oC, 1h); (b) La1-xDyxSrAl3O7 ( where, x = 1 to 15 mol%)
nanophosphors, sintered at 550oC alongwith standard reference data.
expected, Dy3+ ions (0.97 Ao) preferentially substitute Cs symmetry sites of La3+ ions
(1.16 Ao) rather than Sr2+ sites (1.26Ao) in LaSrAl3O7 lattice, on considering the
valence states and ionic radii difference.
198

The average crystallites size, D for La1-xDyxSrAl3O7 powders, was evaluated from
the full width half maxima of the most intense diffraction peak according to
Scherrers equation D 0.941/ cos, where is the wavelength of CuK radiation
(0.1548 nm), is the full width in radians at half-maximum (FWHM) and is the
Braggs angle of an observed X-ray diffraction peak. The values of particle size
calculated from their corresponding FWHM and diffraction angle (2) for 1, 5, 10 and
15 mol% of Dy3+ ions in La1-xSrAl3O7 powder, sintered at 550oC were found to be 65
nm, 69 nm, 72 nm and 74 nm, respectively.
ii) Morphological Studies
SEM image and TEM image of La0.90Dy0.10SrAl3O7 nanophosphor, sintered at
550oC are shown in Fig. 5.2.2 and 5.2.3, respectively. SEM image displays smooth
loosely aggregated tetragonal particles which are packed together by edge-to-edge
conjunctions, resulting in high surface area. Due to uncontrolled dynamics of solution
combustion process several pores are also apparent on the surface alongwith small
particles formed by escaping of large gaseous materials with high pressure. These
combustion synthesized powders with high surface area and porous network are
highly favorable for better luminescence [45]. TEM image reveals aggregated
tetragonal shaped particles with average size ranging between 65 nm and 75 nm. The
average particles size estimation nanophosphor was found to be consistent with that
determined using Scherrers equation.

Fig.5.2.2. SEM image of La0.90Dy0.10SrAl3O7 nanophosphor sintered at 550oC.

199

Fig.5.2.3. TEM image of La0.90Dy0.10SrAl3O7 nanophosphor sintered at 550oC


iii) FT-IR Studies

Fig. 5.2.4 shows the FT-IR spectrum of La0.90Dy0.10SrAl3O7 nanophosphors,


sintered at 550oC in the 4000-400 cm-1 range. IR bands in the region of 1000-400 cm-1
are associated to stretching and bending vibrations of AlO45- group and other M-O
bonds in the melilite lattice.

Fig. 5.2.4 FT-IR spectrum of La0.90Dy0.10SrAl3O7 nanophosphor sintered at 550oC


in the 4000-400 cm-1 range
200

The absorption peaks at 1630 cm-1 is assigned to bending vibrations of absorbing free
H2O while broad band in 2900 to 3700 cm-1 range can be attributed to stretching
vibrations of O-H group in the lattice. The characteristic peak around 1384 cm-1 due
to unreacted nitrates was not detected in the spectrum, which eventually confirm the
XRD results that impurity free pure melilite LaSrAl3O7 structure was obtained at
sintering temperature, 550oC.
iv) Luminescent Studies
The Photoluminescence excitation (PLE) spectrum of La0.90Dy0.10SrAl3O7
nanophosphor, sintered at 550oC, recorded at 574 nm emission wavelength for the
4

F9/2 6H13/2 transition is presented in Fig.5.2.5. It can be clearly seen that the

excitation spectrum comprises of a series of sharp peaks between 300 and 500 nm
attributed to characteristic 4f transitions of Dy3+ ions within its 4f9 configuration while
host related or O2- Dy3+ charge transfer band has not been noticed in short
wavelength region, suggesting weak interactions between dysprosium and oxygen in
the melilite lattice. These excitation peaks in longer wavelength region are assigned to
radiative transitions of Dy3+ ions from 6H15/2 energy state to 4K13/2 + 4H13/2 ,6P3/2 , 6P7/2,
6

P5/2, 4I13/2 , 4G11/2 , 4I15/2 and 4F9/2 at 301 nm, 328 nm, 352 nm, 366 nm, 384 nm, 428

nm, 451 nm and 467 nm respectively in LaSrAl3O7 lattice [11-12].

Fig.5.2.5. Photoluminescence excitation (PLE) spectrum of La0.90Dy0.10SrAl3O7


nanophosphor, sintered at 550oC, on monitoring emission at 574 nm.
201

The Photoluminescence spectra of La0.90Dy0.10SrAl3O7 nanophosphors assynthesized and sintered at 550oC, recorded with 352 nm excitation wavelength for
6

H15/2 6P7/2 in the range of 400-650 nm is shown in Fig.5.2.6. The spectra exhibit

two main emission; blue emission centered at 478 nm in the 450-500 nm region and
yellow emission centered at 574 nm in the range of 550-600 nm [52]. The blue
emission belongs to the magnetic allowed dipole (4F9/26H15/2) transitions of the Dy3+
ions, which gets hardly influenced by the crystal field strength around dysprosium
ions while yellow emission corresponds to forced electric allowed (4F9/26H13/2)
transitions with selection rule J = 2, being hypersensitive strongly affected by the
outside environment. If Dy3+ ions are located at high symmetry site with an inversion
centre, the 4F9/26H15/2 transition should be dominant; otherwise in a low symmetry
site with no inversion centre, the hypersensitive 4F9/26H13/2 transition will be
prominent in the emission spectra [53].

Fig.5.2.6. Photoluminescence (PL) spectra of La0.90Dy0.10SrAl3O7 nanophosphors


as-synthesized and sintered at 550oC, on monitoring excitation at 352 nm
As per the crystal structure principle, Dy3+ ions may easily enter into the low
symmetry sites of La3+ ions (Cs) in the LaSrAl3O7 structure on the basis of similar
valence state and smaller ionic radii that of La3+, hence for the both
202

La0.90Dy0.10SrAl3O7 nanophosphors forced electric transition (yellow emission) found


to be stronger than magnetic dipole transition (blue emission) as clearly visible in
Fig.5.2.6. In both samples, sintering do not induce any significant change in the shape
and positions of emission peaks although a slight increase in the intensity of assynthesized La0.90Dy0.10SrAl3O7 nanophosphor was observed at higher temperature
due to improvement in crystallinity.
The PL spectra of La1-xDyxSrAl3O7, where x = 1 to 15 mol% nanophosphors
sintered at 550oC, recorded at 352 nm as excitation wavelength are depicted in Fig.
5.2.7. For all La1-xDyxSrAl3O7 powders, the spectra are dominated by hypersensitive
yellow (4F9/26H13/2) emission of dysprosium ions. The ratio of yellow to blue
emission can be used as a spectroscopic probe to measure the degree of distortion of
luminescent center from the inversion symmetry in the host. The value of

emission ratio for different dysprosium contents in La1-xSrAl3O7 powders has come
out to be nearly 1.5, indicating that crystal field symmetry of Dy3+ ions in LaSrAl3O7
host does not vary with dopant concentration.

Fig.5.2.7. Photoluminescence (PL) spectra of La1-xDyxSrAl3O7, where x = 1 to 15


mol% nanophosphors sintered at 550oC, on monitoring excitation at 352 nm.
It has been seen that emission intensity due to both transitions increases with
the increasing Dy3+ ions concentration, and maximum intensity approaches at 10
203

mol% of dysprosium contents, then the intensity decreased due to concentration


quenching phenomenon. The possible explanation of this luminescence quenching
behavior is the non-radiative energy transfer from 4F9/2 state of Dy3+ ions via crossrelaxation process between two neighboring luminescent centers i.e. 4F9/2 (Dy3+) +
6

H15/2 (Dy3+) 4F9/2/6H7/2 (Dy3+) + 6F3/2 (Dy3+) [33]. The respective luminescence

decay curves for La1-xDyxSrAl3O7 nanophosphors at different Dy3+ ions


concentrations corresponding to yellow emission (4F9/2 6H13/2) at 574 nm, recorded
at ex = 352 nm are presented in Fig. 5.2.8. The decay curves corresponding to all
dysprosium ion contents can be well fitted into single exponential functions,
represented by the equation I =I0 exp (-t/), where is the radiative decay time, I and
I0 are the luminescence intensities at time t and 0, respectively. It shows Dy3+ ions
senses homogenous senses crystal field environment in LaSrAl3O7 lattice at different
doping concentrations. The average lifetimes determined are 2.61 ms, 2.41 ms, 2.38
ms, 2.23 ms, 2.07 ms and 1.98 ms for 1, 3, 5, 10, 12 and 15 mol% of Dy3+ ions,
respectively in La1-xSrAl3O7 nanophosphors.

