Luminescence properties of Eu 3+ doped CaMoO4 nanoparticle

7
Dalton Transactions Dynamic Article Links Cite this: Dalton Trans., 2011, 40, 7595 www.rsc.org/dalton PAPER Luminescence properties of Eu 3+ doped CaMoO 4 nanoparticles Abdul Kareem Parchur, a Raghumani Singh Ningthoujam,* b Shyam Bahadur Rai, a Gunadhor Singh Okram, c Ram Asaray Singh, d Mohit Tyagi, e S. C. Gadkari, e Raghvendra Tewari f and Rajesh Kumar Vatsa b Received 25th Feb 2011, Accepted 17th May 2011 DOI: 10.1039/c1dt10878f When Eu 3+ ions occupy Ca 2+ sites of CaMoO 4 , which has a body centered tetragonal structure with inversion symmetry, only the magnetic dipole transition ( 5 D 0 7 F 1 ) should be allowed according to Judd–Ofelt theory. Even if there are a few distortions in the Eu 3+ environment, its intensity should be more than that of the electric dipole transition ( 5 D 0 7 F 2 ). We report here the opposite effect experimentally and ascribe this to the polarizability effect of the MoO 4 tetrahedron, which is neighboring to EuO 8 (symmetric environment). The contribution of the energy transfer process from the Mo–O charge transfer band to Eu 3+ and the role of Eu 3+ over the surface of the particle could be distinguished when luminescence decay processes were measured at two different excitations (250 and 398 nm). Further, the luminescence intensities and lifetimes increase significantly with increasing heat-treatment temperature of the doped samples. This is attributed to the reduction of H 2 O from the surface of the particles and a non-radiative process after heat treatment. 1. Introduction Rare-earth ion doped calcium molybdate (CaMoO 4 ) has attracted much interest due to its wide luminescence applications. 1–6 These include optical filters, solid state lasers, LED (light emitting diodes), scintillators, microwave dielectrics, cryogenic detectors and fluorescent lamps. 7,8 Moreover, some of its attractive features include a high melting point (1445–1480 C), refractive index (1.98), effective average decay time (14 ms), photo electron yield (9%) and it is also chemically resistant and non-hygroscopic. 9 CaMoO 4 shows a broad blue and/or green luminescence emission peak in the range 350–650 nm with a peak maximum around 500 nm. The peak maximum however varies with particle size and oxygen vacancies in the lattice. 10 It is related to the electron-hole recombination after excitation through the band gap of the charge transfer (CT) from O to Mo (2p 4d orbital) in the MoO 4 2- complex tetrahedron cluster itself. 11 Liu et al. 12 calculated the band gap of pure CaMoO 4 crystal using a density functional theory calculation as 3.6 eV. Nearly consistently, Ryu et al. 10 found it to be 3.41 eV using a linearized-augmented-plane-wave technique. The experimentally measured value was slightly higher. 10 a Department of Physics, Banaras Hindu University, Varanasi, 221005, India b Chemistry Division, Bhabha Atomic Research Centre, Mumbai, 400085, In- dia. E-mail: [email protected]; Fax: +91-22-25505151; Tel: +91-22-2559321 c UGC–DAE Consortium for Scientific Research, Indore, 452001, India d Department of Physics, Dr Hari Singh Gour University, Sagar, 470003, India e Technical Physics Division, Bhabha Atomic Research Centre, Mumbai, 400085, India f Material Science Division, Bhabha Atomic Research Centre, Mumbai, 400085, India In fact, when a rare-earth dopant like Eu 3+ is doped in calcium molybdate, what is predicted to happen to the luminescence? In our study, the crystal structure of CaMoO 4 has inversion symmetry. 13 Eu 3+ ions occupy Ca 2+ sites and thus only magnetic dipole transi- tion should be allowed because Eu 3+ ions have inversion symmetry according to Judd–Ofelt theory. 14,15 In many reported papers, it is suggested that Eu 3+ ions have an asymmetric environment. 11,16,17 This has been claimed because the electric dipole transition intensity is higher than the magnetic dipole transition of Eu 3+ in the observed spectra. However, the relationship between crystal structure and luminescence behavior has not been reported. The unusual behaviour in this compound has not been discussed in earlier literature so far, to the best of the authors’ knowledge. This is explained based on the polarizability effect of the MoO 4 tetrahedron, which is sharing with EuO 8 /CaO 8 through an O atom in this work. The effect of particle size on luminescence behavior and lifetime are also studied. 2. Experimental details The nanoparticles of CaMoO 4 : Eu 3+ (Eu 3+ = 0, 1, 3, 5, 7, 10 at. %) were prepared at low temperature (150 C for 3 h) using urea hydrolysis in ethylene glycol. 13 Finally, the as-prepared sample was annealed at 500 and 900 C for 4 h each. The photoluminescence (PL) spectra of these powder phosphors were recorded using Hitachi F-4500 spectrometer with a 150 W Xe lamp as a source at the spectral resolution of 3 nm. All the measurements were carried out at room temperature. PL decay was recorded with an Edinburgh instrument F920 equipped with 100 W ms flash xenon lamp as the excitation source. This journal is © The Royal Society of Chemistry 2011 Dalton Trans., 2011, 40, 7595–7601 | 7595 Downloaded by Bhabha Atomic Research Centre on 27 July 2011 Published on 22 June 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10878F View Online

