Photoluminescence properties of Eu3+ ions in yttrium oxide nanoparticles: defect vs. normal sites

I. E. Kolesnikov*ab, A. V. Povolotskiya, D. V. Mamonovaac, E. Lähderantab, A. A. Manshinaa and M. D. Mikhailovc
aSaint Petersburg State University, Universitetskaya nab. 7-9, 199034, Saint Petersburg, Russia. E-mail: ie.kolesnikov@gmail.com
bLappeenranta University of Technology LUT, Skinnarilankatu 34, 53850 Lappeenranta, Finland
cScientific and Technological Institute of Optical Material Science, VNTs S. I. Vavilov State Optical Institute, Babushkina 36-1, 192171, Saint Petersburg, Russia

Received 29th June 2016 , Accepted 8th August 2016

First published on 8th August 2016


Abstract

The position of activator ions in the lattice has a fundamental effect on the luminescent properties of phosphors. In this paper, we describe the normal and defect sites of Eu3+ ions in a Y2O3 crystal lattice, their interaction and difference in physical properties. Analytically detectable amounts of defect sites reached in Y2O3:Eu3+ nanopowders with an average size of 40–50 nm were synthesized by a novel combined Pechini-foaming method. The luminescence properties of Eu3+ ions in defect sites are most pronounced in highly doped nanocrystalline powders (32 and 40 at%). To explain this phenomenon we suggest the mechanism of irreversible energy transfer from normal to defect sites of Eu3+ ions.


1. Introduction

In recent years, materials doped with rare-earth ions (REI) have been widely used as high-performance luminescent devices, magnets, and catalysts due to their unique electronic, optical, and chemical characteristics.1–5 Among the rare-earth compounds, yttrium oxides (Y2O3) have been widely used in a broad range of fields as one of the multifunctional materials. Perfect thermal stability, chemical durability, and low phonon energy are the main advantages of Y2O3 when used as a host matrix for phosphors.6–8 The luminescence properties of the materials largely depend on their size, shape, dimensions, and structure.9 The development of nano- or micromaterials with controlled chemical–physical parameters provides opportunities in exploring new phenomena.10 Therefore, the design and controlled synthesis of inorganic materials with desired sizes and shapes have been attracting attention for many years.11–16 Nowadays, various methods for the synthesis of rare-earth oxide doped phosphors have been developed, including the solution combustion,17 sol–gel,18 spray pyrolysis,19 hydrothermal synthesis,20 and precipitation methods.21

Pechini technique22 (variation of sol–gel method) is an inexpensive and convenient way to prepare REI doped phosphors; but it results in formation of strongly agglomerated nanocrystalline powders.23–25 Optimization of the synthesis process allowed us to modify the procedure and improve properties of the obtained powders. It was found that adding a foaming agent leads to the intense gas release during calcination procedure. This process prevents agglomeration and sintering of the derived particles.

Due to wide application of Y2O3:Eu3+ particles, there are a lot of papers devoted to the study of this material. In the last years manuscripts concerning comparison of different synthesis methods,26 annealing temperature and doping concentration effect27,28 and influence of cold compaction29 were published. However, to the best of our knowledge, there is a lack of papers dealing with wide concentration series of Eu-doped yttria nanoparticles and highly doped samples.

The present paper is focused mainly on the study of structural and luminescence properties of nanocrystalline Y2O3:Eu3+ powders synthesized by the combined Pechini-foaming method which is first reported here. The detailed analysis of REI doping concentration effect on synthesized samples was presented. The most interesting point and one of the novel outcomes of this research is that Eu3+ ions could occupy not only two well-known symmetry sites in cubic Y2O3 host C2 (noncentrosymmetric) and C3i (centrosymmetric) but also defect sites. Photoluminescence properties of ions at various sites were studied at different excitation wavelengths.

2. Experimental

Nanocrystalline Y2O3 doped with Eu3+ ions was prepared by the combined Pechini-foaming method. Previously we synthesized complex oxides doped with rare earth ions using modified Pechini method.30–32 This method reduces agglomeration but requires second calcination procedure in molten salt (potassium chloride).

The proposed combined technique is also an improved standard Pechini method but has the only one calcination stage. The main part of modification is implementation of foaming agent into polymer gel matrix in order to prevent strong agglomeration of synthesized samples. The mixture of aluminium nitrate and potassium chloride was chosen as a foaming agent.

Y(NO3)3 and Eu(NO3)3 prepared by dissolving of Y2O3 and Eu2O3 in nitric acid were used as initial salts for the synthesis process. Then one of foaming agents – aluminium nitrate – was added (molar ratio Al(NO3)3[thin space (1/6-em)]:[thin space (1/6-em)]Y(NO3)3 = 5[thin space (1/6-em)]:[thin space (1/6-em)]3). To the abovementioned solution were added at first saturated solution of citric acid in distilled water (volume ratio 1[thin space (1/6-em)]:[thin space (1/6-em)]1), and then the potassium chloride powder; the resultant mixture was heated up to 150–200 °C. The reaction between citric acid and metal nitrates leads to the chelate complexes formation. The acid excess reacts with the salt, which leads to the potassium dihydrogen citrate formation; oxidized, it forms potassium carbonate. The aforementioned chemical reactions are shown in eqn (1).

 
image file: c6ra16814k-t1.tif(1)

The solution is transformed into polymer gel after the etherification reaction (adding the ethylene glycol in the solution with the ethylene glycol to citric acid solution ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]4). Aluminium, yttrium and europium ions are uniformly distributed in polymer network while polymer cells contain potassium carbonate. The gel was calcinated at 1000 °C for 2 hours to remove organic components. Uniformly distributed aluminium oxide reacts with potassium carbonate, which results in formation of KAlO2 accompanied by intense gas release (eqn (2)).