Fig.5.2.8. Decay curves of La1-xDyxSrAl3O7, (x = 1 to 15 mol%) nanophosphors


sintered at 550oC, recorded at ex = 352 nm and em = 574 nm
As stated earlier, shape of emission curves and Y/B ratio does not vary much
over the range of dysprosium contents in LaSrAl3O7, hence the Commission
204

International De IEclairage chromaticity coordinates corresponding to different


concentrations of dysprosium fall in white region. At optimized concentration i.e
La0.90Dy0.10SrAl3O7 nanophosphor, sintered at 550oC exhibits color coordinates of x
=0.332 and y = 0.351 quite close to other standard color systems such as NTSC
(0.3101, 0.3162), PAL/SECAM/HDTV (0.3127, 0.329), ProPhoto/Color Match
(0.3457, 0.3585) and CIE white light point (0.33, 0.33).
SECTION III
Dy3+ doped GdSrAl3O7 Nanophosphors
As important rare earth oxides, melilite complexes (ABC3O7) have been
received considerable attention due to their stable structural features and high
application potential in solid state laser materials, plasma display panels, mercury
free- high discharge lamps (HID), high vision TV, plasma panel and scintillators [40,
46-47]. The melilite ABC3O7 structure, where A is the rare earth elements, B is the
alkaline earth elements and C is Al, Ga or In, is composed of eightfold coordinated
alternating (A/B)2 and cornersharing tetrahedral anionic CO45- layers. The trivalent
A and divalent B cations occupy Cs symmetry sites while trivalent C cations are
located at two non-equivalent tetragonal sites with S4 and Cs symmetries. It has been
found that the two dimensionally connected tetrahedral layered oxide network
stabilizes oxygen by local relaxation, leading to high conductivity of interstitial oxide
anions in the melilite lattice [49-50, 54]. Up to now, various synthetic routes have
been adopted to synthesize different RE3+ ions doped mellite oxide phosphors,
including solid state method, sol-gel process and combustion approach [1, 41-44, 45556]. As known, solid state process requires long sintering time ( 24h) to increase
diffusivity between precursor materials, resulting in agglomerated and coarse particles
which adversely affect the efficiency of phosphor [48,57]. In sol-gel method, post
sintering treatment at high temperature is needed to obtain the phosphor with single
phase and high luminescence intensity [1]. In order to improve these drawbacks,
solution combustion synthesis has been proved as an efficient, simple and self
sustained process due to high reactivity of raw materials at precursor level, which
yields not only nano-sized oxide powders with large surface area but also allows
homogenous doping of trace amounts of RE3+ ions in host lattice at low sintering
temperature [29, 39, 45].
205

Among the melilite oxides, gadolinium strontium aluminate GdSrAl3O7


having tetragonal crystal structure with space group P-421m, could act as both
efficient host lattice and sensitizer due to presence of Gd3+ as constituent ion, and
offer excellent luminescent potentialities when doped with rare-earth ions (RE3+) like
other isostructural melilite oxides, GdCaAl3O7 and GdSrGa3O7 [44, 56]. Recently, the
doping of Tb3+ ions in GdSrAl3O7 lattice has been investigating by Zhou et. al. [55]
via EDTA sol-gel process, reporting the pure melilite phase formation at higher
sintering temperature (900oC) for 5h but luminescent properties of dysprosium ions in
this host has not been studied so far. In the present work, Dy3+ doped GdSrAl3O7
nanocrystalline phosphor were synthesized at low temperature using effective and
rapid solution combustion approach. Dy3+ ions doped phosphor holds a great promise
in white light emission by proper tuning of yellow to blue emission intensities on
varying the composition of host. Herein, the structural and photoluminescence
features of Dy3+ doped GdSrAl3O7 nanophosphor with energy transfer mechanism
from Gd3+ to Dy3+ ions were also investigated in details.
A. EXPERIMENTAL DETAILS
i) Powder Synthesis
Gd1-xSrAl3O7: xDy3+ nanopowders, where x = 0.01 to 0.15 were synthesized
using urea assisted solution combustion process. The chemical equation for the
reaction was:
1-xGd(NO3)3 + xDy(NO3)3 + Sr(NO3)2 + 3Al(NO3)3 + 11.66CH4N2O (urea)
Gd1-xDyxSrAl3O7(s) + gaseous products.
According to stoichiometric composition Gd1-xDyxSrAl3O7, high purity raw
materials Sr(NO3)2, Gd(NO3)3.6H2O, Al(NO3)3, Dy(NO3)3.6H2O and urea were
dissolved in minimum quantity of deionized water. The amount of urea was
calculated using total oxidizing and reducing valencies according to the concept used
in propellant chemistry [30]. Finally the beaker containing the aqueous paste was
placed in a preheated furnace maintained at 500oC. The mixture of metal nitrates
(oxidizers) and fuel (urea) undergo rapid and self-sustaining combustion process and
the chemical energy released during this exothermic redox reaction results in
dehydration and foaming followed by decomposition. Consequently, the large
amounts of volatile combustible gases generated alongwith flames, yields voluminous
solid within 5-8 minutes. The foam thus obtained was again sintered at 550oC for 1h
206

in order to eliminate unreacted nitrates, resulting in single phased dysprosium doped


GdSrAl3O7 nanophosphor.
ii) Powder Characterization Techniques
To reveal the crystalline phase of Gd1-xDyxSrAl3O7 powders, X-ray diffraction
(XRD) was carried out on Rigaku Ultima-IV X-ray powder diffractometer at 40 kV
tube voltage and 40 mA tube current with CuK radiation in the 2 ranging from15 to
70o. The functional groups in GdSrAl3O7 lattice were identified by the Fourier
transform infra-red spectroscopy (Perkin-Elmer spectrometer) in the spectral range
4000-400cm-1. The morphology and particle size were evaluated using scanning
electron microscopy (SEM) on Jeol JSM-6510 and transmission electron microscopy
(TEM) on Hitachi F-7500. The photoluminescence excitation and emission spectra of
the nanophosphor in the ultraviolet-visible region and decay curves under time scanmode were analyzed by fluorescence spectrophotometer (Hitachi F-7000) equipped
with Xe-lamp as the light source.
B. RESULTS AND DISCUSSION
i) X-ray Studies
GdSrAl3O7 is one of the complex oxide of melilite family, consisting of
tetragonal crystals with the space group P-421m and lattice parameter a = 7.801Ao and
c = 5.132Ao. The GdSrAl3O7 lattice is built up of AlO45- tetrahedral layers and the Sr2+
and Gd3+ ions are randomly distributed in between the layers at Cs symmetry sites.
The XRD patterns of Gd0.90 Dy0.10SrAl3O7 nanophosphor, as-synthesized (500oC) and
sintered (550oC, 1h) along with standard data of melilite oxide, GdSrAl3O7 (JCPDS
No. 50-1817) are presented in Fig.5.3.1a. In the as-synthesized sample, diffraction
peaks at 16.09 (001), 17.30 (111), 23.94 (111), 25.65 (210), 28.84 (201), 31.04 (211),
35.09 (002), 36.50 (310), 43.89 (212) and 51.46 (312) belonging to tetragonal phased

GdSrAl3O7 (JCPDS No. 50-1817) with a shoulder peak marked as * at 19.8 due to
unreacted nitrates were observed. It is quite visible that the sintering of
Gd0.90Dy0.10SrAl3O7 powder at 550oC does not induce any significant phase change so
far except the impurity peak gets disappeared completely. The XRD patterns indicate
that well-crystallized melilite phased GdSrAl3O7: Dy3+ nanophosphor could be easily
obtained under the low temperature synthetic conditions. Hence, it is very easy to
conclude that SCS is an effective approach which allows homogenous mixing of
207

different oxidizers (i.e. metal nitrates) at molecular level during pre-sintering process
due to self sustained combustion of metal nitrate and an organic fuel at low furnace
temperature (500oC).
The XRD profiles of Gd1-xDyxSrAl3O7 nanophosphors doped with different
dysprosium contents, sintered at 550oC alongwith standard reference data (JCPDS No.
50-1817) are depicted in Fig. 5.3.1b. The well defined diffraction peaks confirmed
that, after sintering at 550oC, powder samples crystallize in a pure tetragonal
GdSrAl3O7 phase (JCPDS No. 50-1817). It also implies that the incorporation of Dy3+
ions does not cause any distortion in the melilite lattice as Dy3+ ions (RDy3+ = 0.97 Ao)
can easily enter into Cs symmetry sites for substituting Gd3+ ions (RGd3+ =1.00 Ao) in
GdSrAl3O7. The average particle size, D of Gd1-xDyxSrAl3O7 powders, was
determined from the XRD parameters according to Scherrers equation D 0.941/
cos, where is the wavelength of CuK radiation (0.1548 nm), is the full width in
radians at half-maximum (FWHM) and is the Braggs angle of an observed X-ray
diffraction peak. From the FWHM value of the most intense peak (211), the
calculated particle size corresponding to 1, 5, 10 and 15 mol% of Dy3+ ions in Gd1xSrAl3O7

powder, sintered at 550oC came out to be 49 nm, 48 nm, 45 nm and 42 nm,

respectively. Within all samples, the average size standard deviation was 45

4 nm,

indicating that the amount of Dy3+ ions has no remarkable influence on the particle
size of Gd1-xDyxSrAl3O7 nanophosphors.

208

Fig.5.3.1. XRD patterns of (a) Gd0.90 Dy0.10SrAl3O7 nanophosphors, assynthesized and sintered at 550oC ; (b) of Gd1-xDyxSrAl3O7 nanophosphors (x = 1
to 15 mol%), sintered at 550oC alongwith standard data of GdSrAl3O7
ii) Morphological Studies
The morphological features of Gd0.90Dy0.10SrAl3O7 nanophosphor, sintered at
550oC were investigated by scanning electron microscopy (Fig. 5.3.2) and
transmission electron microscopy (Fig. 5.3.3).

