Investigation of the luminescent properties of Ce3+ doped and Ce3+/Mn2+ co-doped CaAl2Si2O8 phosphors

W. B. Dai*
Institut des Matériaux Jean Rouxel, Université de Nantes, CNRS, 2 rue de la Houssinière, BP32229, 44322 Nantes Cedex 03, France. E-mail: wubin.dai@foxmail.com

Received 29th November 2013 , Accepted 10th February 2014

First published on 10th February 2014


Abstract

A series of Ce3+ doped and Ce3+/Mn2+ co-doped CaAl2Si2O8 phosphors have been synthesized in a reducing atmosphere (Ar/H2) via a solid-state route and their photoluminescent properties have been investigated at room temperature. For the CaAl2Si2O8:Ce3+ phosphor, a broad emission band at 430 nm is observed in addition to characteristic bands of Ce3+ peaking at 334 nm and 356 nm. The extra band is assigned to the presence of cationic vacancies naturally created to counterbalance the substitution of Ce3+ ions for Ca2+ alkaline-earth species. Its intensity can be significantly reduced when Ca2+ cations are replaced by the reaction image file: c3ra47156j-t1.tif i.e. when the amount of alkaline earth vacancies induced by the presence of Ce3+ cations is minimized. Moreover, the co-doping Ce3+/Mn2+ goes along with the appearance of energy transfers that make possible fine tuning of the CIE(x, y) parameters from blue to warm white and even yellow.


1. Introduction

White light-emitting diodes (LEDs) have been extensively investigated because they are promising candidates as light sources for everyday life. In particular, they offer a huge variety of benefits such as low power consumption, high efficiency, long lifetime, high reliability and eco-friendliness.1–5 To achieve white inorganic LEDs, three approaches are commonly considered: (1) the assembly of three LEDS emitting in the red, green and blue, respectively, (2) the use of a UV-LED chip coated with a blend of UV-excitable RGB phosphors, and (3) the combination of a yellow phosphor (namely, Y3Al2Si3O12:Ce3+) deposited at the surface of a blue chip (e.g. InGaN). Unfortunately, such devices still suffer from a poor colour rendering index (CRI of about 70%) that limits the full deployment of such an innovative technology. Moreover, the last technical solution, often favoured so far for cost reasons, exhibits a bright blue-halo effect that is currently subject of many controversies due to its potential risk for the retina of young children and elderly persons.6 Moreover, there still exist many problems to overcome for such devices to capture a share of the lighting market, such as the low stability of phosphor colour temperature with time and temperature, the untimely deterioration of phosphors, their uncorrelated lifetimes when mixed together in a blend, the difficulty in generating warm white lights, etc.7 In that context, an option to provide access to white UV-LEDs with high efficiency and reproducibility could consist in synthesising a single-phase phosphor with a broad emission band mimicking the solar spectrum and easily stimulated by a UV radiation.8–13 This would alleviate the complicated optimization of adjusting the composition of phosphor blends and controlling the absorption of the blue emission, but it would require a good adaptability and flexibility of the host lattice to accommodate many substitutions to finely tune optical characteristics (i.e. emission spectrum, quantum yield, etc.).

In oxide-based luminescent materials, Ce3+ and Eu2+ ions are well-known to efficiently absorb UV rays, and to supply broad emission bands in the visible range.14,15 Nevertheless, to achieve white light, co-doping seems to be inevitable even if the aforementioned activators may occupy distinguishable crystal sites leading to different emission wavelengths. In that context, the use of co-dopants can be of particular interest to adapt the emission spectra to specific requirements, for instance to enrich the emission of red and yellow. Namely, numerous possible energy transfers between luminescent centres can generate warm white light with enhanced colour rendering index. In that framework, Mn2+-doped materials are reported to exhibit emission bands ranging from about 500 to 700 nm depending on the strength of the crystal field experienced by the activator. Namely, a green emission band is observed for a weak crystal field in tetrahedral sites, whereas red luminescence is associated with a strong crystal field in octahedral sites. An intermediate crystal field will give rise to yellow or orange fluorescences, all of which are related to the 4T16A1 transition.14,16,17 However, as the d–d electric dipole absorptions are forbidden by spin and parity, Mn2+ cations are very difficult to pump. A universal solution to figure out this situation consists in associating Mn2+ cations with sensitizers such as Eu2+ or Ce3+, i.e. materials with a high absorption cross-section.8–13 This concept has already been applied to several matrices to provide potential phosphors for white LEDs.

Eu2+-doped alumino-silicate hosts with the CaAl2Si2O8 composition are of great interest for luminescent materials because they exhibit efficient luminescence in the visible spectrum coupled with a high chemical stability. Moreover, they are also quite easy to synthesize.9,18–21 In that context, Yang et al.9 investigated the energy transfer between Eu2+ and Mn2+ in CaAl2Si2O8 and claim that this material could be regarded as a potential phosphor for white light UV-LED. If such an assertion holds under the stringent quality requirements for such a specific application, it is clear that the CIE chromatic parameters can be tuned continuously from (0.17, 0.11) to (0.33, 0.31) going from Ca0.99Eu0.01Al2Si2O8 to Ca0.74Eu0.01Mn0.25Al2Si2O8 compositions under excitation at 354 nm. This prompted us to seek new single phase phosphors with adaptable chromatic characteristics. Besides, it should be recalled that Eu doped CaAl2Si2O8 materials have also been reported as long persistent luminescent materials by Clabau et al.18 Consequently, in addition to producing white light, these materials might be used in the distant future for lighting displays supplied with electrical current in intermittent ways that could save energy. So far, no investigation has been carried out on energy transfers from Ce3+ to Mn2+ in this series of compounds. Thus we embarked on the elucidation of the optical properties of CaAl2Si2O8 host lattices doped and co-doped with Ce3+ and Mn2+ cations.