Transcript of Luminescence properties of Eu 3+ doped CaMoO4 nanoparticle

DaltonTransactions

Dynamic Article Links

Cite this: Dalton Trans., 2011, 40, 7595

www.rsc.org/dalton PAPER

Luminescence properties of Eu3+ doped CaMoO4 nanoparticles

Abdul Kareem Parchur,a Raghumani Singh Ningthoujam,*b Shyam Bahadur Rai,a Gunadhor Singh Okram,c

Ram Asaray Singh,d Mohit Tyagi,e S. C. Gadkari,e Raghvendra Tewari f and Rajesh Kumar Vatsab

Received 25th Feb 2011, Accepted 17th May 2011DOI: 10.1039/c1dt10878f

When Eu3+ ions occupy Ca2+ sites of CaMoO4, which has a body centered tetragonal structure withinversion symmetry, only the magnetic dipole transition (5D0→7F1) should be allowed according toJudd–Ofelt theory. Even if there are a few distortions in the Eu3+ environment, its intensity should bemore than that of the electric dipole transition (5D0→7F2). We report here the opposite effectexperimentally and ascribe this to the polarizability effect of the MoO4 tetrahedron, which isneighboring to EuO8 (symmetric environment). The contribution of the energy transfer process fromthe Mo–O charge transfer band to Eu3+ and the role of Eu3+ over the surface of the particle could bedistinguished when luminescence decay processes were measured at two different excitations (250 and398 nm). Further, the luminescence intensities and lifetimes increase significantly with increasingheat-treatment temperature of the doped samples. This is attributed to the reduction of H2O from thesurface of the particles and a non-radiative process after heat treatment.

1. Introduction

Rare-earth ion doped calcium molybdate (CaMoO4) has attractedmuch interest due to its wide luminescence applications.1–6 Theseinclude optical filters, solid state lasers, LED (light emittingdiodes), scintillators, microwave dielectrics, cryogenic detectorsand fluorescent lamps.7,8 Moreover, some of its attractive featuresinclude a high melting point (1445–1480 ◦C), refractive index(1.98), effective average decay time (14 ms), photo electron yield(9%) and it is also chemically resistant and non-hygroscopic.9

CaMoO4 shows a broad blue and/or green luminescence emissionpeak in the range 350–650 nm with a peak maximum around500 nm. The peak maximum however varies with particle size andoxygen vacancies in the lattice.10 It is related to the electron-holerecombination after excitation through the band gap of the chargetransfer (CT) from O to Mo (2p → 4d orbital) in the MoO4

2-

complex tetrahedron cluster itself.11 Liu et al.12 calculated the bandgap of pure CaMoO4 crystal using a density functional theorycalculation as 3.6 eV. Nearly consistently, Ryu et al.10 found it tobe 3.41 eV using a linearized-augmented-plane-wave technique.The experimentally measured value was slightly higher.10

aDepartment of Physics, Banaras Hindu University, Varanasi, 221005, IndiabChemistry Division, Bhabha Atomic Research Centre, Mumbai, 400085, In-dia. E-mail: [email protected]; Fax: +91-22-25505151; Tel: +91-22-2559321cUGC–DAE Consortium for Scientific Research, Indore, 452001, IndiadDepartment of Physics, Dr Hari Singh Gour University, Sagar, 470003,IndiaeTechnical Physics Division, Bhabha Atomic Research Centre, Mumbai,400085, IndiafMaterial Science Division, Bhabha Atomic Research Centre, Mumbai,400085, India

In fact, when a rare-earth dopant like Eu3+ is doped in calciummolybdate, what is predicted to happen to the luminescence? In ourstudy, the crystal structure of CaMoO4 has inversion symmetry.13

Eu3+ ions occupy Ca2+ sites and thus only magnetic dipole transi-tion should be allowed because Eu3+ ions have inversion symmetryaccording to Judd–Ofelt theory.14,15 In many reported papers, it issuggested that Eu3+ ions have an asymmetric environment.11,16,17

This has been claimed because the electric dipole transitionintensity is higher than the magnetic dipole transition of Eu3+

in the observed spectra. However, the relationship between crystalstructure and luminescence behavior has not been reported. Theunusual behaviour in this compound has not been discussed inearlier literature so far, to the best of the authors’ knowledge.This is explained based on the polarizability effect of the MoO4

tetrahedron, which is sharing with EuO8/CaO8 through an O atomin this work. The effect of particle size on luminescence behaviorand lifetime are also studied.