Fig.5.3.2. SEM image of Gd0.90Dy0.10SrAl3O7 nanophosphor sintered at 550oC.


209

Fig.5.3.3. TEM image of Gd0.90Dy0.10SrAl3O7 nanophosphor sintered at 550oC.


SEM image of Gd0.90Dy0.10SrAl3O7 nanophosphor depicts smooth and densely
packed tetragonal porous particles, as expected arise from the non-uniform
distribution of temperature and mass flow in the combustion flame (Fig. 5.3.2). The
combustion derived products with high surface area and porous network are believed
to be an outcome of large amount of escaping gaseous materials during the selfsustained combustion process [58-59]. TEM image of Gd0.90Dy0.10SrAl3O7
nanophosphor sintered at 550oC, shows monodispersed tetragonal shaped particles
with sizes in 40-50 nm range (Fig.5.3.3). The average particles size estimation of
Gd0.90Dy0.10SrAl3O7 nanophosphor is consistent with that determined from the XRD
patterns.
iii) FT-IR Studies
The FT-IR spectrum of Gd0.90Dy0.10SrAl3O7 nanophosphor, sintered at 550oC in
the frequency region of 4000-400cm-1 is depicted in Fig. 5.3.4. Modes in the lower
region (1000-400cm-1) are assigned to the aluminum-oxygen stretching and bending
vibrations of AlO45- tetrahedron and other metal-oxygen bonds present in the melilite
structure. In addition, absorption bands related to characteristics H-O-H bending
vibrations of absorbing free H2O (1630 cm-1) and O-H stretching vibration (3438cm-1)
also appeared in the sample. It is significant to note that the fundamental residual
NO3- peak around 1384 cm-1 was not detected in the spectrum, confirming the
formation of impurity free pure melilite phased structure at low crystallization
temperature, 550oC.
210

Fig.5.3.4. FT-IR spectrum of Gd0.90Dy0.10SrAl3O7 nanophosphor sintered at


550oC
iv) Luminescent Studies
The photoluminescence excitation (PLE) spectrum of Gd0.90Dy0.10SrAl3O7
nanophosphor, sintered at 550oC, monitored with 574 nm (4F9/2 6H13/2) as emission
wavelength is illustrated in Fig. 5.3.5. The excitation spectrum monitored at yellow
emission consists of several characteristic sharp peaks corresponding to intra-4f
transitions of Gd3+ and Dy3+ in the melilite host. Among these excitations, the peaks
at 274 nm and 312 nm are assigned to transitions of Gd3+ ions from 8S7/2 state to 6I7/2
and 6P7/2 states while the peaks in the 320 -500 nm correspond to ff transitions of
Dy3+ ions within its 4f9 configuration. These excitation peaks in longer wavelength
region at 322 nm, 350 nm, 385 nm, 423 nm, 452 nm and 463 nm are ascribed to
6

H15/26P3/2, 6H15/2 6P7/2, 6H15/2 4I13/2, 6H15/24G11/2, 6H15/24I15/2 and 6H15/2

F9/2 transitions, respectively of Dysprosium ions in GdSrAl3O7 host [11]. The host or

O2- Dy3+ charge sensitized luminescence was not observed in the short wavelength
region indicating weak Dy3+ ions interactions with melilite host while presence of
dominant 8S7/26I7/2 and 8S7/26P7/2 electronic transitions of Gd3+ ions in the
spectrum confirms the efficient energy transfer between Gd3+ and Dy3+ ions in host
lattice [60].
211

Fig.5.3.5. Photoluminescence excitation (PLE) spectrum of Gd0.90Dy0.10SrAl3O7


nanophosphor, sintered at 550oC, monitored with em = 574 nm
The photoluminescence (PL) spectra of Gd0.90Dy0.10SrAl3O7 nanophosphor assynthesized and sintered at 550oC, monitored with 350 nm (6H15/26P7/2) as excitation
wavelength in 400-650 nm region is depicted in Fig. 5.3.6. The two main emission
bands in the 450-500 nm (blue region) and 550-600 nm (yellow region) corresponding
to transitions between well-defined 4f energy states of the Dy3+ ions were observed.
The blue emission containing multiple emission lines, centered at 478 nm and
dominating yellow emission centered at 574 nm are assigned to 4F9/26H15/2 and
4

F9/2 6H13/2 transitions of Dy3+ ions, respectively. With respect to Dy3+ ions,

magnetic allowed dipole 4F9/26H15/2 transition gets hardly influenced by the crystal
field symmetry of dysprosium ions while forced electric allowed transition
4

F9/26H13/2 being hypersensitive appreciably affected by chemical surrounding of the

luminescent center [62].


Both samples exhibits stronger forced electric transition (yellow emission)
than magnetic dipole transition (blue emission) as there is a high probability of
substitution of Dy3+ ions into low inversion symmetric Gd sites (Cs) in this melilite
oxide due to comparable size of Dy3+ and Gd3+ ions. However, the PL spectrum of
212

as-synthesized Gd1.90SrDy0.10Al3O7 nanophosphor shows slightly weak emission with


maintained shape and positions of peaks corresponding to both transitions as
compared to that of sample, sintered at 550oC (1h). This indicates that
Gd1.90SrDy0.10Al3O7 nanophosphors are not well crystallized at 500oC due to presence
of surface impurities in host lattice which may be omitted out after sintering, leading
to improvement in crystallinity as confirmed by the XRD measurements.

Fig.5.3.6. Photoluminescence spectrum of Gd0.90Dy0.10SrAl3O7 nanophosphor assynthesized and sintered at 550oC, monitored with ex= 350 nm
The PL spectra of Gd1-xDyxSrAl3O7 nanophosphors, sintered at 550oC with
dysprosium doping contents ranging from 1 to 15 mol%, monitored with 350 nm as
excitation wavelength are shown in Fig. 5.3.7. For all Gd1-xDyxSrAl3O7 samples, the
hypersensitive yellow (4F9/26H13/2) emission of Dy3+ ions is the prominent one as
Dy3+ ions preferentially substitutes low symmetry Gd3+ sites (Cs) with no inversion
center in GdSrAl3O7 lattice. The relative PL intensities corresponding to both yellow
and blue emission enhanced with the increasing concentration of Dy3+ ions, reaching
the maximum at 10 mol% of dysprosium contents while decreased with the further
increase of dopant concentration. Such behavior is due to cross-relaxation between

213

luminescent centers [33], caused by quenching of the energy state 4F9/2 at higher
dysprosium contents via non-radiative energy transfer process such as;
Dy3+ (4F9/2) + Dy3+ (6H15/2) Dy3+ (4F9/2/6H7/2) + Dy3+ (6F3/2)
It has been noticed that value of yellow to blue emission ratio
(

does not show significant changes with the increasing Dy3+ contents in

GdSrAl3O7 host, indicating hypersensitive electric forced transition (4F9/26H13/2)


senses same crystal field environment at Dy3+ symmetry sites in GdSrAl3O7 lattice at
different doping concentrations.

Fig.5.3.7. Photoluminescence spectra of Gd1-xDyxSrAl3O7 nanophosphors doped


with different Dy3+ contents, sintered at 550oC and monitored with ex = 350 nm
The luminescence decay curves corresponding to 4F9/2 6H13/2 transition at 574
nm for Gd1-xDyxSrAl3O7 nanophosphors in terms of Dy3+ ions concentrations,
monitored with 350 nm excitation wavelength are shown in Fig. 5.3.8. In all samples,
4

F9/26H13/2 transitions show single exponential behavior, represented by the equation

I =I0 exp (-t/), where is the radiative decay time, I and I0 are the luminescence
intensities at time t and 0, respectively. The calculated lifetimes are 2.98, 2.77, 2.65,
2.54, 2.31 and 2.19 for 1, 3, 5, 10, 12 and 15 mol% of Dy3+ ions, respectively. At
higher concentration, the distance between the dysprosium ions shortens subsequently
non-radiative energy transfer between optical active ions become more frequent,
hence life time decreases with the increasing dopant concentration.
214