2. Experimental section

2.1. Raw materials and synthesis process

Conventional high temperature solid state reactions were performed to synthesize Ce3+/Mn2+ co-activated CaAl2Si2O8 compounds (hereafter abbreviated CASO). Oxide and carbonate materials, i.e. CaCO3 (99.997%, Alfa Aesar), SiO2 (99.99% Chempur), γ-Al2O3 (99.99%, Alfa Aesar), Ce2(CO3)3 (99.5%, Rhodia), and MnO (99.99%, Alfa Aesar) were used as starting materials. Precursors were weighed in stoichiometric amounts and ball milled with a Fritsch Pulverisette 7 for two hours in ethanol. After drying at 100 °C, mixtures were annealed at 1350 °C for 50 h in argon/hydrogen (95/5 vol%) atmosphere.

2.2. Structure characterizations

The phase purity and crystal structure of the powder were examined by powder X-ray diffraction (XRD) profiles which were measured with a Bruker AXS D8 advanced automatic diffractometer with Cu K-L3 radiation (germanium monochromator) operated at 40 kV and 40 mA. Furthermore, the structure refinements based on Gebert's crystal investigations22 are carried out with the Jana2006 Beta version software.23 The instrumental function was expressed in terms of the geometry of the diffractometer with the relevant parameters which reported in Table 1. The XRD patterns were collected in the 10–90° 2θ range and all the doped samples diffraction peaks are in good agreement with those reported in JCPDS files 89-1462 (CASO) as exemplified in Fig. 1 for CASO:9%Ce3+, 6%Mn2+. No characteristic peaks of the dopants (Ce3+/Mn2+) were observed after a full pattern matching analysis, which means a solid-state solution is produced in all the samples to be researched. Note here, the experimental spectrum is not fitted well with the calculate spectrum. Three reasons can be put forward to explain, (i) the sample is the powder diffraction and CASO structure is triclinic which has lower symmetry, these factors easily induce overlap of the diffraction peaks and, (ii) due to the distort of the bond angle and length induced by the different radii between the Eu3+ and Ca2+ ions and, (iii) in order to make the charge balance, there would exists lots of holes and many Eu3+ ions may go into the interval of the lattice also caused the bad refinement induced by the different valence between Ce3+ and Ca2+ ions. Meanwhile, the lattice parameters for Ca0.91Mn0.09Al2Si2O8, Ca0.865Ce0.09Al2Si2O8 and Ca0.775Ce0.09Mn0.09Al2Si2O8 powder were also calculated based on the experimental XRD profiles with cell refinement software (Jana2006) by using a Rietveld procedure with the fundamental parameter approach in the I[1 with combining macron] space group (Table 2).
Table 1 Instrumental data used for Rietveld refinements of CASO and its Ce3+, Mn2+-doped derivatives
Primary and second radius 217.5 mm
Receiving slit length 16 mm
Glancing angle 13.65°
Source and sample length 12 mm
Primary soller slit aperture 2.5°
Reception slit divergence angle 0.2°
Receiving slit width 0.1 mm
Peak-shape function Lorentzian



image file: c3ra47156j-f1.tif
Fig. 1 Observed, calculated and difference X-ray diffraction pattern of an Ce3+ (0.09) and Mn2+ (0.06)-co-doped CASO phosphor in the [10–45] 2θ range (inset is given the total pattern).
Table 2 Crystallographic details for Ca0.925Eu0.05Al2Si2O8 and CASO HL
Formula Ca0.91Mn0.09Al2Si2O8 Ca0.865Ce0.09Al2Si2O8 Ca0.775Ce0.09Mn0.09Al2Si2O8
a According to Angel et al., the anorthite form of CASO may crystallize in the I[1 with combining macron] space group with similar cell parameters than those reported in Table 2. Then the multiplicity of the Wyckoff sites occupied by calcium cations is doubled compared to P[1 with combining macron] SG but the occupancy rate is about 50%. So far, based on X-ray patterns collected in this study, we couldn't distinguish between the two models and we systematically privileged the I[1 with combining macron] SG.b Rwp = weighted profile residual factor.c GOF = goodness-of-fit on F2.
Formula weight (g mol−1) 2236.36 2286.26 2373.11
Cryst. syst. Anorthic(triclinic) Anorthic(triclinic) Anorthic(triclinic)
Space groupa I[1 with combining macron] I[1 with combining macron] I[1 with combining macron]
a (Å) 8.1721(3) 8.1917(3) 8.1778(3)
b (Å) 12.8683(4) 12.8763(5) 12.8695(4)
c (Å) 14.1643(5) 14.1741(5) 14.1628(5)
α (deg.) 93.348(3) 93.069(3) 93.312(3)
β (deg.) 115.769(2) 115.688(2) 115.716(2)
γ (deg.) 91.156(2) 91.212(3) 91.135(2)
V3) 1337.39(10) 1343.74(10) 1339.01(10)
Z 8 8 8
Dcalcd (g cm−1) 2.7629(2) 2.7493(3) 2.7590(2)
Rp (%) 10.79 11.10 12.58
Rwp (%)b 13.56 14.50 16.08
GOFc 1.50 1.33 1.59