 
Al2O3 + K2CO3 → 2KAlO2 + CO2↑ (T > 600 °C) (2)

As a result, Y2O3 nanoparticles with homogeneously distributed Eu3+ ions were obtained, reaction byproducts were removed by washing in distilled water. The synthesized particles were collected by centrifugation at 2800 rpm for 5 min.; repeated 3 times. The resulting washed precipitate was dried naturally or in an oven at 110 °C.

Y3+ ions was gradually substituted by Eu3+, to give wide range of Eu doping concentration varying from 2 to 40 at%. The synthesized powder samples (5 mg) and potassium bromide (300 mg) were pressed into pellets (diameter 13 mm) for luminescence studies.

X-ray diffraction patterns were registered with the powder diffractometer UltimaIV (Rigaku) in Bragg–Brentano geometry with CuKα1 radiation (λ = 1.54059 Å) in the 2θ range from 10° to 80°. Phase identification was carried out using a powder diffraction database PowderDiffractionFile (PDF-2, 2011). The unit cell parameters were estimated using TOPAS software by full-profile analysis. Electron micrograph images were obtained with SUPRA 40VP WDS scanning electron microscope (resolution 4 nm). Raman spectra were measured with Bruker SENTERRA Raman Microscope with semiconductor laser 488 nm as an excitation source. Luminescent spectra were recorded with a fluorescence spectrometer Fluorolog-3 equipped with a Xe-arc lamp (450 W power). For the lifetime measurements the europium luminescence was excited with a Xe-flash lamp (150 W power, 3 μs pulse width). All the measurements were performed at the room temperature.

3. Results and discussion

3.1 Structural analysis

Fig. 1a shows the XRD patterns of synthesized concentration series of Y2O3:Eu3+ nanophosphors (doping concentration C(Eu3+) = 2–40 at%). Almost all the peaks in diffraction patterns matched well to cubic phase of Y2O3 (JCPDS 41-1105). Low intensity peak at 2θ ≈ 19° is assigned to the formation of bayerite phase (Al(OH)3) which was not fully removed after synthesis procedure. The broad XRD peaks suggest the formation of nano-sized particles.
image file: c6ra16814k-f1.tif
Fig. 1 (a) XRD patterns of Y2O3:Eu3+ concentration series and the standard card Y2O3 and Al(OH)3; (b) the dependence of the unit cell volume on the Eu3+-doping.

The unit cell parameters were calculated with TOPAS software using full-profile analysis. It was found that both, unit cell parameter and single-crystal cell volume, systematically increase with growth of dopant concentration. This fact can be explained by substitution of yttrium ions (r = 0.89 Å) with larger europium ions (r = 0.95 Å).33 It should be noted that the cell volume has a linear relationship with the content of Eu3+, which is consistent with Vegard's law (Fig. 1b). This result demonstrates that the europium ion has been efficiently and homogeneously incorporated into the host matrix due to the similar ionic radius and chemical reactivity of Eu3+ and Y3+.34–36 So, the obtained results testify that the most of doping Eu3+ ions occupy yttrium sites in crystal lattice.

SEM image of Y2O3:Eu3+ 2 at% phosphor is presented in Fig. 2. As one can see, nanocrystalline powder consists of rather small spherical nanoparticles. Average size of these nanoparticles is about 40–50 nm with standard deviation of σ ≈ 6 nm.


image file: c6ra16814k-f2.tif
Fig. 2 SEM image of Y2O3:Eu3+ 2 at% nanopowder.

Fig. 3 shows Raman spectra of Y2O3:Eu3+ powders with various doping concentration normalized to the most intense band in each of them. These spectra were measured in the spectral region of 80–700 cm−1.


image file: c6ra16814k-f3.tif
Fig. 3 Raman spectra of Y2O3:Eu3+ concentration series.

According to the factor group theory investigation,37 twenty two Raman modes have been predicted for the cubic Y2O3 lattice belonging to the Ia3 space group with C type structure:

 
ΓRS = 4Ag + 4Eg + 14Fg (3)

The major peak assigned as Ag + Fg mode is observed at about 370 cm−1.38 The high intensity of this band compared to the others indicates a large polarizability variation during the vibration.39 The low-intensity line at about 590 cm−1 could be assigned to Fg mode. Lines centered at the spectral region 450–550 cm−1 are quite rarely observed in the Raman spectrum of Eu3+-doped Y2O3 samples. One of these lines has been previously reported in Y2O3:Eu3+ nanoparticles40 and was attributed to the Eu3+ doping effect without any further explanation. It should be noted that these bands do not correspond to any of the first-order Raman-active modes of cubic or monoclinic Y2O3 and Eu2O3 clusters.41–44 In order to determine the origin of these bands Raman spectra were measured using solid state laser 532 nm as an excitation source. It turned out that no line was observed in the spectral range 450–550 cm−1. Therefore, these bands could not be Raman lines and most likely correspond to the luminescence of Eu3+ ions located at positions with C3i symmetry sites. Energy levels of Eu3+ ions that occupy C3i site are higher than that for C2 site. So, the energy transfer to the C2 site occurs, and the transfer efficiency increases with increasing of Eu3+ concentration.45,46 Thus, the proposed mechanism also explains the observed intensity decrease with increasing of doping concentration.