Fig.5.3.8. Decay curves of Gd1-xDyxSrAl3O7 nanophosphors doped with different


Dy3+ contents, sintered at 550oC and monitored with ex = 350 nm
As stated earlier, shape of emission curves and Y/B ratio does not vary much
over the range of dysprosium contents in GdSrAl3O7, hence the Commission
International De IEclairage chromaticity coordinates corresponding to different
concentrations of dysprosium fall in yellowish-white region, with color coordinates of
x =0.361 and y = 0.436 for Gd0.90Dy0.10SrAl3O7 nanophosphor, sintered at 550oC
which is found to be better than that of other reported Dy3+ doped phosphor such as
Gd2O3: Dy3+ (0.41, 0.44) phosphor [62].
SECTION IV
Dy3+ doped CaZrO3 Nanophosphors
Metal oxides ABO3 (where A = Ca, Pb, Sr, Ba, Zn, Ni, Fe; B = Ti, Zr, Si, Hf)
with perovskite structure have attracted significant attention owing to their
superconductivity and remarkable electrical properties such as ferroelctricity and
piezoelectricity [63-67]. Alkaline-earth zirconates are well known for their structural
diversity including high thermal and chemical stability, single phase crystalline
structure, high refractive index and wide band gap which results in their potential
technological applications in electronic ceramic industry, gas sensors, optical
coatings, filters and so on [68-70]. Among the perovskite zirconates CaZrO3, seems to
215

be considerably important for both mechanical and electrical applications due to its
high permittivity, insulation resistance, high ionic conductivity, good corrosion
resistance, hydrocarbon sensing and high thermal sensitivity [71-78]. In addition,
ABO3 perovskites doped with acceptor-ions exhibits proton conductivity at high
temperature, making them suitable for electrochemical devices [79].
A perovskite ABO3 lattice consists of slightly deformed BO6 octahedra with BO-B angles at 180o and 12-oxygen coordinated A2+ cations with m3m symmetry at
dodecahedral site. Calcium zirconate oxide appears to be potential host matrix as it
possesses most distorted structure favorable for the doping of trivalent rare earth ions
(RE3+) [80]. Rare earth ions are known for their excellent luminescent properties
when doped in different inorganic host lattices. Among the RE3+ ions, Dy3+ ions
exhibits strong luminescence in a variety of host lattices and show characteristic both
blue (4F9/2 6H15/2) and yellow (4F9/26H13/2) emissions, necessary for generation of
white light very useful for high resolution optical display system [25-28]. However,
only limited information regarding the doping of dysprosium ions in calcium
zirconate perovskite oxide are available in the literature. In the past, sol-gel process
has been exploited to synthesize CaZrO3:Dy3+ phosphor and the photoluminescence
properties of Dy3+ and Tm3+ co-activated CaZrO3 phosphor were investigated [82-83].
But no reports on the solution combustion synthesis of Dy3+ doped CaZrO3 nanocrystalline phosphors has come into the notice. The present work reports synthesis
and characterization of CaZrO3: Dy3+ using urea assisted solution combustion
process. Moreover, dependence of structural and photoluminescence properties of
these nanophosphors on various parameters such as temperature and dysprosium
stoichiometry have also been studied in details.
A. EXPERIMENTAL
i) Powder Synthesis
Ca1-xZrO3: xDy3+ nanophosphors were synthesized by solution combustion
method using high purity Ca(NO3)2.4H2O, ZrN2O7, Dy(NO3)3.6H2O and urea as
starting reagents. The chemical equation for the reactions is:
(1-x)Ca(NO3)2 + xDy(NO3)3 + ZrN2O7 + ~3.5 CH4N2O Ca1-xDyxZrO3 (s) + gaseous
products.
According to nominal composition of Ca1-xZrO3: xDy3+ (x = 0.01 to 0.05), a
stoichiometric amount of metal nitrates were dissolved in minimum quantity of
216

deionized water in 200 mL capacity pyrex beaker. Then urea was added in this
solution with molar ratio of urea to oxidizer based on total oxidizing and reducing
valencies of oxidizer and fuel (urea) according to concept used in propellant
chemistry [30]. This aqueous paste containing calculated amount of metal nitrates and
urea was then placed in a preheated furnace maintained at 500oC. The mixture of
metal nitrates (oxidizers) and fuel (urea) undergo rapid and self-sustaining
combustion process and the chemical energy released during this exothermic redox
reaction results in

dehydration and foaming followed by decomposition.

Consequently, the large amounts of volatile combustible gases generated alongwith


flames, yields voluminous solid within 5-8 minutes. The powders obtained were again
fired from 800C to 1200C for 3h in order to evaluate the effect of sintering on the
structural and luminescent features on Ca1-xZrO3: xDy3+ nanophosphors.
ii) Powder Characterization Techniques
Crystal phase and particle size of CaZrO3: Dy3+ powders were examined by
Rigaku Ultima-IV X-ray powder diffractometer with CuK radiation to record the
patterns in 2 range of 20-65. Surface morphology was evaluated using Jeol JSM6510 scanning electron microscope. The surface morphology of the samples was
examined using Jeol JSM-6510 scanning electron microscope (SEM). The crystallite
size and shape has been evaluated using Tecnai G2 transmission electron microscope
(TEM). The photoluminescence excitation and emission in the ultraviolet-visible
region and decay curves under time scan mode were carried out on Hitachi F-7000
fluorescence spectrophotometer equipped with Xe-lamp at room temperature.
B. RESULTS AND DISCUSSION
i) X-ray Studies
CaZrO3 is one of the most distorted orthorhombic perovskite structures where
divalent Ca cation occupies distorted dodecahedral site. In this perovskite lattice,
rotation of slightly deformed ZrO6 octahedra distorts the Ca2+ site causing reduction
in its symmetry site from Pm3m to 1 along with decrease in coordination from 12 to 8
[82]. On introducing in CaZrO3 host, the dopant Dy3+ (0.092 nm) has a single choice
between substitution sites of Ca2+ ion (0.099 nm) and Zr4+ (0.072 nm) due to
comparable ionic size difference, hence preferably substitute Ca2+ site in this lattice.
XRD patterns of Ca0.96Dy0.04ZrO3 nanophosphor as-synthesized and sintered at
217

different temperatures from 800C to 1200C alongwith the standard JCPDS No. 350790 are represented in Fig.5.4.1a. At 1200C, Ca0.96Dy0.04ZrO3 nanophosphor
crystallizes in pure orthorhombic pervoskite structure having space group Pnma (62)
and lattice parameters a = 5.755, b = 8.010 and c = 5.592 belonging to JCPDS No.
35-0790. The main diffraction peaks at 22.12 (101), 24.76 (111), 31.02 (200),
31.49 (121), 31.92 (002), 37.27 (031), 45.13 (202), 50.17 (301), 50.81 (222),
51.49 (103), 55.47 (321), 55.79 (240), 56.38 (042), 56.64 (123) and 65.74 (242)
due to orthorhombic CaZrO3 lattice were observed. No traces of other peaks due to
additional phases apart from single pervoskite CaZrO3 phase was detected at this
temperature. While at lower temperatures, weak reflex lines of ZrO2 (JCPDS No.491746) marked as * were also apparent alongwith all the main diffraction peaks of
CaZrO3. The XRD profiles of the Ca0.96Dy0.04ZrO3 nanophosphors at lower
temperatures (800C and 1000C) indicated a high degree of peak broadening which
shows that sample is amorphous with only weak evidence of crystallinity. It is clearly
seen that on increasing the temperature, the intensity of main peak (121) plane
enhanced with the decrease in line width, concluding that the crystalline degree of
dysprosium doped CaZrO3 nanocrystals is improved apparently with the rise in
sintering temperature.
The XRD patterns of Ca(1-x)DyxZrO3 powders sintered at 1200C, doped with
different contents of dysprosium ions alongwith the standard JCPDS No. 35-0790 are
depicted in Fig. 5.4.1b. It is quite evident that all the samples exhibit single
orthorhombic CaZrO3 phase belonging to space group Pnma, indicating small amount
of doped Dy3+ ions had no influence on crystal structure of this lattice. The crystallite
size, D of Ca(1-x)ZrO3 : xDy3+ powders was estimated by Scherrers formula, D
0.941/ cos, where is the wavelength of CuK radiation (0.1548 nm), is the full
width in radians at half-maximum (FWHM) and is the Braggs angle of an observed
X-ray diffraction peak. The calculated average crystallite sizes, by taking main peak
(121) of Ca0.96Dy0.04ZrO3 powders came out to be 45.3 nm, 53.2 and 64.5 nm due to
pervoskite CaZrO3 phase as-synthesized, at sintering temperatures 800C, 1000C and
1200C, respectively. It was observed that all the powders are in nano-regime range
although an increase in particle size with the rise in sintering temperature was noticed
due to enhanced atomic mobility of particles which leads to faster grain growth at
higher temperature.

218

Fig.5.4.1. XRD patterns of (a) Ca0.96Dy0.04ZrO3 powders sintered at various


temperatures (b) Ca1-xDyxZrO3 ( x = 1 to 5 mol%), sintered at 1200C along with
standard data of CaZrO3 (JCPDS no. 35-0790).
219

ii) Morphological Characteristics


SEM and TEM studies were carried out to investigate the morphology and
particle size of the dysprosium doped CaZrO3 nanophosphors. SEM image of
Ca0.96Dy0.04ZrO3 powder sintered at 1200C clearly depicts narrow size distribution of
slightly agglomerated spherical shaped particles (Fig.5.4.2).

Fig.5.4.2. SEM micrographs of Ca0.96Dy0.04ZrO3 powders sintered at 1000oC


TEM image of Ca0.96Dy0.04ZrO3 powder sintered at 1200C also show smooth slightly
agglomerated spherical morphology particles having size in diameter range 50 - 60
nm (Fig.5.4.3). The particle size estimation from TEM studies is well consistent with
that of XRD results.

Fig.5.4.3. TEM micrograph of Ca0.96Dy0.04ZrO3 powders sintered at 1200oC.