In Fig. 2 are reported the evolution of refined lattice volume for CASO matrixes with different dopant concentrations (Ce3+ and Mn2+ separately). For the CASO matrix, the slight difference of ionic radii24 between Ce3+ and Ca2+ (r (Ce3+) = 1.01 Å and r (Ca2+) = 1.00 Å when CN = 6, r (Ce3+) = 1.07 Å and r (Ca2+) = 1.06 Å when CN = 7) leads to a continuous increase of the lattice volume. At the opposite, a shrink of the cell volume is observed when Ca2+ cations are replaced by Mn2+ ones (r (Mn2+) = 0.67 Å and 0.90 Å when CN = 6 and 7, respectively). All these results indicate that in all cases, a solid solution is formed because the evolution of the cell parameters regularly evolves with dopant ions concentration.


image file: c3ra47156j-f2.tif
Fig. 2 The variations in the lattice volume with Ce3+ and Mn2+ concentrations in CASO.

2.3. Optical measurements

The photoluminescence (PL) and PL excitation (PLE) spectra were recorded at room temperature on a Fluorolog-3 spectrofluorometer (Jobin Yvon Instruments) equipped with a xenon light as excitation source. Emission and excitation spectra were corrected for the contribution of the monochromator and the wavelength dependence of the lamp output. Chromaticity coordinates in the CIE 1931 color space were calculated by using the OptPropMatlab toolbox.25 The room PL decays were acquired with a streak camera Hamamatsu C7700 coupled to an imaging spectrograph. The excitation line was obtained with a Spectra Physics Hurrican X laser system (800 nm, 100 fs, 1k Hz). Samples were excited at λex of 267 nm by the frequency-tripled Ti sapphire laser line.

3. Results and discussion

From structure point of view, CASO has both layered and framework structure. In the layered structure, the Ca2+ ions are located in the interlayers of the double tetrahedral layer and are expected to exchange readily through the interlayers. According to the Wyckoff, CASO structure contains four crystallographically different Ca sites at the same Wyckoff position, 4i (Fig. 3). Due to the charge balance, all the four sites are not full occupied and induced the disorder of bond angle and length. We can divide the four Ca2+ sites into two groups in accord to their different number of folders. Namely, one type of Ca2+ ion occupies an octahedral site with six oxygen atoms and the average Ca–O bond distance is 2.485 Å. Other Ca2+ ions occupy three kinds of polyhedral sites with seven coordinated oxygen atoms and their average bond distances are 2.508, 2.531, and 2.562 Å, respectively (Fig. 4). Therefore, it is expected that the Ce3+ ions doped in this host will exhibit two kinds of luminescence properties according to their different energy level. Al and Si atoms both occupy tetrahedral sites with four coordinated oxygen atoms, and the average bond distances for Al–O and Si–O are 1.735 and 1.611 Å, respectively. On the basis of steric considerations, it is clear that Ce3+ cations will occupy exclusively alkaline-earth crystallographic sites like Eu2+ cations in Eu2+ MASO derivatives (M = Ca, Sr, Ba). In the same way, the ionic radii of Mn2+ cation in the different polyhedra tend to suppose their presence in the alkaline-earth sites. This assumption was a posteriori verified by Yang and co-workers9 and their investigations of optical properties of Mn2+-doped CASO phases.
image file: c3ra47156j-f3.tif
Fig. 3 CASO structure represented in cell along the c axis.

image file: c3ra47156j-f4.tif
Fig. 4 Calcium environments and bond length in CASO (Wyck.: 4i).

3.1. Optical properties of C1−3x/2ASO:xCe3+

The PLE and PL spectra of CASO:2%Ce3+ are displayed in Fig. 5. PLE spectra are composed of two broad bands peaking at 250 and 294 nm for CASO (λem = 356 nm). Without doubt, these absorption bands are associated with 4f–5d intra-atomic electronic transitions of Ce3+, the d-block being split as usual by the ligand field. Since the optical gap of the CASO host lattice is estimated around 4 eV (310 nm),18 it can be speculated that the 4f-block of Ce3+ lies at the top of the valence band. Formally, on the basis of optical and XPS investigations carried out on CaAl2Si2O8:Eu,18 the Eu-d block is expected to lie just below the conduction band and the 4f7 block of Eu2+ cations. Thus, as the position in energy of d-orbitals is not expected to significantly change within the lanthanide series26 this leads us to conclude on the aforementioned positioning of the 4f7 block of Ce3+.
image file: c3ra47156j-f5.tif
Fig. 5 PLE and PL spectra for phosphors with composition of CASO:2%Ce3+.

Emission spectra were monitored at both excitation wavelengths, i.e. 294 and 250 nm. No significant evolution in the shape of PL was noticed vs. excitation wavelengths. The PL spectra of the CASO:Ce samples exhibit an intense band at 356 nm with a shoulder at 334 nm and a much broader band around 430 nm. The two emission bands placed at low wavelength can be assigned to Ce3+ cations occupying the Ca2+ sites. Formally, in sake of clarity, PL spectra of CASO:9%Ce sample has been deconvoluted by Gaussian functions taking into account the ∼0.25 eV energy separation between the 5d → 2F5/2 and 5d → 2F7/2 transitions due to spin-orbit coupling.14 This clearly leads for CASO:Ce compound to distinguish the contributions of six-fold and seven-folded coordinated Ce3+ cations (Fig. 6), the former giving rise to emissions at 352 and 379 nm, the latter to emissions at 327 and 355 nm. This assignment, based only on the widespread postulate that the higher the coordination of Ce3+ cations, the lower the crystal field splitting and the shorter the emission wavelength, must be moderated by the existence of counter-examples.27,28


image file: c3ra47156j-f6.tif
Fig. 6 Emission spectra of CASO:9%Ce3+ with decomposition in Gaussian curves under λex = 294 nm.