2. Experimental details

The nanoparticles of CaMoO4 : Eu3+ (Eu3+ = 0, 1, 3, 5, 7, 10 at.%) were prepared at low temperature (150 ◦C for 3 h) using ureahydrolysis in ethylene glycol.13 Finally, the as-prepared sample wasannealed at 500 and 900 ◦C for 4 h each.

The photoluminescence (PL) spectra of these powder phosphorswere recorded using Hitachi F-4500 spectrometer with a 150 WXe lamp as a source at the spectral resolution of 3 nm. All themeasurements were carried out at room temperature. PL decaywas recorded with an Edinburgh instrument F920 equipped with100 W ms flash xenon lamp as the excitation source.

This journal is © The Royal Society of Chemistry 2011 Dalton Trans., 2011, 40, 7595–7601 | 7595

Dow

nloa

ded

by B

habh

a A

tom

ic R

esea

rch

Cen

tre

on 2

7 Ju

ly 2

011

Publ

ishe

d on

22

June

201

1 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

1DT

1087

8FView Online

3. Results and discussion

3.1 Luminescence study

Fig. 1 shows excitation and emission spectra of as-prepared,500 and 900 ◦C heated samples of pure CaMoO4. The emissionwavelength is monitored at 450 nm and excitation wavelengthis fixed at 250 nm because Mo–O absorption occurs at 250–300 nm.18–22 In the excitation spectra, a rise in intensity starts below300 nm. In the emission spectra, a broad peak at 360–600 nmis observed. The peak maximum changes from 405 to 420 nmas heat-treatment temperature (particle size) increases, consistentqualitatively with earlier reports.21,22 The increase in intensity issuch that the 500 and 900 ◦C heated samples show a 2 and 10fold increase, respectively, as compared to that of the as-preparedsample. It is due to reduction of the surface effect, unwanted H2Omolecules and an organic moiety (EG) adsorbed on the surface ofparticles compared to the as-prepared sample.

Fig. 1 Excitation (monitoring emission at 450 nm), (b) emission (afterexcitation at 250 nm) spectra of as-prepared, 500 and 900 heated pureCaMoO4 samples.

Fig. 2 shows excitation spectra of as-prepared, 500 and 900 ◦Cheated samples of 5 at. % Eu3+ doped CaMoO4 nanophosphors.The emission wavelength is monitored at 615 nm. In the excitationspectra, there is a broad peak at 220–300 nm and sharp peaks withlow intensity at 320–500 nm. The broad peak arises from Eu–Oand Mo–O charge transfers (CT).11,16,18–24 This can be clearly seenin the inset of Fig. 2 for the 900 ◦C heated sample and its dataare fitted with a Gaussian distribution (discussed later). Two mainpeaks could be resolved. The calculated peaks at 247 and 272 areassigned to Eu–O and Mo–O CT bands, respectively. These peakscould not be distinguished clearly in the case of as-prepared or500 ◦C heated samples. This is possibly due to a quenching effectfrom the surface or H2O over the particles. The sharp intensepeaks at 398 (7F0→5L6) and 468 nm (7F0→5D2) are due to the f–ftransitions of Eu3+.23,24

Fig. 3 shows emission spectra of as-prepared, 500 and 900 ◦Cheated samples of 5 at. % Eu3+ doped CaMoO4 nanophosphors.The excitation wavelength used is 250 nm. In the emission spectra,a broad peak at 360–500 nm and sharp peaks at 592 and 614 nmcould be observed. The broad peak is assigned to the emissiondue to host CaMoO4. The sharp peaks at 592 and 614 nm are

Fig. 2 Excitation spectrum (monitoring emission at 615 nm) of 5 at. %Eu3+ doped CaMoO4 as-prepared, 500 and 900 ◦C heated samples. Inset(a) shows expansion of excitation spectrum of 900 ◦C heated sample inthe range 220–300 nm and fitting to data in the range 220–300 nm afterbase line correction. Inset (b) shows expansion of excitation spectrum of900 ◦C heated sample in the range 325–500 nm.

Fig. 3 Emission spectra (after excitation at 250 nm) of 5 at. % Eu3+

doped CaMoO4 as-prepared, 500 and 900 ◦C heated samples, Inset showsexpansion of 900 ◦C heated sample and data are fitted with a Gaussianequation. The fitted parameters are given in the figure itself.

assigned to the magnetic (5D0→7F1) and electric (5D0→7F2) dipoletransitions of Eu3+, respectively. The relative intensity of Eu3+

with respect to the host increases with increasing heat-treatmenttemperature. This is related to extent of reduction of quenchingeffect from surface or H2O over particles with heat treatment.