220

iii) Luminescent Studies


The photoluminescence excitation (PLE) spectrum of Ca0.96Dy0.04ZrO3
nanophosphors sintered at 1200oC, monitored with yellow emission, 576 nm
corresponding to 4F9/26H13/2 transition in 200-500 nm range is shown in Fig. 5.4.4.
The PLE spectrum can be divided into a very weak broad band in the ultraviolet
region (200-275 nm) with a maximum at 258 nm, which might be overlapping of
attributed to charge transfer states (CTS) due to O2- Dy3+ interactions and host
absorption band (HAB) and a series of sharp excitation lines in the 320-500 nm range,
which are assigned to the intra- 4f transitions of the Dy3+ ions. These excitation peaks
in longer wavelength region are located at 326, 353, 366, 388, 427, 453 and 470 nm
corresponding to transitions of Dy3+ ions from 6H15/2 energy state to 6P3/2 , 6P7/2, 6P5/2,
4

I13/2 , 4G11/2

I15/2 and 4F9/2 respectively, peak with maxima at 353 nm being the

dominating [53].

Fig.5.4.4. Photoluminescence excitation (PLE) spectrum of Ca0.96Dy0.04ZrO3


nanophosphors sintered at 1200oC, monitored with em = 576 nm
The photoluminescence (PL) spectra of Ca0.96Dy0.04ZrO3 nanophosphors assynthesized and sintered at different temperatures, monitored at 353 nm excitation in
425-675 nm region is depicted in Fig. 5.4.5. The PL spectra consist of two sharp
emission lines in the blue region (470-490 nm) and yellow region (530-650 nm). The
221

blue emission centered at 484 nm corresponds to 4F9/26H15/2 while yellow emission


with maxima at 576 nm is ascribed 4F9/26H13/2 transitions of Dy3+ ions, respectively
[27].

Fig.5.4.5. Photoluminescence (PL) spectra of Ca0.96Dy0.04ZrO3 nanophosphors


sintered at different temperatures, monitored with ex = 353 nm
It is quite apparent that relative PL intensity of yellow emission (4F9/26H13/2)
corresponding to forced electric transition is stronger than that of blue emission
(4F9/26H15/2) belonging to magnetic dipole transition. Being hypersensitive yellow
emission with the selection rule, J = 2, get greatly influenced by the outside
surroundings while blue emission is hardly influenced by crystal field symmetry of
Dy3+ ions [32]. Increase in PL intensity of as-synthesized Ca0.96Dy0.04ZrO3
nanophosphors corresponding to both blue and yellow transitions with the rise in
sintering temperature from 800 to 1100oC is quite visible. This may be attributed to
reduction in non-radiative recombination effects, quenching sites and surface defects
in the crystal structure at higher temperature, hence leading to improvement in doping
and better crystallinity [2].
The PL spectra of Ca1-xDyxZrO3 doped with different Dy3+ contents sintered at
1200C, monitored at 353 nm excitation are illustrated in Fig. 5.4.6. Variable contents
of dysprosium ions have no distinct influence on the shape and positions of emission
222

lines in PL spectra. In all Ca1-xDyxZrO3 samples, hypersensitive yellow (4F9/26H13/2)


emission was found to be greater than blue (4F9/26H15/2) emission. However,
increase in the relative PL intensities corresponding to both transitions with increasing
dysprosium concentration was noticed. The maximum emission intensity was
observed at 4 mol% of Dy3+ ions contents in Ca1-xDyxZrO3 nanophosphors and
decreased afterwards owing to well-known concentration quenching phenomenon.
This concentration quenching may be attributed to the non-radiative cross relaxation
process between adjacent Dy3+ ions at higher concentration, causing depopulation of
the 4F9/2 energy state of Dy3+ ions as; Dy3+ (4F9/2) + Dy3+ (6H15/2) Dy3+ (4F9/2/6H7/2)
+ Dy3+ (6F3/2) [33]. In addition, constant yellow to blue emission ratio (Y/B

) with

the increasing Dy3+ contents in CaZrO3 lattice leads to assume that the hypersensitive
electric forced transition (4F9/2 6H13/2) senses no remarkable variation in the crystal
field symmetry of Dy3+ ions in this pervoskite CaZrO3 lattice.

Fig.5.4.6. Photoluminescence (PL) spectra of Ca1-xDyxZrO3 powders doped with


different dysprosium contents sintered at 1200oC and monitored at ex = 353 nm.

223

Fig.5.4.7. Decay curves Ca1-xDyxZrO3 powders doped with different dysprosium


contents sintered at 1200oC, monitored at ex = 353 nm and em = 576 nm
The luminescence decay curves for Ca1-xDyxZrO3 nanophosphors where x =
0.01 to 0.05 corresponding to yellow emission (4F9/26H13/2) monitored at 353 nm
excitation (6H15/2 6P7/2) are displayed in Fig. 5.4.7. All the decay curves were found
to be mono- exponential, represented by the equation I =I0 exp (-t/), where is the
radiative decay time, I and I0 are the luminescence intensities at time t and 0,
respectively. The calculated average lifetimes for 0.5, 1, 3, 4, 5 and 7 mol% of Dy3+
ions in Ca1-xDyxZrO3 nanophosphors are 1.88 ms, 1.82 ms, 1.71 ms, 1.57 ms and 1.50
ms, respectively. This decrease in average lifetimes with the increasing dysprosium
contents may also be due to enhanced non-radiative energy transfer between dopant
ions at higher concentration. The Commission International De IEclairage (CIE)
chromaticity coordinates for Ca1-xDyxZrO3 nanophosphors doped with 0.1 to 5 mol%
dysprosium contents were calculated from their corresponding PL spectra, monitored
with 353 nm excitation and represented in Fig. 5.4.8. The color coordinates (x, y) of
Ca1-xDyxZrO3 samples located at (0.360, 0.371), (0.330, 0.348), (0.333, 0.365),
(0.348, 0.379) and (0.367, 0.366) for 1, 2,3, 4 and 5 mol%, respectively are found to
be very close to the standard color systems such as ProPhoto/Color Match (0.3457,
0.3585) and CIE white light point (0.33, 0.33) [25]. It is quite apparent that white light
is slightly tending towards yellowish hue with the increasing dysprosium
concentration in Ca1-xZrO3 nanophosphors.
224

Fig.5.4.8. CIE color (x, y) coordinates for (a) 2 mol%, (b) 3 mol%, (c) 4 mol%,
and (e) 5mol%, of Dy3+ ions in Ca1-xDyxZrO3 nanophosphors sintered at 1200oC
after excitation at 353 nm
SECTION V
Dy3+ doped BaZrO3 Nanophosphors
Barium zirconate having pervoskitetype structure possess an extraordinary
potential for applications in different fields as thermal barrier coating (TBC) for
supersonic air jets [83], material for interface engineering of alumina fiber composites
[84-85] and proton conducting material in steam electrolyzer, humidity sensor and
especially in solid oxide fuel cells (SOFCs) [87-88]. All these promising applications
of this excellent refractory ceramic are attributed to its good mechanical strength, high
thermal stability, low chemical reactivity with corrosive materials, high melting point
(2600 oC) and low coefficient of thermal expansion ( = 87

10-7 IoC) [89-91].

BaZrO3 has cubic structure belonging to P23 (195) space group in which Zr atoms are
bonded to six oxygen atoms forming octahedral ZrO6 units while Ba atoms are bonded
to 12 oxygen atoms in cuboctahedral manner as BaO12. Within the ZrO6 octahedron,
zirconium atom resides in centrosymmetric positions [90].

225

Nowadays, intensive research has been made to explore potential of BaZrO3


host matrix because doping of divalent or trivalent rare earth ions is quite easy in this
single phased crystalline material. Rare earth (RE) doped barium zirconate is well
known for light emission in the visible region via up-conversion or down-conversion
mechanism. Several routes are reported in the literature including solid-state reaction,
sol-gel, hydrothermal reaction, vapor phase synthesis and microwave assisted
hydrothermal reaction for the synthesis of barium zirconate doped with different RE
ion. Recently, Borja-Urby et al. [92] have exploited wet synthesis hydrothermal
method to prepare Ce, Eu, Dy, Er, Yb doped BaZrO3 phosphors. In the present report,
Dy3+ doped BaZrO3 nanophosphors have been synthesized adopting solution
combustion synthesis (SCS) as well as their structural, morphological and
luminescent features were also investigated in details. Solution combustion synthesis
has been emerged as low cost, rapid and self sustained process which yields highly
crystalline and homogenous oxide powders with large surface in a single step at lower
temperature than the conventional synthesis method. Dy3+ ions show characteristic
white light emission arising from appropriate yellow to blue (

emission ratio.