If the assignment of Ce3+ emission bands to the luminescence spectrum of Ce doped CASO compound is elucidated, there still remains an unexpected broad peak (hereafter called “defect” peak) at 430 nm for CASO. To shed light on the origin of this unassigned peak, PLE spectra were monitored at these two wavelengths (Fig. 7). Their examination clearly shows that they differ from the PLE of Ce3+. At very first sight, based on previous investigations carried out on Eu activated CASO phosphor, emissions at 430 nm might be attributed to 4f65d1 → 4f7(8S7/2) transitions of Eu2+ possibly present as undesired traces since this activator gives rise to luminescence bands peaking around 440 nm for CASO host lattices.18 To address this issue, samples were prepared in new alumina crucibles with fresh precursors. They systematically led to materials with the same optical characteristics, i.e. a broad “defect” emission band in addition to those of Ce3+. The determination of the decay times for this band in CASO was then initiated and revealed lifetimes of about 22 ns for CASO:Ce samples which turns to be very different to that commonly expected for Eu2+ (ref. 29) and that collected for CASO:Eu materials by Yang et al. (i.e. 60–70 μs).9 Consequently, the assignment of these unforeseen bands to the presence of Eu2+ luminescent centers can be unambiguously eliminated. This is in good agreement with the observation that room temperature excitation spectra of CASO:Ce3+ at 430 nm differ from those of CASO:Eu2+ at 440.


image file: c3ra47156j-f7.tif
Fig. 7 PLE spectra for CASO:2%Ce3+ monitored at 356, 430 nm.

To explain the presence of these broad bands in the absence of Eu2+ cations, one hypothesis can be proposed. Namely, the substitution of Ce3+ ions for Ca2+ alkaline-earth specie must be combined with the creation of vacancies according to the formal equation 3Ca2+ → 2Ce3+ + VCa to respect the charge balance. Thus, “defect” emissions would be correlated with the presence of these alkaline-earth vacancies (probably distributed in close proximity of image file: c3ra47156j-t2.tif for electrostatic reasons). This would explain why the intensity of such bands increases with cerium concentration as observed experimentally (Fig. 8).


image file: c3ra47156j-f8.tif
Fig. 8 Evolution PL of CASO:Ce3+ and overlap between Ce3+ and ‘defects’ (insert).

To test this hypothesis, investigations were undertaken on CASO:La3+, CASO:La3+, K+ and CASO:Ce3+, K+ materials (prepared with K2CO3 (99.99% Alfa Aesar) and La2O3 (99.9% Alfa Chempur) precursors), as well as CASO host lattices themselves. Fig. 9 displays the PL of CASO, CASO:5%Ce, CASO:5%La, CASO:5%Ce, 5%K and CASO:5%La, 5%K monitored at 294 nm. First, undoped CaAl2Si2O8 host lattices do not luminesce at all. Second, CASO:La3+ materials exhibit only the defect emissions, i.e. similar spectra than those of CASO:Ce3+ missing the 5d → 4f contributions. Third, the intensities of the defect bands strongly decrease when K+ ions are co-inserted in the host lattices with lanthanide cations in equimolar concentration. Namely, the alkali metal plays the role of charge compensator according to the image file: c3ra47156j-t3.tif formal reaction that diminishes the amount of alkaline earth vacancies naturally created under La3+ substitution. On the basis of the examination of emission spectra of CASO:Ce and coupled with decay time measurements, are clearly indicative that the emission bands at 430 nm in Ce doped CASO are related to alkaline earth vacancies and not to the presence of undesired Eu2+ cations. CASO:Ce, K, it is worth noting that the intensities of the 5d → 2F5/2 and 5d → 2F7/2 Ce contributions are not affected by the insertion of K+ and that the “defect” bands cannot be totally annihilated by the insertion of K+ cations suggesting that VCa vacancies still exist. Thus, we can conclude that Ce3+ (La3+) cations are unfortunately systematically introduced in larger amount than K+ cations. Regardless, the three aforemention observations, coupled with decay time measurements, are clearly indicative that the emission bands at 430 nm in Ce doped CASO is related to alkaline earth vacancies and not to the presence of undesired Eu2+ cations.


image file: c3ra47156j-f9.tif
Fig. 9 PL spectra monitored at 294 nm for CASO HL and Ce3+, La3+, Ce3+/K+ and La3+/K+ co-doped materials.

The dependence of the PL of Ca1−3x/2ASO:xCe3+ materials (x lower than 16%) are depicted in Fig. 8. The emissions associated with Ce3+ lines increase up to quenching concentrations of about 9% for CASO. Beyond these limits, the intensity of 5d → 2F5/2 and 5d → 2F7/2 lines decreases whereas the one of the “defect” bands continues to slightly increase and to tend towards a plateau. This trend maybe explained from the fact that (i) the higher the Ce3+ concentration the higher the Ca2+ vacancies, (ii) the existence of possible energy transfers from Ce3+ to VCa as suggested by the spectral overlap between the Ce3+ PL spectra and the “defect” PLE spectra in CASO:Ce3+ materials (the inset in Fig. 8), and (iii) a quenching phenomenon specific to defects themselves.