It should be noted that the spectra do not show the expected number of bands. Some bands are quite broad, and some of them are likely overlapped, at least partially. Probably, nanoscale dimension of samples lead to the inhomogeneous strains and to the band broadening.47 Furthermore, some Raman bands can overlap with luminescence lines. As can be seen from Fig. 3, doping concentration also affects the position of Raman lines. Red shift of peaks was observed along with the increase of Eu3+ ions level (inset of Fig. 3). This shift can be explained by changes in the effective mass and effective potential.48,49 Substitution to an element with larger mass and larger electrons number leads to shift toward a lower wavenumber. It was found that 40% substitution of yttrium (89 Da (dalton), 36 electrons) ions for europium (152 Da, 60 electrons) ions results in 13 cm−1 shift of the most intense vibration wavenumber.

3.2 Luminescent properties

Photoluminescence spectrum of Y2O3:Eu3+ 12 at% nanopowder is presented in Fig. 4a. The measurement was performed in the spectral range 500–750 nm upon 265 nm. Observed spectrum consists of characteristic narrow lines corresponding to the intra-configurational f–f transitions. It is well known that the luminescence intensity of the transitions between different j-levels depends on the site symmetry of Eu3+. Eu3+ ions could be located at two symmetry sites in cubic Y2O3 host: C2 (noncentrosymmetric) and C3i (centrosymmetric). In general, electric dipole transitions are forbidden for centrosymmetric site because the energy levels have the same parity. In this case, only weak magnetic dipole transitions occur. Therefore, for the Eu3+ located at C3i site, only the magnetic dipole transitions can be observed. For noncentrosymmetric sites, the crystal-field introduces an opposite-parity part to the potential energy of the crystal levels, which makes the electric dipole transitions possible.50,51 So, for the case of Eu3+ ions at C2 site, the electric dipole transitions are also allowed.46
image file: c6ra16814k-f4.tif
Fig. 4 (a) Emission spectrum of Y2O3:Eu3+ 12 at% nanophosphor; (b) concentration dependence of 5D07F2 intensity.

Emission spectrum is dominated by the forced electric dipole transition 5D07F2 with maximum at 610.2 nm. Maximum of magnetic dipole transition 5D07F1 slightly differs for different symmetry sites. Very weak spectral line centered at 581.2 nm was observed at low-doped samples. This line can be assigned to the Eu3+ ions located at C3i site. Other three lines centered at 586.2, 592 and 598.6 nm correspond to the Eu3+ ions located at C2 site. Transitions 5D07F0 (579.2 nm), 5D07F3 (649.6, 656.6, 661.4 nm) and 5D07F4 (686.4, 692.4, 706, 707.6, 711.2 nm) are also observed in emission spectrum. Moreover, the spectrum contains a low-intensity transition from a higher excited level 5D1: 5D17F1 (532.2 and 537 nm).

Concentration dependence of the integrated intensity of the most strong transition 5D07F2 (610.2 nm) is presented in Fig. 4b. The optimal doping concentration is determined by two competitive effects: on the one hand, an increase of the doping concentration means an increase of luminescence centers number and thus radiative recombination. On the other hand, there is also an increase of the cross-relaxation probability which enhances the efficiency of the nonradiative processes.29 In the Fig. 4b we observe that increase of Eu3+ concentration leads to the increase of the luminescence intensity only up to 12 at%. Further increase of doping concentration results in the intensity reduction. This fact indicates the concentration of quenching effects. It has been reported in earlier studies that there are possibilities of pairing and aggregation of dopant ions at higher doping concentrations where fraction of activators can act as quenchers in yttria lattice.52 It was also shown that optimum doping concentration depends on crystallite size, surface area, synthesis method etc. Dependence of quenching concentration on particle size was studied by Zhang et al.53 It was found that quenching concentration is 6 mol%, 13 mol% and 18 mol% for 3000 nm, 40 nm and 5 nm size particles respectively. In our research optimum concentration was determined to be 12 at% which coincide with optimum concentration for nanophosphors with similar particles dimensions prepared by colloidal precipitation method.54

Excitation spectrum of nanocrystalline phosphor Y2O3:Eu3+ 12 at% for electric–dipole transition 5D07F2 monitored at λem = 610.2 nm is shown in Fig. 5. The spectrum consists of intensive broad band and several sharp weak lines in the longer-wavelength region. The observed broad band can be assigned to the charge transfer (CT) between Eu3+ and O2−, an electron transfers from O2− (2p6) orbital to the empty orbital of 4f6 for Eu3+.55,56 Other weak lines are attributed to the typical intra-configurational f–f transitions of the Eu3+ ion: 7F05L8 (322 nm), 7F05D4 (362.5 nm), 7F05L7 (381.5 nm), 7F05L6 (393.5 nm), 7F05D3 (415 nm), 7F05D2 (465 nm), 7F05D4 (362.5 nm), 7F05L7 (381.5 nm), 7F05L6 (393.5 nm), 7F05D3 (415 nm), 7F05D2 (465 nm), 7F05D1 (532.5 nm) and 7F15D0 (587 nm).


image file: c6ra16814k-f5.tif
Fig. 5 Excitation spectrum of Y2O3:Eu3+ 12 at% nanophosphor (λem = 610.2 nm).