As-prepared nanoparticles of 5 at. % Eu3+ doped CaMoO4

were excited at four different excitation wavelengths at240, 250, 300 and 398 nm to choose the optimum excitationwavelength (Fig. 4a). The intensity increases in the followingorder of excitations: 300, 398, 250 and 240 nm. Thus, excitation at240 nm gives the highest emission intensity. In the case of samesample heated at 500 or 900 ◦C (Fig. 4b, 4c), the intensity increasesin the following order of excitations: 398, 300, 240 and 250 nm.This order is different from the as-prepared sample. It means thatenergy transfer process from the host or Eu–O to Eu3+ shifts to

7596 | Dalton Trans., 2011, 40, 7595–7601 This journal is © The Royal Society of Chemistry 2011

Dow

nloa

ded

by B

habh

a A

tom

ic R

esea

rch

Cen

tre

on 2

7 Ju

ly 2

011

Publ

ishe

d on

22

June

201

1 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

1DT

1087

8FView Online

Fig. 4 Luminescence spectra of (a) as-prepared, (b) 500 ◦C and (c) 900 ◦C heated 5 at. % Eu3+ doped CaMoO4 nanoparticles at different excitationwavelengths. (d) Luminescence spectra of as-prepared sample of 10 at. % Eu3+ doped CaMoO4.

higher wavelength (from 240 to 250 nm) with increasing heat-treatment temperature. For 500 or 900 ◦C heated samples, Mo–Ocharge transfer excitation peak (300 nm) dominates over the directEu3+ (398 nm) excitation. There is large gap between emissionintensities from different excitation wavelengths in the case ofthe as-prepared sample. However, this gap reduces with heat-treatment temperature up to 900 ◦C. This is related to the extentof reduction of the non-radiative process arising from danglingbonds, H2O and the organic moiety (EG) on the particles.

In order to see the optimum concentration of Eu3+ in CaMoO4

samples for the highest luminescence, luminescence spectra wererecorded for as-prepared, 500 and 900 ◦C heated samples at fourdifferent excitation wavelengths 240 (host or Eu–O CT), 250 (hostor Eu–O CT), 300 (Mo–O CT) and 398 nm (direct Eu3+). Typicalemission spectra of 900 ◦C heated samples are shown in Fig. 5.

Fig. 5 Luminescence spectra of 900 ◦C heated CaMoO4 : Eu3+ (Eu3+ = 1,3, 5, 7 and 10 at. %) nanoparticles at 250 nm excitation.

In order to compare emission intensity, the Gaussian equationis used.23–30 The data are fitted in range 583 to 630 nm inwhich magnetic dipole and electric dipole allowed transitions (592and 614 nm) occur. The parameters such as area (A) and fullwidth at half maximum intensity (FWHM, W ) for the magneticdipole and electric dipole allowed transitions are characterizedby (A1, W 1) and (A2, W 2), respectively. One typical fit to dataof 5 at. % Eu3+ doped CaMoO4 is shown as inset of Fig. 3using a Gaussian equation after background correction. Thefitting parameters are shown in Fig. 3, inset. Since electric dipoletransition is sensitive to the asymmetric environment of Eu3+, onlythe variations of A2 and W 2 with Eu3+ concentrations are discussedhere.

In Fig. 6a, A2 of the as-prepared samples decreases with Eu3+

up 10 at. %. The decrease in luminescence with Eu3+ is due theconcentration quenching effect. It was reported that 8 at. % Eu3+

gave the optimum luminescence intensity in CaMoO4.18 In lowerEu3+ doping (1 and 3 at. %), the luminescence intensity increasesin the following order of excitation wavelengths 398, 300, 240and 250 nm. For 5 at. % Eu3+ doping, luminescence intensityincreases in the following order of excitation wavelengths 300,398, 250 and 240 nm. This variation in emission intensity withexcitation wavelength can be correlated with appearance of extraphases (MoOn) and Eu–O–Mo for 5 at. % Eu3+ doping in theXRD patterns wherein 1 and 3 at. % Eu3+ doped CaMoO4 samplesshow substitution of Ca2+ by Eu3+. In Fig. 6a (inset), W 2 of as-prepared samples increases from 7.5 to 13 nm with increasing ofEu3+ concentration up to 10 at. %. Larger W 2 means a broaderpeak and usually, a less luminescence intensity. This corroborateswith a decrease in luminescence for 7 to 10 at. % Eu3+ doping (Fig.6a).

This journal is © The Royal Society of Chemistry 2011 Dalton Trans., 2011, 40, 7595–7601 | 7597

Dow

nloa

ded

by B

habh

a A

tom

ic R

esea

rch

Cen

tre

on 2

7 Ju

ly 2

011

Publ

ishe

d on

22

June

201

1 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

1DT

1087

8FView Online

Fig. 6 Variation of A2 and W 2 for (a) as-prepared, (b) 500 and (c) 900 ◦C heated CaMoO4 : Eu nanophosphors at different excitations.