3.2. Optical properties of C1−yASO:yMn2+

The PL and PLE spectra for purely Mn2+ activated CASO host lattices were also recorded (Fig. 10). They tend to be very similar to those reported in the literature.9,21 In brief, excitation spectra consist for both host matrix of several bands in the UV and visible regions corresponding to the 4A1(6S) to 4E(4D), 4T2(4D), [4A1(4G), 4E(4G)], 4T2(4G), and 4T1(4G) transition levels, respectively. Concerning the emission spectra, the main peak is located at 563 nm for CASO. Due to the photoluminescence property of Mn2+ ions and the low dopant concentration in the CASO matrix, it is difficult to observe the PL and PLE spectra for the C0.995ASO:0.5%Mn2+ sample. However, excitation and emission can be observed for higher dopant rate. Mn2+ ion in solid materials is usually characterized by a green emission in tetrahedral coordination and by a yellow-red emission in octahedral coordination.10,21 Therefore, the emission at 568 nm also suggests that Mn2+ cations are located at the alkaline-earth site instead of being at the Al or Si sites as already proposed by for CASO9 which was expected on the basis of steric considerations. As the d–d transitions of Mn2+ ion are difficult to pump, the transitions energy between the ground state and excited sub-levels states are weak. Thus, we can suppose that the energy differences between the Mn1 (occupying the Ca1 site) and the other sites Mn2,3,4 (occupying the Ca2,3,4 sites) are also weak and thus not easy to distinguish. Under this consideration, we cannot give proper assignments to the emission and excitation spectra of Mn2+. Furthermore, the large Stocks shift (0.9 eV) of Mn2+ centers in CASO matrix indicates that a strong electron–lattice interaction occurs inducing a bad resolution of emission band of Mn2+.
image file: c3ra47156j-f10.tif
Fig. 10 PL and PLE spectra for purely Mn2+ activated CASO.

3.3. Overlap of the PL of CASO:Ce and PLE of CASO:Mn

The comparison of the PL spectra for CASO:Ce3+ and the PLE spectra for CASO:Mn2+ reveals a heavy spectral overlap (Fig. 11). Thus, energy transfers from Ce3+ to Mn2+ can be anticipated as already exemplified in many phosphors including Ca8MgLa(PO4)7,30 Ca3Sc2Si3O12,31 SrAl2O4,32 BaAl2O4,33 Ba2Ca(BO3)2,8 Ca9Y(PO4)7 (ref. 10) or LiCaBO3.34 Nevertheless, in contrasts to these materials, transfers could also take place in the case of CASO from alkaline-earth-vacancy based defects to Mn2+ and from Ce3+ to Mn2+ via VCa according to the proposed Fig. 12.
image file: c3ra47156j-f11.tif
Fig. 11 Spectral overlap between the photoluminescence spectrum of C0.865ASO:9%Ce3+ (solid line) and photoluminescence excitation spectrum of C0.91ASO:9%Mn2+.

image file: c3ra47156j-f12.tif
Fig. 12 Possible energy transfers in CASO:Ce, Mn compounds.

Moreover, Fig. 13 gathers the PLE of CASO:9%Ce3+ monitored at 353 nm (Ce3+ emission), and CASO:9%Ce3+, 9%Mn2+ monitored at 353 nm and 568 nm (Mn2+ emission). It can be observed for CASO:Ce, Mn that photoluminescence excitation spectrum monitored by the Mn2+ emission (λem = 568 nm, green line) has a profile very similar to that of the excitation band monitored by the Ce3+ emission (λem = 353 nm, red line). This reveals that the Mn2+ ions can be essentially excited in the same energy range than Ce3+ ions. This observation, coupled with the strong decrease of the PLE going from CASO:9%Ce (black line) to CASO:9%Ce, 9%Mn (red line) confirms also that an ET occurs from image file: c3ra47156j-t4.tif to Mn2+. The photoluminescence excitation spectra of the Mn2+ ions in the Ce3+ and Mn2+ co-activated sample extend from 230 to 340 nm, indicating that this phosphor could be, in principle, used in a UV white-light LED.


image file: c3ra47156j-f13.tif
Fig. 13 Photoluminescence excitation of CASO:9%Ce3+ (black line), CASO:9%Ce3+, 9%Mn2+ (red line) under λem = 353 nm and CASO:9%Ce3+, 9%Mn2+ (green line) under λem = 568 nm.

3.4. Optical properties of co-doped CASO:Ce, Mn

Emission spectra of Ce3+/Mn2+ co-doped CASO is reported in Fig. 14. Ce3+ concentration has been fixed at 9% for CASO, namely, at the quenching concentration to favour energy transfers. On PL spectra, it can be observed that the insertion of Mn2+ in CASO:Ce compounds triggers a decline of the intensity of the emission bands associated with both Ce3+ and “defects” concomitant with an enhancement of the Mn2+ emission. The decrease of the Ce3+ emission is regular and continuous with Mn2+ concentration, whereas the Mn2+ emission progressively increases to a Mn insertion rate of about 9% before decreasing due to concentration quenching.
image file: c3ra47156j-f14.tif
Fig. 14 Photoluminescence spectra for C0.865−yASO:9%Ce3+, yMn2+ phosphors versus the Mn2+ doping content (y), under λex = 294 nm.

To quantify the optical characteristics of phosphors and to judge the impact of Mn2+ co-doping on the generated color, the CIE(x, y) parameters of prepared materials have to be determined. These values are reported below as shown in Tables 3 and 4 for CASO:Ce, CASO:Ce, Mn, respectively. All data were collected with an excitation wavelength of 294 nm.