In order to study effect of the excitation wavelength upon the luminescence properties, emission spectra of Y2O3:Eu3+ concentration series were measured. 4 various pumping wavelengths were used: (1) λex = 265 nm (CT); (2) λex = 322 nm (Eu3+, 7F05L8); (3) λex = 393.5 nm (Eu3+, 7F05L6); (4) λex = 465 nm (Eu3+, 7F05D2). Spectra normalized to the intensity of 610.2 nm peak are presented in Fig. 6. Normalized emission spectra do not depend on Eu3+ content upon 265 and 465 nm. However, the increase of Eu3+ doping concentration leads to the significant changes when excited at 322 and 393.5 nm. Shoulder of the forced electric dipole transition 5D07F2 is appeared in the longer-wavelength region. Also, intensity of the very weak line centered at 699 nm rise dramatically. The observed changes could be assigned to the luminescence of Eu3+ ions in a third site completely different to the well-known C2 or C3i sites.


image file: c6ra16814k-f6.tif
Fig. 6 Emission spectra of Y2O3:Eu3+ concentration series upon (a) λex = 265 nm, (b) λex = 322 nm, (c) λex = 465 nm and (d) λex = 393.5 nm.

Excitation spectra of Y2O3:Eu3+ 40 at% nanophosphors for lines centered at 711 and 699 nm (forced electric dipole transition 5D07F4) are presented in Fig. 7. Monitored emission bands are assigned to C2 and unknown site of Eu3+ ions, respectively. Shown excitation spectra were normalized in respect to the intensity of 393.5 nm line. As it can be seen, the positions of the observed excitation bands do not depend on luminescence center site. However, the excitation efficiency of luminescence is different. So, we can draw a conclusion that Eu3+ ions situated at unknown sites are mostly excited through Eu3+ ions located at C2 and C3i sites.


image file: c6ra16814k-f7.tif
Fig. 7 Excitation spectra of Eu3+ ions situated at different sites in Y2O3:Eu3+ 40 at% nanophosphor (λem1 = 699 nm; λem2 = 711 nm).

3.3 Kinetics of luminescence

In addition to the steady-state measurements, important information about luminescence properties of synthesized nanophosphors could be obtained from fluorescence kinetics. Fig. 8 shows emission spectra of Y2O3:Eu3+ 2, 12 and 40 at% nanopowders and wavelength of spectral lines where lifetimes were measured. All these lines are assigned to the transitions from metastable excited level 5D0 to the lower levels.
image file: c6ra16814k-f8.tif
Fig. 8 Emission spectra and position of spectral lines for lifetime measurements of Eu3+-doped Y2O3 phosphors (λex = 393.5 nm).

Lifetimes monitored at different spectral lines are listed in the Table 1. Analysing measured lifetimes in Y2O3:Eu3+ 2 at% and Y2O3:Eu3+ 12 at% we can see that there are two very different observed lifetimes of 5D0 level for the same sample. It is well known that the occupation of different symmetry sites leads to the distinction of lifetimes. So, the presence of two different lifetimes could be explained by various positions of Eu3+ ions in crystal lattice. Measured lifetimes are similar for all emission lines except 699 nm. This suggests that the luminescence bands are attributed to Eu3+ ions occupying C2 sites in crystal lattice. 699 nm emission peak has different lifetime and hence it can be assigned to the Eu3+ ions situated at defect sites.

Table 1 Lifetime of 5D0 level measured at different transitions for Eu3+-doped Y2O3 phosphors
Wavelength (nm) Lifetime (ms)
2 at% 12 at% 40 at%
592 1.29 ± 0.03 0.90 ± 0.02 0.22 ± 0.01
610 1.34 ± 0.04 0.91 ± 0.02 0.15 ± 0.01
630 1.26 ± 0.03 0.99 ± 0.02 0.14 ± 0.01
699 0.36 ± 0.01 0.32 ± 0.01 0.23 ± 0.01
708 1.22 ± 0.03 0.90 ± 0.02


Photoluminescence decays of Y2O3:Eu3+ 2, 12 and 40 at% nanopowders are presented in Fig. 9a. Doping concentration effect on the 5D0 level lifetime of Eu3+ ions occupying both normal and defect positions in crystal lattice was studied to understand further the luminescence properties (Fig. 9b). Emission was excited using 393.5 nm pumping in both cases. Fluorescence decays of Eu3+ ions at C2 sites were measured for the electric dipole transition 5D07F2 (610.2 nm). Experimental curves of Y2O3 samples doped up to 16 at% were fitted by single exponential function. However, higher doped samples show non mono-exponential decay. Correct fits required a double-exponential model to be used. The average lifetime was calculated by a double exponential fit.57

 
image file: c6ra16814k-t2.tif(4)
where A1 and A2 are pre-exponential factors; τ1 and τ2 are lifetimes. The change of fitting model can be explained by the following reason. An increasing concentration of the dopant forces allocation of Eu3+ ions in surface sites. So the non-exponential decay can occur due to the different decays of the Eu3+ ions at the surface and Eu3+ ions in the volume of the nanoparticles.58–60


image file: c6ra16814k-f9.tif
Fig. 9 (a) Experimental data and single exponential fit of Eu3+-doped Y2O3 phosphors (λex = 393.5 nm); (b) 5D0 level lifetime of Eu3+ ions occupied different positions in crystal lattice as a function of Eu3+ doping level.

As it can be seen from Fig. 9b, concentration dependence consists of two parts. Eu3+ concentration increase up to 24 at% leads to a significant decrease of 5D0 lifetime. So, we observed the drop from 1.31 ms to 0.23 ms. Further increase of Eu3+ concentration results in much more smooth lifetime decrease.

Compared to the lifetime of Eu3+ at C2 site, there are only small changes for the lifetime of Eu3+ at defect sites. When doping level reaches 24 at%, 5D0 lifetime of Eu3+ ions at both positions becomes almost the same.