The asymmetric environment of europium ion (Eu3+) in thehost lattice can be calculated by measuring the ratio of the electricto the magnetic dipole transitions (known as asymmetric factor(A21)).23–30 A21 values for the as-prepared samples at differentexcitation wavelengths (240, 250, 300 and 398 nm) are calculated.For lower concentration of Eu3+ (1, 3, 5 at. %), A21 is found to bemaximal (7–9) while for 7 to 10 at. % Eu3+ it is minimal (1–2). Thisis related to variation in A2. In highly Ln3+ doped samples and/orsystems with highly quenching environment such as H2O, C–Hgroups or surface dangling bonds, the following can be concluded:(1) intensities of the magnetic and electric dipole transitions aredominated by cross-relaxation among Ln3+ ions and quenchingenvironment and thus, intensity is reduced significantly. (2) Theirintensity ratio (A21) is close to 1 or 2. In this situation, occupancyof Ln3+ ions in asymmetric environment could not be determinedby observing the A21 alone. (3) Their FWHM (W 1 and W 2) arevery large as compared to that of diluted Ln3+ doping or crystallinesamples having reduction of quenching effect from surroundingor surface of particles. For example, equal values of intensity in A1

and A2 or FWHM in W 1 and W 2 are shown by as-prepared samplewith 10 at. % Eu3+ dopant after different excitation wavelengths(Fig. 4d). Their intensities (A1 or A2) are very weak due to the con-centration quenching effect and the presence of H2O over particles.

Fig. 6b shows the variations of A2 and W 2 of 500 ◦C heatedsamples at different excitation wavelengths. A2 increases withEu3+ up to 3 at. % and decreases with further increase of Eu3+.This is related to the concentration quenching effect above 3at. % Eu3+. W 2 increases from 7.2 to 10 nm with increasingof Eu3+. A21 values are found in the range 2–10 depending onEu3+ concentration and excitation wavelength. Fig. 6c shows thevariations of A2 and W 2 of 900 ◦C heated samples at differentexcitation wavelengths. A2 increases with Eu3+ up to 5 at. % anddecreases with further increase of Eu3+. W 2 decreases from 7.4to 7 nm with increasing Eu3+. A21 values are found in the range4.2–10. From heat-treatment study, it can be concluded that A2 orA21 increases, whereas W 2 decreases with heat-treatment.

It is worth mentioning that CaMoO4 has a tetragonal structure,with space group I41/a (C6

4h).13 This structure has body centered

symmetry, similar to the reported one.1 In CaO8, there aretwo different Ca–O bond lengths, but with inversion symmetry.Likewise, there is only one Mo–O bond length with no inversionsymmetry in MoO4 (tetrahedron structure). It is expected thatonly the magnetic dipole transition should be allowed if Eu3+ ionsoccupy Ca2+ sites according to Judd–Ofelt theory.14,15,31 If there isa slight distortion from symmetry, distribution of magnetic andelectric dipole transitions of Eu3+ could be observed. From the

7598 | Dalton Trans., 2011, 40, 7595–7601 This journal is © The Royal Society of Chemistry 2011

Dow

nloa

ded

by B

habh

a A

tom

ic R

esea

rch

Cen

tre

on 2

7 Ju

ly 2

011

Publ

ishe

d on

22

June

201

1 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

1DT

1087

8FView Online

Table 1 Parameters obtained after bi-exponential equation fitting to the decay data of as-prepared, 500 and 900 ◦C heated samples at 250 and 398 nmexcitations

Sample Excitation l (nm) Eu3+ (at. %) I 1(%) t 1 (ms) I 2(%) t 2 (ms) t av (ms) c2a

As-prepared 250 1 39 304 61 731 566 1.8843 56 280 44 733 481 1.835 54 204 46 670 417 2.3610 54 72 46 280 167 1.529

398 1 5 3 95 403 383 1.7643 7 2 93 385 357 1.1235 6 2 94 336 316 1.22110 3 2 97 130 126 2.018

500 ◦C 250 1 39 341 61 1034 764 2.1993 45 334 55 956 676 1.9225 44 249 56 723 513 0.9910 69 160 31 712 332 1.891

398 1 18 3 82 600 495 1.683 12 2 88 496 438 1.7125 8 2 92 478 439 1.59110 5 2 95 210 201 2.135

900 ◦C 250 1 40 377 60 803 631 3.0223 72 502 28 876 606 1.215 86 497 14 835 545 1.1410 73 162 27 420 231 0.832

398 1 8 2 92 607 557 1.5583 3 118 97 596 580 1.9855 94 499 6 3021 641 3.30610 50 119 50 277 198 1.157

a c 2 = −∑w X F nk

k

k k

22[ ] , c2 is goodness of fit, wk is a weighting factor for data points ( w Fk k=1/ ), X k is the fitted data and Fk is the

luminescence decay data.