Influence of varying sintering temperature and dysprosium ions on BaZrO3


nanophosphors has also taken in account in order to determine the exact optimal
conditions for synthesizing these nanophosphors with superior luminescent properties.
A. EXPERIMENTAL
i) Powder Synthesis
BaZr1-xO3: xDy3+ nanophosphors were synthesized by solution combustion
method using high purity Ba(NO3)2.4H2O, ZrN2O7, Dy(NO3)3.6H2O and urea as
starting reagents. The chemical equation for the reactions is:
Ba(NO3)2 + xDy(NO3)3 + (1-x)ZrN2O7 + ~3.5 CH4N2O BaZr1-xDyxO3 (s) + gaseous
products.
According to nominal composition of BaZr1-xO3: xDy3+ (x =0.01 to 0.05), a
stoichiometric amount of metal nitrates were dissolved in minimum quantity of
deionized water in 200 mL capacity pyrex beaker. Then urea was added in this
solution with molar ratio of urea to oxidizer based on total oxidizing and reducing
valencies of oxidizer and fuel (urea) according to concept used in propellant
chemistry [30]. This aqueous paste containing calculated amount of metal nitrates and
226

urea was then placed in a preheated furnace maintained at 500oC. The mixture of
metal nitrates (oxidizers) and fuel (urea) undergo rapid and self-sustaining
combustion process and the chemical energy released during this exothermic redox
reaction results in

dehydration and foaming followed by decomposition.

Consequently, the large amounts of volatile combustible gases generated alongwith


flames, yields voluminous solid within 5-8 minutes. The powders obtained were again
fired from 800C to 1200C for 3h in order to increase brightness.
ii) Powder Characterization Techniques
The structural characterization of BaZrO3: Dy3+ powders was done by high
resolution X-ray diffraction (XRD) using Rigaku Ultima IV diffractometer with
CuK radiation at 40 kV tube voltage and 40 mA tube current in the 2 range between
10-80o. The morphology of the particles was evaluated using Jeol JSM-6510 scanning
electron microscope (SEM) The excitation and emission spectra of the phosphor were
measured in the ultraviolet-visible region on Hitachi F-7000 fluorescence
spectrophotometer with Xe-lamp at room temperature. The life time calculations of
the phosphor were done by the software of the spectrophotometer (FL solution for F7000).
B. RESULTS AND DISCUSSION
i) X-ray Studies
BaZrO3 possess cubical perovskite structure where divalent 12-oxygen
coordinated Ba cation resides in cuboctahedral site while 6-oxygen coordinated Zr
atoms are present at centrosymmetric locations within the octahedron. XRD profiles
of BaZr0.97Dy0.03O3 nanophosphors as-synthesized and sintered at different
temperatures alongwith the standard JCPDS No. 74-1299 are depicted in Fig.5.5.1a.
All resolved diffraction peaks of BaZr0.97Dy0.03O3 powder sintered at 1200C, well
matched the barium zirconate cubic phase (JCPDS No. 74-1299) having space group
P23 [195]. No additional peaks corresponding to the impurity phases were detected,
indicating complete crystallization in single cubic pervoskite BaZrO3 structure at this
temperature. XRD profiles of BaZr0.97Dy0.03O3 powder sintered at 800C and 1100C
shows the presence of minor phase ZrO2 (JCPDS No. 86-1499) along with diffraction
peaks of main pervoskite phase. Although in as-synthesized BaZr0.97Dy0.03O3 sample,
weak reflex lines of unreacted nitrates were also apparent. However, it is quite
227

noticeable that with the rise in temperature, peaks due to additional phases get
diminished while intensity of main peak (110) enhanced with the decrease in full
width half maximum (FWHM), indicating improvement in doping and crystallinity.
The XRD patterns of BaZr(1-x)DyxO3 powders sintered at 1200C, doped with
different contents of dysprosium ions alongwith the standard JCPDS No. 74-1299 are
depicted in Fig. 5.5.1b. All the samples crystallize in single cubic BaZrO3 phase
having space group P23 [195], confirming no influence of small amount of Dy3+ ions
on crystal structure of this lattice. The results also indicate that despite of ionic radii
difference, Dy3+ (0.091nm) occupies Zr4+ (0.072nm) ion site as compared to larger
Ba2+ (0.135nm) in BaZrO3 host lattice. The crystallite size, D of BaZr(1-x)DyxO3
powders was evaluated by Scherrers formula, D 0.941/ cos, where is the
wavelength of CuK radiation (0.1548 nm), is the full width in radians at halfmaximum (FWHM) and is the Braggs angle of an observed. The calculated average
crystallite sizes, by taking main peak (110) of BaZr0.97Dy0.03O3 powders were found to
be 9.7 nm, 24.0 nm. 29.4 nm and 43.5 nm at 500C, 800C, 1000C and 1200C,
respectively. It can be observed from the calculated results as expected that with the
increase of sintering temperature crystallite size also increases.

228

Fig.5.5.1. XRD patterns of (a) BaZr0.97Dy0.03O3 powders sintered at various


temperatures; (b) BaZr(1-x)DyxO3 (x =1 to 5 mol%) along with standard data of
BaZrO3 (JCPDS no. 74-1299)
ii) Morphological Studies
The SEM images of BaZr0.97Dy0.03O3 powders3 as sintered at temperatures 1000C
and 1200C are represented in Fig. 5.5.2(a-b) respectively.

Fig.5.5.2. SEM images of BaZr0.97Dy0.03O3 sintered at (a) 1000oC; (b) 1200oC.

229

For BaZr0.97Dy0.03O3 sample sintered at 1000oC as shown in Fig. 5.5.2(a)


cubical shaped particles having high agglomeration phenomenon are observed. Some
voids and pores characteristic of combustion synthesized product are also apparent.
With further rise of the sintering temperature upto 1200oC, clear morphology of
tetragonal shape particles could be observed. Hence, a uniform distribution of
tetragonal particles having small size distribution is clearly visible in Fig. 5.5.2(b).
iii) Luminescent Studies
The photoluminescence excitation (PLE) spectrum of BaZr0.97Dy0.03O3
nanophosphors, sintered at 1200oC, monitored with 575 nm (4F9/2 6H13/2) as
emission wavelength presented in Fig. 5.5.3. The PLE spectrum monitored at yellow
emission comprised of several characteristic sharp peaks in 300 to 500 nm range
corresponding to intra-4f transitions Dy3+ in the perovskite host. These excitation
peaks in longer wavelength region at 327 nm, 354 nm, 366 nm, 389 nm, 429 nm,
453nm and 470 nm are assigned to 6H15/26P3/2, 6H15/2 6P7/2, 6H15/2 6P5/2, 6H15/2
4I13/2, 6H15/2 4G11/2, 6H15/2 4I15/2 and 6H15/2 4F9/2 transitions, respectively of
Dy3+ ions in BaZrO3 lattice [11]. In the short wavelength region, weak host or O2-
Dy3+ charge sensitized luminescence indicates very weak Dy3+ ions interactions with
perovskite host.
The photoluminescence (PL) spectra of BaZr0.97Dy0.03O3 nanophosphor assynthesized and sintered at different temperatures on monitoring excitation
wavelength at 354 nm is depicted in Fig. 5.5.4. The two main emission bands in the
470- 490nm (blue region) and 560-580 nm (yellow region) corresponding to
transitions of the Dy3+ ions from well-defined 4F9/2energy states were observed. The
blue emission band centered at 482 nm is assigned to 4F9/26H15/2 while yellow
emission band centered at 575 nm is ascribed to 4F9/2 6H13/2 transitions of Dy3+ ions,
respectively. All samples shows prominent hypersensitive forced electric transition
(yellow emission) in as compared to magnetic dipole transition (blue emission) which
gets hardly influenced by the crystal field symmetry of dysprosium ions. It is quite
noticeable that rise in sintering temperature enhanced the relative PL intensity of assynthesized BaZr0.97Dy0.03O3 nanophosphors with maintained shape and positions of
peaks corresponding to both transitions. This indicates that recombination defects and
surface defects in perovskite lattice are omitted out after sintering of nanophosphors.

230

Fig.5.5.3. Photoluminescence excitation (PLE) spectrum of BaZr0.97Dy0.03O3


nanophosphors, sintered at 1200oC, monitored with em = 575 nm.

Fig.5.5.4. Photoluminescence spectra (PL) of BaZr0.97Dy0.03O3 nanophosphors,


sintered at different temperatures, monitored with em = 575 nm.

231

The PL spectra of BaZr1-xDyxO3 nanophosphors, sintered at 1200oC with


dysprosium doping contents ranging from 1 to 5 mol%, monitored with 354 nm as
excitation wavelength are shown in Fig. 5.5.5. For all BaZr1-xDyxO3 samples, the
hypersensitive yellow (4F9/26H13/2) emission of Dy3+ ions is the prominent one while
relative PL intensities corresponding to both yellow and blue emission enhanced with
the increasing concentration of Dy3+ ions, reaching the maximum at 3 mol% of and
decreased with the further increase of dopant contents. Non-radiative cross-relaxation
mechanism between luminescent centers at higher dopant contents is primarily
responsible for such kind of concentration quenching. It has been noticed that value of
yellow to blue emission ratio (

has not been much influenced by varying

dysprosium contents indicating that hypersensitive electric forced transition


(4F9/26H13/2) senses same crystal field environment at Dy3+ symmetry sites in
BaZrO3 lattice upto 5 mol%.