To understand this fact, relative positions of europium ions should be considered. At low doping concentrations the distance between the Eu3+ is very large. Therefore, there is almost no interaction between the Eu3+ ions occupied different (normal and defect) sites. However, the distances between the ions become shorter with the increase of doping level and the interaction between them becomes stronger.61 So, the probability of energy transfer between different sites increases.

3.4 Energy transfer processes of Eu3+ ions

Excitation spectra of luminescence lines centered at 610.2 nm, 699 nm and 711 nm were measured in order to study the interaction between Eu3+ ions situated at normal and defect positions. For all the above-mentioned luminescence lines the ratio of integrated intensities of 7F05L6 (393.5 nm) and 7F05D2 (465 nm) excitation transitions was calculated as a function of Eu3+ doping concentration (Fig. 10a). It was found that in case of 610.2 nm and 711 nm emission this ratio remains constant for all concentration range, whereas in case of 699 nm emission there are two different parts of the studied dependence. In the first part, when doping concentration does not exceed 16 at% the intensity ratio also remains constant. It should be noted that calculated intensity ratio is about 4 times higher than for other emission wavelengths. In the second part, gradual increase of intensity ratio with increase of Eu3+ content was observed. It can be seen that there is linear dependence between the studied ratio and the doping level for high Eu3+ concentration. Growth of the intensity ratio can be explained as follows. As mentioned above, the increase of doping concentration leads to the decrease of distances between Eu3+ ions. Hence, the probability of energy transfer between Eu3+ ions which occupied normal and defect sites in crystal lattice is enhanced. This fact is also confirmed by kinetics measurements (for Y2O3:Eu 24 at% sample lifetimes of Eu3+ ions, which situated at different positions in crystal lattice, become almost the same). The Eu3+ energy levels depend on occupied site due to the crystal field variation. Scheme of energy levels is presented in Fig. 10b. Based on emission spectra (Fig. 6) we can conclude that energy levels of Eu3+ ions occupied defects sites shift towards a longer-wavelength region compared to the Eu3+ ions occupied normal sites. It is well known that resonance energy transfer is more probable than the phonon-assisted one. Due to the fast thermalization between sublevels and level positions, energy transfer from normal to defect sites can be resonant whereas back transfer is always phonon-assisted. Thus, the probability of energy transfer from normal to defect sites is much higher than the probability of back transfer. The Eu3+ 5L8 and 5L6 energy levels reached by 322 and 393.5 nm excitation, respectively, are quite broad and probably these levels overlaps or are close to the defect Eu3+ excited state levels and hence the energy transfer can occur. The Eu3+ 5D2 energy level reached by 465 nm excitation is narrow and therefore there could be no overlap with the Eu3+ defect 5D2 levels. Energy transfer requires a multiphonon process and is unlikely. Increase of the Eu3+ concentration will increase energy transfer from Eu3+ in the C3i and C2 sites to the Eu3+ defect site, and the studied intensity ratio in Fig. 10a is expected to increase.
image file: c6ra16814k-f10.tif
Fig. 10 (a) Ratio of 7F05L6 and 7F05D2 intensities as a function of Eu3+ doping level; (b) scheme of possible energy transfer.

To gain some insight of the luminescence properties of Eu3+ ions in Y2O3 nanophosphors, 4f–4f intensity theory was applied to determinate the radiative and nonradiative transition rates and quantum efficiencies. These parameters were obtained from emission spectra using the technique first reported by Kodaira et al.62 Calculations have been performed according to the procedure described in articles.63,64 Briefly, the Einstein coefficients A0−λ for spontaneous emission can be obtained from the relationship:

 
image file: c6ra16814k-t3.tif(5)
where ν0−λ and I0−λ are, respectively, the frequency and intensity of the corresponding transition 5D07Fλ in the emission spectrum. Due to the magnetic character of the 5D07F1 transition and its weak dependence on crystal field effect, the value of the A0−1 coefficient is taken as a constant and is equal to 50 s−1. The radiative transition rate (Arad) is the sum of all the A0−λ values. The observed lifetime of excited state, radiative and nonradiative (Anrad) transition rates are related through the following equation:
 
image file: c6ra16814k-t4.tif(6)

The quantum efficiency of the 5D0 level may be calculated from:

 
image file: c6ra16814k-t5.tif(7)

We studied effect of excitation wavelength on the calculated values. The obtained results are listed in the Table 2. It should be noted that uncertainties in calculated values do not exceed 10%.

Table 2 Radiative (Arad), nonradiative (Anrad) and total (Atotal) transition rates of the 5D0 level, and quantum efficiencies (η) obtained upon different excitations as a function of Eu3+ doping concentration in Y2O3:Eu3+
C(Eu3+) (at%) Atotal (s−1) λex (nm)
Arad (s−1) Anrad (s−1) η (%)
265 322 393.5 465 265 322 393.5 465 265 322 393.5 465
2 760 380 350 370 400 380 410 400 360 50 46 48 52
4 760 340 350 330 390 420 410 430 370 45 46 43 51
8 960 420 410 390 420 540 550 570 550 44 42 41 43
12 1100 420 410 380 420 680 690 720 680 38 37 35 38
16 1500 420 410 380 420 1100 1100 1100 1100 28 28 25 28
24 4400 420 390 360 410 3900 4000 4000 3900 10 9 8 9
32 5000 400 340 320 390 4600 4700 4700 4600 8 7 6 8
40 6700 410 320 310 370 6300 6300 6400 6300 6 5 5 6


Analyzing the calculated radiative transition rate we can see that it slightly changed with the increase of doping level. Slight variations of this probability could be explained by the crystal field changes. In spite of similar values of the ionic radii, substitution of Y3+ ions (r = 0.89 Å) to Eu3+ ions (r = 0.95 Å)18 affects the position of Eu3+ energy levels and the oscillator strength. It should be noted that the calculated radiative probabilities were almost the same for all excitation wavelengths. Special attention should be paid to the highly-doped samples (32 and 40 at%). Their radiative transition rates are smaller when excited at 322 and 393.5 nm, compared with 265 and 465 nm. This can be explained by significant emission of Eu3+ ions at defect sites upon 322 and 393.5 nm (Fig. 6). Due to the partial overlap of emission lines attributed to the electric dipole transitions of normal and defect ions, we obtained averaged radiative transition rate of Eu3+ ions occupied both normal and defect sites in crystal lattice.