XRD study, there is substitution of Ca2+ sites by Eu3+ up to 3 at.%. If Eu3+ ions do not occupy Ca2+ sites (and instead occupy a grainboundary), broadening in the XRD peak could be observed. Thiswas not observed. We feel that the intensity of the magnetic dipoletransition should be more than an electric dipole transition even ifthere is a slight distortion from symmetry. However, the oppositeeffect is observed. Similar unusual luminescence observations withthe respect to structural model were reported in REPO4 : Ln3+

(RE = Y, La, Gd; Ln3+ = Eu3+, Dy3+, Tb3+).24,32–34 A possibleway to answer this is due to MoO4 polarization in luminescence.This polarization changes the crystal field of EuO8 and should beconsidered when Judd–Ofelt parameters are calculated.18,35

3.2 Luminescence decay study

The luminescence decay curves of the level 5D0 (615 nm) has beenmeasured for as-prepared, 500 and 900 ◦C heated samples of Eu3+

doped CaMoO4. Fig. 7, 8 and 9 show the decay curves for 5D0→7F2

emission for these samples when excited at 250 and 398 nm. Thedecay curves are not well fitted by the mono-exponential equation(I = I 1 e(-t/t1)). Fig. 10a shows the mono-exponential curve fittingof 5 at. % Eu3+ doped CaMoO4. The fitted parameter are givenin the figure itself with c2 = 19.35. The definition of c2 is given inTable 1. The fitting behavior can be clearly understood by plottingln(I) vs. t, which is shown in Fig. 10a (inset). The fitted straightline does not match with the decay data points. This suggests thatthe environment of the Eu3+ ions in the lattice are not the same indifferent positions.

All decay data are fitted by using bi-exponential decay equation,which is expressed as

I = I 1 e(-t/t1) + I 2 e(-t/t2), (1)

Fig. 7 Luminescence decay spectra of as-prepared Eu3+ doped CaMoO4

(Eu3+ = 1, 3, 5 and 10 at. %) nanoparticles at (a) 250 and (b) 398 nmexcitations.

This journal is © The Royal Society of Chemistry 2011 Dalton Trans., 2011, 40, 7595–7601 | 7599

Dow

nloa

ded

by B

habh

a A

tom

ic R

esea

rch

Cen

tre

on 2

7 Ju

ly 2

011

Publ

ishe

d on

22

June

201

1 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

1DT

1087

8FView Online

Fig. 8 Luminescence decay spectra of 500 ◦C heated Eu3+ doped CaMoO4

(Eu3+ = 1, 3, 5 and 10 at. %) nanoparticles at 250 and 398 nm excitations.

where I 1 and I 2 are the intensities at different time intervals and t 1

and t 2 their corresponding lifetimes. The average lifetime can becalculated using the equation,

t t tav

I I

I I=

++

1 1 2 2

1 2

(2)

Typical plot of I versus time for as-prepared 5 at. % Eu3+

doped CaMoO4 fitted using bi-exponential decay is shown in Fig.10b with goodness of fitting parameter c2 = 2.36. Bi-exponentialdecay behavior is observed when energy transfer takes placefrom donor/host to activator.11 It is observed that lifetime valuedecreases with increasing Eu3+ concentration. This is due toconcentration quenching effect. The average lifetime (t av) valuesfor the 5D0 level are 291, 175, 146 and 94 ms for 1, 3, 5 and 10 at.% Eu3+ ion concentration, respectively under 250 nm excitation.Moreover, the lifetime value measured for as-prepared samples ofEu3+ doped CaMoO4 at 398 nm excitation is shorter as comparedto that at 250 nm excitation (see Table 1). This is due to decreasein luminescence intensity at 398 nm excitation compared to that at250 nm excitation. Longer lifetime value or higher intensity afterexcitation at 250 nm is related to the high population of excitedphotons at the 5D0 level due to energy transfer from the host orEu–O CT band to Eu3+.

Table 1 gives the parameters (t 1, I 1, t 2, I 2) obtained afterbi-exponential equation fitting to data of as-prepared, 500 and900 ◦C samples. In the as-prepared sample (1 at. % Eu3+), t 2

(I 2%) is found to be 731 ms (61%), which is two times t 1 (I 1%) =304 ms (39%) after excitation at 250 nm. The same sample afterexcitation at 398 nm gives t 2 (I 2%) = 403 ms (95%), which is 134times t 1 (I 1%) = 3 ms (5%). Thus, t 1 from 250 nm excitation has

Fig. 9 Luminescence decay spectra of 900 ◦C heated Eu3+ doped CaMoO4

(Eu3+ = 1, 3, 5 and 10 at. %) nanoparticles at 250 and 398 nm excitations.