Table 3 CIE chromaticity coordinates for phosphors C1−3x/2ASO:xCe3+ (λex = 294 nm)
C1−3x/2ASO:xCe3+ CIE(x, y)
x = 0.005 (0.1976, 0.1564)
x = 0.02 (0.2162, 0.1791)
x = 0.05 (0.2148, 0.1876)
x = 0.09 (point 1 in Fig. 15) (0.2270, 0.2212)
x = 0.15 (0.2022, 0.1868)


Table 4 CIE chromaticity coordinates for phosphors C0.865−yASO:0.09Ce3+, yMn2+ (0.01 < y < 0.15)
C0.865−yASO:0.09Ce3+, yMn2+ CIE(x, y)
x = 0.01 (0.2968, 0.3142)
x = 0.015 (0.3176, 0.3242)
x = 0.02 (0.3416, 0.3342)
x = 0.03 (0.3577, 0.4016)
x = 0.06 (0.3982, 0.4535)
x = 0.09 (0.4272, 0.4794)
x = 0.12 (0.2022, 0.1868)
x = 0.15 (0.4551, 0.4797)


For the C0.865−yASO:9%Ce3+, yMn2+ (y = 0 to 15%) series, a regular shift from blue for y = 0% to white (y = 1, 1.5 and 2%) and to yellow (y > 3%) is clearly evidenced (Fig. 15 and Tables 3 and 4). This corresponds to a change from (0.2270, 0.2212) to (0.4551, 0.4797) of the chromatic coordinates for Mn2+ concentrations increasing from 0 to 15%. This clearly evidences the role of ET from Ce (and image file: c3ra47156j-t5.tif defect) towards Mn2+ cations in these materials and the importance to control the Mn concentration to generate a white color. The white color would be obtained for a Ca0.845Ce0.090Mn0.02Al2Si2O8 composition, this composition being probably not optimal.


image file: c3ra47156j-f15.tif
Fig. 15 Evolution of CIE chromaticity coordinates for C0.865−yASO:0.09Ce3+, yMn2+ excited at 294 nm. The numbers from 1–9 correspond to: 1, y = 0; 2, y = 0.01; 3, y = 0.015; 4, y = 0.02; 5, y = 0.03; 6, y = 0.6; 7, y = 0.09; 8, y = 0.12; 9, y = 0.15.

3.5. Mechanism of the ET from Ce3+ image file: c3ra47156j-t6.tif to Mn2+

At present, we will examine the efficiency (ηT) of the ET from image file: c3ra47156j-t7.tif to Mn2+. Generally, the ET efficiency from a sensitizer to activator can be expressed by the following eqn (1):10,19,33
 
image file: c3ra47156j-t8.tif(1)
where ηT is the ET efficiency and I0 and IS are the luminescence intensity of a sensitizer in the absence and presence of an activator, respectively. In our case, IS0 is the intrinsic emission intensity of the sensitizer (Ce3+) and IS is the emission intensity of the sensitizer (Ce3+) in the presence of the activator (Mn2+) for identical cerium concentrations. Fig. 16 shows the results of ET efficiency from Ce3+ to Mn2+ calculated by the previous eqn (1) for the C0.865−yASO:9%Ce3+, yMn2+ (y = 1 to 15%) samples. As shown in Fig. 16, the ET efficiency increases with increasing Mn2+ concentration. However, the slope of the emission intensity gradually decreases when Mn2+ concentration increases. This indicates that the quantity of energy that can transfer from Ce3+ to Mn2+ is gradually restricted for high Mn2+ concentration due to: (1) the fixed concentration of Ce3+ and (2) the concentration quenching of Mn2+ to a lesser extent.

image file: c3ra47156j-f16.tif
Fig. 16 Dependence of the energy transfer efficiency ηT on the Mn2+ content (y) for the C0.865−yASO:9%Ce3+, yMn2+ samples under λex = 294 nm.

Normally, two types of resonant ET mechanism exist. One corresponds to exchange interactions and the other to multipolar interactions.10,14,19 It is well-known that if ET is based on exchange interaction, the critical distance between the sensitizer and activator should be shorter than about 5 Å. We previously saw that the critical distance can be calculated from the quenching concentration. From Fig. 16, the critical concentration can be determined to be about 0.161 (= 0.09 (Ce3+ concentration) + 0.071 (Mn2+ concentration)) from the total concentration of Ce3+ and Mn2+ for which the ET efficiency is 0.5. Therefore, the critical distance RC (Ce3+–Mn2+) was calculated to be ∼15.21 Å by using the eqn (2),10,11,13,14 where V is the unit-cell volume of the sample, N is the number of formula units per unit cell and XC is the critical concentration. This value is too large to lead to exchange interaction via orbital overlaps. Consequently, the electric multipolar interaction must be considered for ET between Ce3+ and Mn2+ ions in the CASO matrix.

 
image file: c3ra47156j-t9.tif(2)

On the basis of the Dexter's ET formula of multipolar interaction and Reisfeld's approximation, the following eqn (3) can be obtained:9–11,21

 
image file: c3ra47156j-t10.tif(3)
where IS0 and IS have the same meaning as above. C is the concentration of Mn2+. α equal to 6, 8, 10 for dipole–dipole, dipole–quadrupole, quadrupole–quadrupole interactions, respectively. Plots of IS0/IS based on the above equation are shown in Fig. 17. As one can see, linear behavior is only observed when α = 6, which implies that ET from Ce3+ to Mn2+ occurred via a dipole–dipole mechanism. This is quite surprising because a dipole–quadripole interactions was expected however this was already observed in the literature for several series of materials.35


image file: c3ra47156j-f17.tif
Fig. 17 Dependence of IS0/IS of Ce3+ versus α = 6, 8 and 10 for the samples C0.865−yASO:9%Ce3+, yMn2+.