Nonradiative transition rate is monotonically raised together with doping level. However, the dramatic increase (∼3000 s−1) was observed for 24 at% doped sample. It should be noted that from this doping level the decay curve became non exponential. Therefore, we can draw the conclusion that decay curve shape is probably changed due to the nonradiative processes. There are different types of these processes: multiphonon relaxation, quenching on impurities remaining after synthesis (e.g. OH group), migration and cross-relaxation. Detailed description of these processes was given in our previous paper.65,66 Taking into account low energy of phonons in Y2O3 host and large energy gap between the excited level 5D0 and the nearest lower level it can be concluded that multiphonon relaxation is impossible. Probability of quenching on impurities cannot be neglected due to the presence of a few OH groups that can be adsorbed from the atmosphere. Cooperative processes could lead to the diffusion process between europium ions when the involved levels are identical (migration) or to self-quenching when they are different (cross-relaxation).67 In case of Eu3+ ions the 5D0 level is not involved in the self-quenching processes while 5D1 is. However, both 5D0 and 5D1 levels are involved in diffusion processes.

Only cooperative processes are strongly dependent on the doping level among all considered nonradiative processes. So, increase of nonradiative transition rate is most likely connected with the growth of energy transfer efficiency to impurities.

As can be seen, quantum efficiencies do not depend on the excitation wavelength. The same values of radiative transition rate and growth of nonradiative transition rate results in decrease of quantum efficiencies with the increase of doping level. It was found that η decreased from 50% for 2 at% to 6% for 40 at%, almost nine times.

4. Conclusions

Nanocrystalline Eu3+-doped Y2O3 materials were synthesized by novel combined Pechini-foaming technique. Structural analysis indicated the formation of pure cubic phase nanocrystals without any impurity. The obtained samples consist of nanoparticles with average size about 40–50 nm. Vibrational spectrum was dominated by phonon centered at about 370 cm−1. It was found that the increasing Eu3+ doping concentration leads to the red shift of Raman lines. Intense emission of the nanophosphors due to doping with the Eu3+ ions are assigned to transitions occurring between the 5D0 excited state and 7FJ ground states of these ions. Most efficiently emission was activated by charge transfer between Eu3+ and O2− and subsequent energy transfer to the dopant ions. Optimal doping concentration was determined to be 12 at% in terms of luminescence intensity. Steady-state and kinetic measurements confirm existence of normal and defect sites in Y2O3 crystal lattice for europium ions. The excitation mechanism of Eu3+ ions occupying defect sites was proposed. It was found that energy transfer from normal to defect sites increases with growth of doping level due to the decrease of distance between Eu3+ ions. Radiative and nonradiative transition rates were calculated using the model of f–f transition intensities upon different excitation wavelengths. It was found that radiative transition rate slightly depends on excitation wavelength and doping level. Averaged radiative transition rate of Eu3+ ions occupying both normal and defect sites in crystal lattice was obtained for high doping concentration upon 322 and 393.5 nm excitations. Concentration quenching is caused by the growth of energy transfer efficiency to impurities.

Acknowledgements

This research has been supported by The Ministry of Education and Science of the Russian Federation (# 14.604.21.0078, RFMEFI60414X0078). Experimental investigations were carried out in the “Center for Optical and Laser materials research”, “Research Centre for X-ray Diffraction Studies”, and “Interdisciplinary Resource Center for Nanotechnology” (Saint Petersburg State University).