a contribution (39%) in the average lifetime value while t 1 from398 nm excitation has a minimum contribution (5%) in the averagelifetime value. Longer t 1 with significant contribution from 250 nmexcitation arises from the energy transfer (ET) process from thehost or Eu–O CT band to Eu3+. It was reported that lifetimevalue of the ET process was found to be ~15 ms.11 Shorter t 1

with lower contribution from 398 nm excitation arises from eithersurface Eu3+ or inhomogeneity. The reported lifetime values ofEu3+ (5 at. %) in CaMoO4 were 494 and 455 ms.11,18 Some reportedlifetimes values of other Eu3+ doped compounds: Y2O3 : Eu,Gd2O3 : Eu, YVO4 : Eu, GdVO4 : Eu, YPO4 : Eu, GdPO4 : Eu andLaPO4 : Eu are 1.6, 1.3, 0.247, 0.583, 2.07, 1.0 and 3.56 ms,respectively.23,24,30,33,36–38 In the case of 500 or 900 ◦C heated samples,the lifetime value is longer than that of the as-prepared samples.This is related to the increase of particle size and reduction ofsurface to volume ratio and surface defects after heat-treatment.Also, the contribution of non-radiative rate from O–H (3410 cm-1)reduces with increasing heat-treatment temperature.

4. Conclusions

Luminescence intensity increases with Eu3+ up to 5 at. % and de-creases with further increase of Eu3+. The decrease of luminescencewith increasing Eu3+ is related to a concentration quenching effect.The luminescence intensity improves with heat-treatment due toa reduction in non-radiative processes from the surface and O–H groups. From luminescence decays on direct excitation at 398and indirect excitation at 250 nm, energy transfer process fromMo–O CTB to Eu3+ and surface contribution to lifetime could be

7600 | Dalton Trans., 2011, 40, 7595–7601 This journal is © The Royal Society of Chemistry 2011

Dow

nloa

ded

by B

habh

a A

tom

ic R

esea

rch

Cen

tre

on 2

7 Ju

ly 2

011

Publ

ishe

d on

22

June

201

1 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

1DT

1087

8FView Online

Fig. 10 (a) Mono- and (b) bi-exponential fittings to luminescence decaydata of as-prepared 5 at. % Eu3+ doped CaMoO4 nanoparticles. Fittingparameters are shown in their figures. Inset of (a) shows the ln(I) vs. t plot).

distinguished. The intensity of the electric dipole transition is morethan that of the magnetic dipole transition of Eu3+ even when thecrystal structure has inversion symmetry. This can be explainedfrom the polarizability effect of the MoO4 tetrahedron.

Acknowledgements

A. K. Parchur acknowledges the financial support provided byUniversity Grants Commission, Govt. of India through the D.S. Kothari Post Doctoral Fellowship. The authors thank Dr T.Mukherjee, Chemistry Group and Dr D. Das, Chemistry Division,BARC for their support and encouragement during this work.

References

1 Y. Zhang, N. A. W. Holzwarth and R. T. Williams, Phys. Rev. B:Condens. Matter, 1998, 57, 12738.

2 L. F. Johnson, L. G. Van Uitert, J. J. Rubin and R. A. Thomas, Phys.Rev., 1964, 133, A494.

3 A. S. Barker Jr, Phys. Rev., 1964, 135, A742.4 S. P. S. Porto and J. F. Scott, Phys. Rev., 1967, 157, 716.5 Y. B. Abraham, N. A. W. Holzwarth, R. T. Williams, G. E. Matthews

and A. R. Tackett, Phys. Rev. B: Condens. Matter, 2001, 64,245109.

6 A. Senyshyn, H. Kraus, V. B. Mikhailik, L. Vasylechko and M. Knapp,Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 014104.

7 V. S. Marques, L. S. Cavalcante, J. C. Sczancoski, A. F. P. Alcantara,M. O. Orlandi, E. Moraes, E. Longo, J. A. Varela, M. S. Li and M. R.M. C. Santos, Cryst. Growth Des., 2010, 10, 4752.

8 V. M. Longo, A. T. de Figueiredo, A. B. Campos, J. W. M. Espinosa, A.C. Hernandes, C. A. Taft, J. R. Sambrano, J. A. Varela and E. Longo,J. Phys. Chem. A, 2008, 112, 8920.

9 A. N. Annenkov, O. A. Buzanov, F. A. Danevich, A. S. Georgadze, S.K. Kim, H. J. Kimd, Y. D. Kim, V. V. Kobychev, V. N. Kornoukhov,M. Korzhik, J. I. Lee, O. Missevitchg, V. M. Mokina, S. S. Nagorny,A. S. Nikolaiko, D. V. Poda, R. B. Podviyanukb, D. J. Sedlak, O. G.Shkulkova, J. H. So, I. M. Solsky, V. I. Tretyakb and S. S. Yurchenkob,Nucl. Instrum. Methods Phys. Res., Sect. A, 2008, 584, 334.