Fig.5.5.5. Photoluminescence spectra of BaZr1-xDyxO3 nanophosphors with


different Dy3+ contents, sintered at 1200oC and monitored with ex = 354 nm.
232

The luminescence decay curves for BaZr1-xDyxO3 nanophosphors in terms of


varying Dy3+ ions concentrations corresponding to prominent yellow emission at 575
nm, monitored with 354 nm excitation wavelength are displayed in Fig. 5.5.6. In all
samples, 4F9/26H13/2 transitions show single exponential behavior, represented by the
equation I =I0 exp (-t/), where is the radiative decay time, I and I0 are the
luminescence intensities at time t and 0, respectively. The calculated average lifetimes
are 1.64, 1.60, 1.52, 1.44 and 1.35 for 1, 2, 3, 4, 5 mol% of Dy3+ ions, respectively.

Fig.5.5.6. Decay curves of BaZr1-xDyxO3 nanophosphors doped with different


Dy3+contents, sintered at 1200oC and monitored with ex = 354 nm
The Commission International De IEclairage (CIE) color coordinates (x, y)
calculated from the corresponding PL spectra of BaZr1-xDyxO3 nanophosphors, where
x = 0.01 to 0.05 are displayed in Fig. 5.5.7. The Color coordinates for all BaZr1xDyx3O3

samples lie in white region at (0.344, 0.362), (0.333, 0.358), (0.309, 0.332),

(0.325, 0.331) and (0.333, 0.346) corresponding to 1, 2, 3, 4, and 5 mol%,


respectively. Comparable value of color coordinates of these dysprosium doped
nanophosphors to other standard color systems such as NTSC (0.3101, 0.3162),
PAL/SECAM/HDTV (0.3127, 0.329), ProPhoto/Color Match (0.3457, 0.3585) and
CIE white light point (0.33, 0.33), makes BaZr1-xDyxO3 nanophosphor an excellent
candidate for white light emission in LEDs applications.
233

Fig. 5.5.7. CIE color (x, y) coordinates for (a) 1 mol%, (b) 2 mol%, (c) 3 mol%,
and (d) 4 mol%, of Dy3+ ions in BaZr1-xDyxO3 nanophosphors sintered at 1200oC
after excitation at 354 nm
For comparing the luminescence performance of Dy3+ doped nanophosphors
investigated in this chapter, CIE color coordinates calculated from their corresponding
PL spectra with stronger yellow emission attributed to 4F9/26H13/2 transitions of Dy3+
ions were taken into the consideration. Dysprosium doped nanophosphors exhibit
color coordinates in yellowish white region such as BaY1.92Dy0.o8ZnO5 (0.323, 0.399),
La0.90Dy0.10SrAl3O7

(0.332,

0.351),

Gd0.90Dy0.10SrAl3O7

(0.361,

0.436),

Ca0.96Dy0.04ZrO3 (0.348, 0.379) and BaZr0.97Dy0.03O3 (0.309, 0.332) comparable to


various color systems standards. Among these nanophosphors BaZrO3: Dy3+ (3 mol%)
displays excellent white emission having color coordinates quite closer to NTSC
values (0.3101, 0.3162).
CONCLUSION
In brief, BaY2ZnO5: Dy3+, LaSrAl3O7: Dy3+, GdSrAl3O7: Dy3+, CaZrO3: Dy3+
and BaZrO3: Dy3+ nanophosphors have been successfully synthesized by urea assisted
solution combustion method and their structural as well as luminescent characteristics
were studied for various sintering temperature and dysprosium concentration. Highly
234

crystalline single phased BaY2(1-x)DyxZnO5, La1-xDyxSrAl3O7, Gd1-xDyxSrAl3O7,


Ca1-xDyxZrO3 and BaZr1-xDyxO3 nanophosphors were obtained at optimum
temperature 1100C, 550C, 550C, 1200C and 1200C, respectively as revealed by
X-ray diffraction studies. Morphological analysis of these dysprosium doped powders
indicates narrow distribution of highly crystalline particles in nano-regime.
Luminescent features of these nanophosphors were studied in details using
photoluminescence excitation (PLE) and photoluminescence emission (PL) alongwith
luminescence decay curves and color (x, y) coordinates in order to determine the exact
optimal conditions for superior luminescence. Several characteristics sharp excitation
peaks attributed to intra-4f transitions of the Dy3+ ions in 300-500 nm range were
observed in each studied nanophosphors, the peak ascribed to 6H15/2 6P7/2 transition
( 355 nm) being the most prominent. Photoluminescence spectra of each studied
nanophosphors exhibit both blue and yellow emission attributed to 4F9/26H15/2 and
4

F9/26H13/2 transitions, respectively of Dy3+ ions. The relative PL intensity of

hypersensitive electric forced transition (4F9/26H13/2) found to be stronger than the


magnetic dipole transition (4F9/26H15/2) in each investigated dysprosium doped
nanophosphors. The dependence of the emission intensity of BaY2(1-x)Dy2xZnO5,
La1-xDyxSrAl3O7, Gd1-xDyxSrAl3O7, Ca1-xDyxZrO3 and BaZr1-xDyxO3 nanophosphors
on the x value has also been investigated and found to be maximum at 4 mol%,
10 mol%, 10 mol%, 4 mol% and 3 mol%, respectively of dysprosium ions. These
Dy3+ ions doped nanophosphors are potential candidates for white light emission
which may be achieved by proper tuning of yellow to blue emission intensities on
varying the composition of host.
REFERENCES
1. A. Bao, H. Yang, C. Tao, Curr. Appl. Phys., 9 (2009)1252.
2. Sonika, S.D. Han, S.P. Khatkar, M. Kumar, V.B. Taxak, Mater. Sci. Eng. B, 178
(2013) 1436.
3. Z. Zhu, D. Liu, H. Liu, G. Li, J. Du, Z. He, J. Lumin., 132 (2012) 261.
4. Y. Zheng, Y. Huang, M. Yang, N. Guo, H. Qaio, Y. Jia, H. You, J. Lumin., 132
(2012) 362.
5. X. Zhang, Z. Zhang, H.J. Seo, J. Alloys Compds., 509 (2011) 4875.
6. R.G.A. Kumar, S. Hata, K.-I. Ikeda, K.G. Gopchandran, Ceram. Inter., 40 (2014)
2915.
235

7. M.J.J. Lammers, H. Donker, G. Blasse, Mater. Chem. Phys., 1313 (1985) 527.
8. C. Guo, J. Yu, J.H. Jeong, Z. Ren, J. Bai, Phys B, 406 (2011) 916.
9. L. Shi, and H.J. Seo, J. Lumin., 131(2011) 523.
10. Y. Huang, L. Shi, E.S. Kim, H.J. Seo, J. Appl. Phys., 105 (2009) 013512.
11. B. Tian, B. Chen, Y. Tian, J. Sun, X. Li, J. Zhang, H. Zhang, L. Cheng, R. Hua, J.
Chem. Phys. Solids, 73 (2012)1314.
12. C.H. Liang, L.G. Teoh, K.T. Liu, Y.S. Chang, J. Alloys Compds., 517 (2012) 9.
13. C. Guo, X. Ding, Y. Xu, J. Am. Ceram. Soc., 93 (2010)1708.
14. I. Etchart, I. Hernndez, A. Huignard, M. Brard, M. Laroche, W.P. Gillin, R.J.
Curry, A.K. Cheetham, J. Mater. Chem., 21 (2011)1387.
15. H.Y. Chen, R.Y. Yang, S.J.Chang, J. Phys. Chem. Solids, 74 (2013) 344.
16. S. Kunimin, and S. Fujihara, J. Electrochem. Soc., 157 (2010) J175.
17. H.Y. Chen, R.Y. Yang, S.J. Chang, Mater. Lett., 64 (2010) 2548.
18. S. Singh, S.P. Khatkar, M. Kumar, V.B. Taxak, J. Mater. Sci: Mater. Electron., 24,
(2013) 4677.
19. G.K. Cruz, H.C. Basso, M.C. Terrile, R.A. Carvalho, J. Lumin., 86 (2000) 155.
20. C.H. Liang, Y.C. Chang, Y.S Chang, Appl. Phys. Lett., 93 (2008) 211902.
21. H.Y. Chen, M.H. Weng, R.Y. Yang, S.J. Chang, Ceram. Inter., 37 (2011) 1521.
22. Sonika, S.P. Khatkar, M. Kumar, V.B. Taxak, J. Mater. Sci., 49 (2014)572.
23. I. Etchart, M. Brard, M. Laroche, A. Huignard, I. Hernndez, W.P. Gillin, R.J.
Curry, A.K. Cheetham, Chem. Commun., 47 (2011) 6263.
24. J.A. Kaduk, W. Wong-Ng, W. Greenwood, J. Dillingham, B.H. Toby, J. Res. Natl.
Inst. Stand. Technol.,104 (1999)147.
25. G.S.R. Raju, J.Y. Park, H.C. Jung, B.K. Moon, J.H. Jeong, J.H. Kim, Curr. Appl.
Phys. 9 (2009) e92.
26. D. Gao, Y. Li, X. Lai, Y. Wei, J. Bi, Y. Li, M. Liu, Mater. Chem. Phys., 126
(2011)391.
27. S.D. Han, S.P. Khatkar, V.B. Taxak, G. Sharma, D. Kumar, Mater. Sci. Eng. B,
129 (2006)126.
28. C.R. Kesavulu, and C.K. Jayasankar, Mater. Chem. Phys., 130 (2011)1078.
29. S.T. Aruna and A.S. Mukasyan, Curr. Opin. Solid State Mater. Sci. 12 (2008) 44.
30. S. Ekambaram and K.C. Patil, J. Alloys Compds., 448 (1997)7.
31. G. Blasse and B.C. Grabmaier, Luminescent Materials, Springer-Verlag (1994) 33
32. J. Mulak and M. Mulak, J. Phys A: Math Theor., 40 (2007)2063.
236