References

  1. A.-W. Xu, Y. Gao and H.-Q. Liu, J. Catal., 2002, 207, 151–157 CrossRef CAS.
  2. Z. Lin, X. Liang, Y. Ou, C. Fan, S. Yuan, H. Zeng and G. Chen, J. Alloys Compd., 2010, 496, L33–L37 CrossRef CAS.
  3. E. Nogales, B. Méndez, J. Piqueras and J. A. García, Nanotechnology, 2009, 20, 115201 CrossRef CAS PubMed.
  4. G. Wang, Q. Peng and Y. Li, Acc. Chem. Res., 2011, 44, 322–332 CrossRef CAS PubMed.
  5. W. Qin, D. Zhang, D. Zhao, L. Wang and K. Zheng, Chem. Commun., 2010, 46, 2304–2306 RSC.
  6. T. S. Atabaev, O. S. Jin, J. H. Lee, D.-W. Han, H. H. T. Vu, Y.-H. Hwang and H.-K. Kim, RSC Adv., 2012, 2, 9495–9501 RSC.
  7. J. Silver, M. I. Martinez-Rubio, T. G. Ireland, G. R. Fern and R. Withnall, J. Phys. Chem. B, 2001, 105, 948–953 CrossRef CAS.
  8. G. Glaspell, J. Anderson, J. R. Wilkins and M. S. El-Shall, J. Phys. Chem. C, 2008, 112, 11527–11531 CAS.
  9. G. A. Sotiriou, M. Schneider and S. E. Pratsinis, J. Phys. Chem. C, 2012, 116, 4493–4499 CAS.
  10. M. J. MacLachlan, I. Manners and G. A. Ozin, Adv. Mater., 2000, 12, 675–681 CrossRef CAS.
  11. D. Yang, X. Kang, M. Shang, G. Li, C. Li and J. Lin, Nanoscale, 2011, 3, 2589–2595 RSC.
  12. S. Huang, X. Zhang, L. Wang, L. Bai, J. Xu, C. Li and P. Yang, Dalton Trans., 2012, 41, 5634–5642 RSC.
  13. W. Xu, Y. Wang, X. Bai, B. Dong, Q. Liu, J. Chen and H. Song, J. Phys. Chem. C, 2010, 114, 14018–14024 CAS.
  14. C. Li, Z. Quan, P. Yang, S. Huang, H. Lian and J. Lin, J. Phys. Chem. C, 2008, 112, 13395–13404 CAS.
  15. G. Li, C. Peng, C. Li, P. Yang, Z. Hou, Y. Fan, Z. Cheng and J. Lin, Inorg. Chem., 2010, 49, 1449–1457 CrossRef CAS PubMed.
  16. S. Zeng, G. Ren, C. Xu and Q. Yang, CrystEngComm, 2011, 13, 1384–1390 RSC.
  17. W.-W. W.-P. Zhang, X. Mei, W.-W. W.-P. Zhang, Y. Min, Q. Ze-Ming, X. Shang-Da, C. Garapon, Z. Wei-Wei, X. Mei, Z. Wei-Ping, Y. Min, Q. Ze-Ming, X. Shang-Da and C. Garapon, Chem. Phys. Lett., 2003, 376, 318–323 CrossRef CAS.
  18. M. K. Chong, K. Pita and C. H. Kam, J. Phys. Chem. Solids, 2005, 66, 213–217 CrossRef CAS.
  19. K. Y. Jung and K. H. Han, Electrochem. Solid-State Lett., 2005, 8, H17–H20 CrossRef CAS.
  20. P. K. Sharma, M. H. Jilavi, R. Nar and H. Schmidt, J. Mater. Sci. Lett., 1998, 17, 823–825 CrossRef CAS.
  21. Y. He, Y. Tian and Y. Zhu, Chem. Lett., 2003, 32, 862–863 CrossRef CAS.
  22. M. P. Pechini, US Pat., 3,330,697, 1967.
  23. T. Zaki, K. I. Kabel and H. Hassan, Ceram. Int., 2012, 38, 4861–4866 CrossRef CAS.
  24. X. Tao, X. Chen, Y. Xia, H. Huang, Y. Gan, R. Wu, F. Chen and W. Zhang, J. Mater. Chem. A, 2013, 1, 3295–3301 CAS.
  25. A. Z. Simões, A. Ries, C. S. Riccardi, A. H. Gonzalez, M. A. Zaghete, B. D. Stojanovic, M. Cilense and J. A. Varela, Mater. Lett., 2004, 58, 2537–2540 CrossRef.
  26. S. Som, S. Das, S. Dutta, H. G. Visser, M. K. Pandey, P. Kumar, R. K. Dubey and S. K. Sharma, RSC Adv., 2015, 5, 70887–708898 RSC.
  27. A. M. Khachatourian, F. Golestani-Fard, H. Sarpoolaky, C. Vogt, E. Vasileva, M. Mensi, S. Popov and M. S. Toprak, J. Lumin., 2016, 169, 1–8 CrossRef CAS.
  28. S. Nigam, C. S. Kamal, K. R. Rao, V. Sudarsan and R. K. Vatsa, J. Lumin., 2016, 178, 219–225 CrossRef CAS.
  29. R. Kubrin, J. J. do Rosário and G. A. Schneider, RSC Adv., 2015, 5, 25555–25564 RSC.
  30. I. E. Kolesnikov, A. V. Povolotskiy, D. V. Tolstikova, A. A. Manshina and M. D. Mikhailov, J. Phys. D: Appl. Phys., 2015, 48, 075401 CrossRef.
  31. I. E. Kolesnikov, D. V. Tolstikova, A. V. Kurochkin, N. V. Platonova, S. A. Pulkin, A. A. Manshina and M. D. Mikhailov, Mater. Res. Bull., 2015, 70, 799–803 CrossRef CAS.
  32. I. E. Kolesnikov, E. V. Golyeva, A. V. Kurochkin and M. D. Mikhailov, J. Alloys Compd., 2016, 654, 32–38 CrossRef CAS.
  33. D. Tu, Y. Liu, H. Zhu, R. Li, L. Liu and X. Chen, Angew. Chem., Int. Ed., 2013, 52, 1128–1133 CrossRef CAS PubMed.
  34. A. Huignard, T. Gacoin and J.-P. Boilot, Chem. Mater., 2000, 12, 1090–1094 CrossRef CAS.