10 J. H. Ryu, B. G. Choi, J.-W. Yoon, K. B. Shim, K. Machib and K.Hamadab, J. Lumin., 2007, 124, 67.

11 Z. Hou, R. Chai, M. Zhang, C. Zhang, P. Chong, Z. Xu, G. Li and J.Lin, Langmuir, 2009, 25, 12340.

12 T. Liu, J. Chen and F. Yan, J. Lumin., 2009, 129, 101.13 A. K. Parchur and R. S. Ningthoujam, Dalton Trans., 2011, 40,

DOI: 10.1039/C1DT10327J.14 B. R. Judd, Phys. Rev., 1962, 127, 750.15 G. S. Ofelt, J. Chem. Phys., 1962, 37, 511.16 S. Yan, J. Zhang, X. Zhang, S. Lu, X. Ren, Z. Nie and X. Wang, J.

Phys. Chem. C, 2007, 111, 13256.17 D. Gao, Y. Li, X. Lai, Y. Wei, J. Bi, Y. Li and M. Liu, Mater. Chem.

Phys., 2011, 126, 391.18 F. Lei and B. Yan, J. Solid State Chem., 2008, 181, 855.19 P. Yang, C. Li, W. Wang, Z. Quan, S. Gai and J. Lin, J. Solid State

Chem., 2009, 182, 2510.20 Y. Yu-Ling, L. Xue-Ming, F. Wen-Lin, L. Wu-Lin and T. Chuan-Yi, J.

Alloys Compd., 2010, 505, 239.21 A. Xie, X. Yuan, S. Hai, J. Wang, F. Wang and L. Li, J. Phys. D: Appl.

Phys., 2009, 42, 105107.22 Y. Yin, Y. Gao, Y. Sun, B. Zhou, L. Ma, X. Wu and X. Zhang, Mater.

Lett., 2010, 64, 602.23 M. Niraj Luwang, R. S. Ningthoujam, Jagannath, S. K. Srivastava and

R. K. Vatsa, J. Am. Chem. Soc., 2010, 132, 2759.24 N. Yaiphaba, R. S. Ningthoujam, N. Shanta Singh, R. K. Vatsa, N. R.

Singh, S. Dhara, N. L. Misra and R. Tewari, J. Appl. Phys., 2010, 107,034301.

25 N. S. Singh, R. S. Ningthoujam, S. D. Singh, B. Viswanadh, N. Manojand R. K. Vatsa, J. Lumin., 2010, 130, 2452.

26 L. R. Singh, R. S. Ningthoujam, V. Sudarsan, I. Srivastava, S. D. Singh,G. K. Dey and S. K. Kulshreshtha, Nanotechnology, 2008, 19, 055201.

27 M. N. Luwang, R. S. Ningthoujam, S. K. Srivastava and R. K. Vatsa,J. Am. Chem. Soc., 2011, 133, 2998.

28 R. S. Ningthoujam, Chem. Phys. Lett., 2010, 497, 208.29 R. S. Ningthoujam, V. Sudarsan, S. V. Godbole, A. K. Tyagi, L. Kienle

and S. K. Kulshreshtha, Appl. Phys. Lett., 2007, 90, 173113.30 R. S. Ningthoujam, R. Shukla, R. K. Vatsa, V. Duppel, L. Kienle and

A. K. Tyagi, J. Appl. Phys., 2009, 105, 084304.31 C. Gorller-Walrand and K. Binnemans, Spectral intensities of f–f

transitions, in: Handbook on the Physics and Chemistry of Rare Earths,ed. K. A. Gschneidner Jr and L. Eyring, North-Holland Publishers,Amsterdam, 1998, vol. 25, ch. 167, pp. 101–264.

32 N. Yaiphaba, R. S. Ningthoujam, R. K. Vatsa and N. R. Singh, Eur. J.Inorg. Chem., 2010, 2682.

33 G. Phaomei, R. S. Ningthoujam, W. R. Singh, N. S. Singh, M. N.Luwang, R. Tewari and R. K. Vatsa, Opt. Mater., 2010, 32, 616.

34 G. Pan, H. Song, Q. Dai, R. Qin, X. Bai, B. Dong, L. Fan and F. Wang,J. Appl. Phys., 2008, 104, 084910.

35 W. Wang, G. K. Liu, M. G. Brik, L. Seijo and D. Shi, Phys. Rev. B:Condens. Matter Mater. Phys., 2009, 80, 155120.

36 N. S. Singh, R. S. Ningthoujam, L. R. Devi, N. Yaiphaba, V. Sudarsan,S. D. Singh, R. K. Vatsa and R. Tewari, J. Appl. Phys., 2008, 104,104307.

37 D. Lahiri, R. S. Ningthoujam, D. Bhattacharyya and S. M. Sharma, J.Appl. Phys., 2010, 107, 054316.

38 M. N. Luwang, R. S. Ningthoujam, S. K. Srivastava and R. K. Vatsa,J. Mater. Chem., 2011, 21, 5326.

This journal is © The Royal Society of Chemistry 2011 Dalton Trans., 2011, 40, 7595–7601 | 7601

Dow

nloa

ded

by B

habh

a A

tom

ic R

esea

rch

Cen

tre

on 2

7 Ju

ly 2

011

Publ

ishe

d on

22

June

201

1 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

1DT

1087

8FView Online