To confirm these results based on intensity measurements only, the decay curves of the C0.865−yASO:9%Ce3+, yMn2+ (y = 0 to 15%) samples were also collected. These ones are depicted in Fig. 18 while decay times are reported in Table 5. For the first three samples, the decay curve was best fitted with a monoexponential function whereas for others samples the decays curves were fitted with a biexponential function.


image file: c3ra47156j-f18.tif
Fig. 18 Decay curve obtained under excitation λex = 267 nm and emission λem = 360 nm (centered at the maximum emission of Ce3+ in the CASO matrix). Dotted lines represent the experimental value and solid lines represent the fitting value.
Table 5 The results of the decay times for the samples C0.865−yASO:9%Ce3+, yMn2+ (y = 0 to 15%) (λem = 360 nm)
Phosphors A1 A2 τ1 (ns) τ2 (ns) τ〉 (ns)
Ce9Mn0 23.524 23.524
Ce9Mn1 22[thin space (1/6-em)]665 22.665
Ce9Mn3 20[thin space (1/6-em)]280 20.280
Ce9Mn6 8.1 91.9 0.682 19[thin space (1/6-em)]654 18.117
Ce9Mn9 7.0 93.0 3.067 17.963 16.920
Ce9Mn12 12.8 87.2 0.663 17.268 15.143
Ce9Mn15 17.5 82.5 4.697 17.803 12.073


From Table 5 and Fig. 18, one can also notice that the meantime value 〈τ〉 decreases when the concentration of Mn increases. Thus we can further confirm that the ET from Ce3+ to Mn2+ occurs in the CASO matrix. Furthermore, we can also conclude that the ET is a resonant non-radiative transfer because the decay times of the Ce3+ decreases with the Mn2+ ion concentration. The plots of τS0/τS that are calculated from the previous eqn (3) are reported in Fig. 19. This plot also proves that the dipole–dipole interactions are at work (α = 6) even if dipole–quadripole interactions cannot be now totally rebutted (α = 8).


image file: c3ra47156j-f19.tif
Fig. 19 Dependence of τS0/τS of Ce3+ on α = 6, 8 and 10 for the samples C0.865−yASO:9%Ce3+, yMn2+.

As mentioned above, an ET also occurs between the defects emission image file: c3ra47156j-t11.tif and the Mn2+ ions in the CASO matrix. The ET efficiency from image file: c3ra47156j-t12.tif to Mn2+ ions in the samples C0.865−yASO:9%Ce3+, yMn2+ (y = 1 to 15%) phosphors were also calculated and the results are presented in Fig. 20. As one can see, the ET efficiency increases gradually with the increase of Mn2+ concentration. The rate of increase of the emission intensity dramatically decreases with increasing Mn2+ concentration and the efficiency in the image file: c3ra47156j-t13.tif transfer is higher than in the Ce3+–Mn2+ when compared to the ET from Ce3+ to Mn2+ shown in Fig. 16. These phenomena in the image file: c3ra47156j-t14.tif maybe due to a better overlap between the emission of the image file: c3ra47156j-t15.tif (∼420 nm) with the excitation of the Mn2+ (∼403 nm) compared to the emission of Ce3+ at 356 nm.


image file: c3ra47156j-f20.tif
Fig. 20 Dependence of the energy transfer efficiency ηT from image file: c3ra47156j-t16.tif to Mn2+ (on the Mn2+ content (y)).

The critical concentration image file: c3ra47156j-t17.tif for energy transfer from image file: c3ra47156j-t18.tif to Mn2+ ions was also calculated and estimated to be around 17 Å for a image file: c3ra47156j-t19.tif concentration of ∼0.045 and a quenching concentration for Mn2+ of 2.43%. In the same way as previously described, the mechanism of multipolar interactions can be expected. Fig. 21 was used to determine the intensities of the emissions from the defects image file: c3ra47156j-t20.tif for samples C0.865−yASO:9%Ce3+, yMn2+. For those samples, the Ce3+ concentration is fixed and consequently, the content of defects also image file: c3ra47156j-t21.tif. As one can see in Fig. 21, a linear behavior is only observed when α is equal to 6, meaning that a dipole–dipole interactions also occurs between vacancies image file: c3ra47156j-t22.tif and Mn2+ ions.


image file: c3ra47156j-f21.tif
Fig. 21 Dependence of IS0/IS of defects emission image file: c3ra47156j-t23.tif. versus α = 6, 8 and 10 (fixed Ce3+ (thus, image file: c3ra47156j-t24.tif), change the concentration of Mn2+).

4. Conclusions

In summary, we have synthesized and investigated the luminescent properties of CaAl2Si2O8 phosphors (co)-activated with Ce3+ and Mn2+. The results have shown that the substitution of the alkaline-earth ions by Ce3+ (and La3+) implies the systematic creation of cationic vacancies even when alkali metals are introduced as counter-cations to prevent their formation. These vacancies strongly impact the photoluminescence, leading to broad emission bands at 430 nm in CaAl2Si2O8 host lattice. Their intensities are naturally reinforced by enhancing their concentration when Ce doping increases (at least up to 16%Ce) but also by energy transfer mechanism from Ce3+ cations. Furthermore, Ce doped CASO compounds give rise to a cold blueish white luminescence. Changes in the spectral characteristics can be initiated by the use of Mn2+ cations as co-dopants. Then, due to energy transfers towards Mn2+ centres, emission shifts continuously from blue-white, to white, and even yellow-white vs. Ce and Mn concentrations.