  35. Y. Jun, Y. Jung and J. Cheon, J. Am. Chem. Soc., 2002, 124, 615–619 CrossRef CAS PubMed.
  36. L. Zhu, Y. Liu, X. Fan, D. Yang and X. Cao, J. Lumin., 2011, 131, 1380–1385 CrossRef CAS.
  37. W. B. White and V. G. Keramidas, Spectrochim. Acta, Part A, 1972, 28, 501–509 CrossRef CAS.
  38. Y. Repelin, C. Proust, E. Husson and J. M. Beny, J. Solid State Chem., 1995, 118, 163–169 CrossRef CAS.
  39. A. Ubaldini and M. M. Carnasciali, J. Alloys Compd., 2008, 454, 374–378 CrossRef CAS.
  40. K. Zhang, A. K. Pradhan, G. B. Loutts, U. N. Roy, Y. Cui and A. Burger, J. Opt. Soc. Am. B, 2004, 21, 1804–1808 CrossRef CAS.
  41. S. Ray, S. F. León-Luis, F. J. Manjón, M. A. Mollar, Ó. Gomis, U. R. Rodríguez-Mendoza, S. Agouram, A. Muñoz and V. Lavín, Curr. Appl. Phys., 2014, 14, 72–81 CrossRef.
  42. J. Lancok, C. Garapon, C. Martinet, J. Mugnier and R. Brenier, Appl. Phys. A, 2004, 79, 1263–1265 CrossRef CAS.
  43. A. Kremenovic, J. Blanusa, B. Antic, P. Colomban, V. Kahlenberg, C. Jovalekic and J. Dukic, Nanotechnology, 2007, 18, 145616 CrossRef.
  44. N. Dilawar, S. Mehrotra, D. Varandani, B. V. Kumaraswamy, S. K. Haldar and A. K. Bandyopadhyay, Mater. Charact., 2008, 59, 462–467 CrossRef CAS.
  45. M. Buijs, A. Meyerink and G. Blasse, J. Lumin., 1987, 37, 9–20 CrossRef CAS.
  46. C. Hang, Z. Pei-Fen, Z. Hong-Yang, L. Hong-Dong and C. Qi-Liang, Chin. Phys. B, 2014, 23, 57801 CrossRef.
  47. J. E. Spanier, R. D. Robinson, F. Zhang, S.-W. Chan and I. P. Herman, Phys. Rev. B: Condens. Matter Mater. Phys., 2001, 64, 245407 CrossRef.
  48. M. T. Dove, Introduction to lattice dynamics, Cambridge university press, 1993, vol. 4 Search PubMed.
  49. M. Born and K. Huang, Dynamical Theory of Crystal Lattices, Clarendon press, oxford, 1954 Search PubMed.
  50. B. R. Judd, Phys. Rev., 1962, 127, 750 CrossRef CAS.
  51. M. Buijs, A. Meyerink and G. Blasse, J. Lumin., 1987, 37, 9–20 CrossRef CAS.
  52. X. Qin, Y. Ju, S. Bernhard and N. Yao, J. Mater. Res., 2005, 20, 2960–2968 CrossRef CAS.
  53. W.-W. W.-P. W. P. Zhang, W.-W. W.-P. W. P. Zhang, P.-B. B. Xie, M. Yin, H.-T. T. Chen, L. Jing, Y.-S. S. Zhang, L.-R. R. Lou and S.-D. Da Xia, J. Colloid Interface Sci., 2003, 262, 588–593 CrossRef CAS PubMed.
  54. G. S. Gowd, M. K. Patra, S. Songara, A. Shukla, M. Mathew, S. R. Vadera and N. Kumar, J. Lumin., 2012, 132, 2023–2029 CrossRef.
  55. E. V. Golyeva, D. V. Tolstikova, I. E. Kolesnikov and M. D. Mikhailov, J. Rare Earths, 2015, 33, 129–134 CrossRef CAS.
  56. B. Liu, M. Gu, X. Liu, C. Ni, D. Wang, L. Xiao and R. Zhang, J. Alloys Compd., 2007, 440, 341–345 CrossRef CAS.
  57. T. Fujii, K. Kodaira, O. Kawauchi, N. Tanaka, H. Yamashita and M. Anpo, J. Phys. Chem. B, 1997, 101, 10631–10637 CrossRef CAS.
  58. Y. Zheng, H. You, G. Jia, K. Liu, Y. Song, M. Yang and H. Zhang, Cryst. Growth Des., 2009, 9, 5101–5107 CAS.
  59. V. Sudarsan, F. C. J. M. van Veggel, R. A. Herring and M. Raudsepp, J. Mater. Chem., 2005, 15, 1332 RSC.
  60. N. S. Singh, R. S. Ningthoujam, M. N. Luwang, S. D. Singh and R. K. Vatsa, Chem. Phys. Lett., 2009, 480, 237–242 CrossRef CAS.
  61. D. L. Dexter, J. Chem. Phys., 1953, 21, 836–850 CrossRef CAS.
  62. C. A. Kodaira, H. F. Brito, O. L. Malta and O. A. Serra, J. Lumin., 2003, 101, 11–21 CrossRef CAS.
  63. D. Hreniak, W. Strek, J. Amami, Y. Guyot, G. Boulon, C. Goutaudier and R. Pazik, J. Alloys Compd., 2004, 380, 348–351 CrossRef CAS.
  64. R. J. Wiglusz, T. Grzyb, S. Lis and W. Strek, J. Lumin., 2010, 130, 434–441 CrossRef CAS.
  65. I. E. Kolesnikov, D. V. Tolstikova, A. V. Kurochkin, A. A. Manshina and M. D. Mikhailov, Opt. Mater., 2014, 37, 306–310 CrossRef CAS.
  66. I. E. Kolesnikov, D. V. Tolstikova, A. V. Kurochkin, S. A. Pulkin, A. A. Manshina and M. D. Mikhailov, J. Lumin., 2015, 158, 469–474 CrossRef CAS.
  67. F. Auzel, in Radiationless processes, Springer, 1980, pp. 213–286 Search PubMed.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.