33. L.A. Diaz-Torres, E.D.L. Rosa, P. Salas, V.H. Romero, A. Angeles-Chavez, J.


Solid State Chem., 181 (2008) 75.
34. X. Zhang and H.J. Seo, J. Alloys Compds., 509 (2011) 2007.
35. G. Sharma, S.P. Lochab, N. Singh, Phys. B., 406 (2011) 2013.
36. N. Dhananjaya, H. Nagabhushana, B.M. Nagabhushana, R.P.S. Chakradhar, C.
Shivakumara, B. Rudraswamy, Phys. B., 405 (2010) 3795.
37. M. Ayvacikli, A. Ege, N. Can, Opt. Mater., 34 (2011)138.
38. Y. Shang, P. Yang, W. Wang, Y. Wang, N. Niu, S. Gai, J. Lin, J. Alloys Compds.,
509 (2011) 837.
39. L.G. Jacobsohn, B.L. Bennett, R.E. Muenchausen, J.F. Smith, D.W. Cooke,
Radiat. Meas., 42 (2007) 675.
40. M.A. Kale, C.P.Joshi, S.V. Moharil, P.L. Muthal, S.M. Dhopte, J. Lumin., 128
(2008) 1225.
41. V. Singh, V.V.R. Kumar, R.P.S. Chakradhar, H.Y. Kwak, Philos. Mag., 90 (2010)
3095.
42. Sheetal, V.B. Taxak, Mandeep, S.P. Khatkar, J. Alloys Compds., 549 (2013) 135140.
43. V. Singh, S. Watanabe, T.K. Gundurao, H.Y. Kwak, J. Fluoresc., 21 (2011) 313.
44. L. Zhou, W.C.H. Choy, J. Shi, M. Gong, H. Liang, T.I. Yuk, J. Solid State Chem.,
178 (2005) 3004.
45. V. Singh, V.K. Rai, K.A. Shamery, J. Nordmann, M. Hasse, J. Lumin., 131 (2011)
2679.
46. S. Kubota, M. Izumi, H. Yamane, M. Shimada, J. Alloys Compds., 283 (1999) 95101.
47. N. Kodama, Y. Tanii, M. Yamaga, J. Lumin., 87 (2000) 1076.
48. X. Zhang, J. Zhang, L. Liang, Q. Su, Mater. Res. Bull., 40 (2005) 281.
49. M. Karbowiak, P. Gnutek, C. Rudowicz, WR Romanowski, Chem. Phys., 387
(2011) 69.
50. W.R. Romanowski, S. Golab, W.A. Pisarski, G.D. Dzik, M. Berkowski, A.
Pajaczkowska, Int. J. Electron., 81 (1996) 457.
51. V.R. Bandi, B.K. Grandhe, K. Jang, H.-S. Lee, D.-S. Shin, S.-S. Yi, J.-H. Jeong, J.
Alloys Compds., 512 (2012) 264.
52. B. Yan and X.-Q. Su, J. Non- Cryst. Solids., 352 (2006) 3275.

237

53. G.S.R. Raju, H.C. Jung, J.Y. Park, C.M. Kanamadi, B.Y. Moon, J.H. Jeong, S.-M.
Son, J.H. Kim, J. Alloys Compds., 481 (2009) 730.
54. B. Liu, D. Ding, Z. Liu, F. Chen, C. Xia, Solid State Ionics, 191 (2011) 68.
55. L. Zhou, W.C.H. Choy, J. Shi, M. Gong, H. Liang, J. Alloys Compds., 463 (2008)
302.
56. J.G. Mahakhode, S.J. Dhoble, C.P. Joshi, S.V. Mohari, J. Alloys Compds., 438
(2007) 293.
57. P.S. Anil Kumar, J.J. Shrotri, S.D. Kulkarni, C.E. Deshpande, S.K. Date, Mater.
Lett., 27 (1996) 293.
58. V. Singh, S. Watanabe, T.K. Gundu Rao, H.Y. Kwak, Solid State Sci., 12 (2010)
1981.
59. S.S. Pitale, V. Kumar, I. Nagpure, O.M. Ntwaeaborwa, H.C. Swart, Curr. Appl.
Phys.,11 (2011) 341.
60. J. Zhang, L. Wang, Y. Jin, X. Zhang, Z. Hao, X.J. Wang, J. Lumin., 131 (2011)
429.
61. F.G. Meng, X.M. Zhang, H.J. Seo, Opt. Laser Technol., 44 (2012) 185.
62. P. Lingling, H. Tao, C. Hui, Z. Tiejun, J. Rare Earths, 31 (2013) 235.
63. V. Singh, S. Watanabe, T.K. Gungu Rao, K. Al-Shamery, M. Haase, J. Lumin.,
132 (2012) 2036.
64. N. Pal, M. Paul, A. Bhaumik, Appl. Catal. A, 393 (2011) 153.
65. R. Parra, R. Savu, L.A. Ramajo, M.A. Ponce, J.A. Varela, M.S. Castro, P.R.
Bueno, E. Joanni, J. Solid State Chem., 183 (2010) 1209.
66. Z. Lu, L. Chen, Y. Tang, Y. Li, J. Alloys Compd., 387 (2005) L1.
67. W.Y. Jia, W.L. Xu, I. Rivera, A. Perez, F. Fernandez, Solid State Commun., 126
(2003) 153.
68. F.H. Norton, Fine Ceramics, Mcgraw-Hill, New York, (1970) 40.
69. T. Yajima, K. Koide, H. Takai, N. Fukatsu, H. Iwahara, Solid State Ionics, 79
(1995) 333.
70. Y. Suzuki, P. E. D. Morgan, T. Ohji, J. Am. Ceram. Soc., 83 (2000) 2091.
71. S.K. Manik and S.K. Pradhan, J. Appl. Crystallogr., 38 (2005) 291.
72. Y. Suzuki, P.E.D. Morgan, T. Ohji, Mater. Sci. Eng. A-Struct., 304306 (2001)
780.
73. A.M. Azad, S. Subramaniam, T.W. Dung, J. Alloys Compd., 334 (2002) 118.
74. M. Pollet, S. Marinel, G. Desgardin, J. Eur. Ceram. Soc., 24 (2004) 119.
238

75. K. Kiyoshi, Y. Shu, I.Yoshiaki, Solid state Ion., 108 (1998) 355.
76. T. Yamagushi, Y. Komatsu, T. Otobe, Y. Murakami, Ferroelectrics, 27 (1980)
273.
77. G. Rog, M. Dudek, A. Kozlowska-Rog, M. Bucko, Electrochim. Acta, 47 (2002)
4523.
78. W. Engelen, A. Buekenhoudt, J. Luyten, A. D. Shutter, Solid State Ion., 96 (1997)
55.
79. H. Iwahara, Y. Asakura, K. Katahria, M. Tanaka, Solid State Ion., 168 (2004) 229.
80. J. Kung, R.J. Angel, N.L. Ross, Phys. Chem. Miner., 28 (2000) 35.
81. Y. Li and Y.Wang, Proceedings of International Meeting on Information Display,
pp. 372-373.
82. W. Li, G. Zhou, A. Zhang, Q. Du, H. Zhou, J. Zhang, J. Chin. Ceram. Soc., 39
(2011) 1729.
83. R. Vassen, X. Cao, F. Tietz, D. Basu, D. Stover, J. Am Ceram. Soc., 3 (200) 2023.
84. A. Erb, E. Walker, R. Flukiger, Phys. C, 245 ((1995) 245.
85. Z. Chen, S. Duncan, K.K. Chawla, M. Koopman, G.M. Janowski, Mater. Charact.,
48 (2002) 305.
86. H.G. Bohn and T. Schober, J. Am Ceram. Soc., 83 (2000)768.
87. M. Viviani, M.T. Buscaglia, V. Buscagalia, M. Leoni, P. Nanni, J. Eur. Ceram.
Soc., 21 (2001) 1981.
88. K. Katahira, Y. Kohchi, T. Shiramura, H. Iwahara, Solid State Ionics, 138 (2001)
91.
89. H. Padma Kumar, C.V. Kumar, C.N. Geirge, S. Soloon, R. Jose, J.K. Thomas, J.
Koshy, J. Alloys Compds., 458 (20008) 528.
90. S. Parida, S.K. Rout, L.S. Cavalcante, E. Sinha, M.S. Li, V. Subramanian, N.
Gupta, V.R. Gupta, J.A. Varela, E. Longo, Ceram. Inter., 38 (2012) 2129.
91. M. Enhessari, S. Khanahmadzadeh, K. Ozaee, J. Iran. Chem. Res.,3 (2011) 11.
92. R. Borja-Urby, L.A. Diaz-Torres, P. Salas, C.Angeles-Chavez, O. Meza, Mater.
Sci. Engg. B, 176 (2011) 1388.

239

You might also like