Acknowledgements

The author acknowledges S. Jobic, P. Deniard, H. Brault and F. Massuyeau for the technical assistance and fruitful discussion in IMN, Nantes University, CNRS, FRANCE. This work has been supported by the Chinese Scholarship Council (CSC), no. 2009615018.

Notes and references

  1. C. You, Opt. Eng., 2005, 44, 111307 CrossRef PubMed.
  2. T.-S. Chan, R.-S. Liu and I. Baginskiy, Chem. Mater., 2008, 20, 1215–1217 CrossRef CAS.
  3. C. Feldman, T. Justle, C. R. Ronda and P. J. Schmidt, Adv. Funct. Mater., 2003, 13, 511 CrossRef.
  4. H. A. Hoppe, Angew. Chem., Int. Ed. Engl., 2009, 48, 3572–3582 CrossRef PubMed.
  5. R.-J. Xie, N. Hirosaki, T. Suehiro, F.-F. Xu and M. Mitomo, Chem. Mater., 2006, 18, 5578–5583 CrossRef CAS.
  6. N. Cheung and T. Y. Wong, Surv. Ophthalmol., 2007, 52, 180–195 CrossRef PubMed.
  7. D. A. Steigerwald, J. C. Bhat, D. Collins, R. M. Fletcher, M. O. Holcomb, M. J. Ludowise, P. S. Martin and S. L. Rudaz, IEEE J. Sel. Top. Quantum Electron., 2002, 8, 310–320 CrossRef CAS.
  8. C. Guo, L. Luan, Y. Xu, F. Gao and L. Liang, J. Electrochem. Soc., 2008, 155, J310–J314 CrossRef CAS PubMed.
  9. W.-J. Yang, L. Luo, T.-M. Chen and N.-S. Wang, Chem. Mater., 2005, 17, 3883–3888 CrossRef CAS.
  10. C.-H. Huang, T.-W. Kuo and T.-M. Chen, ACS Appl. Mater. Interfaces, 2010, 2, 1395–1399 CAS.
  11. N. Guo, Y. Huang, H. You, M. Yang, Y. Song, K. Liu and Y. Zheng, Inorg. Chem., 2010, 49, 10907–10913 CrossRef CAS PubMed.
  12. S. Ye, F. Xiao, Y. X. Pan, Y. Y. Ma and Q. Y. Zhang, Mater. Sci. Eng., R, 2010, 71, 1–34 CrossRef PubMed.
  13. G. Li, D. Geng, M. Shang, C. Peng, Z. Cheng and J. Lin, J. Mater. Chem., 2011, 21, 13334–13344 RSC.
  14. G. Bless and B. C. Grabmaire, Luminescent Materials, Springer Verlag, Berlin, 1994 Search PubMed.
  15. C. Ronda, Luminescence: From Theory to Applications, Willey-VCH Verlag GmbH & Co., Weinheim, 2008 Search PubMed.
  16. S. C. Gedam, S. J. Dhoble and S. V. Moharil, J. Lumin., 2007, 124, 120–126 CrossRef CAS PubMed.
  17. S. H. M. Poort, A. Meijerink and G. Blasse, Solid State Commun., 1997, 103, 537–540 CrossRef CAS.
  18. F. Clabau, A. Garcia, P. Bonville, D. Gonbeau, T. Le Mercier, P. Deniard and S. Jobic, J. Solid State Chem., 2008, 181, 1456–1461 CrossRef CAS PubMed.
  19. J. K. Park, J. M. Kim, E. S. Oh and C. H. Kim, Electrochem. Solid-State Lett., 2005, 8, H6–H8 CrossRef CAS PubMed.
  20. C. Zhang, J. Yang, C. Lin, C. Li and J. Lin, J. Solid State Chem., 2009, 182, 1673–1678 CrossRef CAS PubMed.
  21. S. Ye, Z.-S. Liu, X.-T. Wang, J.-G. Wang, L.-X. Wang and X.-P. Jing, J. Lumin., 2009, 129, 50–54 CrossRef CAS PubMed.
  22. W. Z. Gebert, Kristallografiya, 1972, 135, 437 CAS.
  23. V. D. petricek and L. Palatinus, The Crystallographic Computing System JANA, 2006 Beta, Academy of Sciences, Praha, Czeck Republic, 2006 Search PubMed.
  24. R. D. Shannan, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751 CrossRef.
  25. J. Wagberg, Matlab Toolbox for Calculation of Color Related Optical Properties-version 2.1, More Research and DPC Digital Printing Center, 2007 Search PubMed.
  26. P. Dorenbos, J. Lumin., 2003, 105, 117–119 CrossRef CAS.
  27. G. Denis, P. Deniard, E. Gautron, F. Clabau, A. Garcia and S. Jobic, Inorg. Chem., 2008, 47, 4226–4235 CrossRef CAS PubMed.
  28. G. Denis, X. Rocquefelte, P. Deniard, M.-H. Whangbo and S. Jobic, J. Mater. Chem., 2009, 19, 9170–9175 RSC.
  29. C. Fouassier, Luminescence in Encyclopedia of Inorganic Chemistry, New York, J. Wiley, 2nd edn, 2005 Search PubMed.
  30. Y. N. Xue, F. Xiao and Q. Y. Zhang, Spectrochim. Acta, Part A, 2011, 78, 1445–1448 CrossRef CAS PubMed.
  31. Y. Liu, X. Zhang, Z. Hao, Y. Luo, X. Wang, L. Ma and J. Zhang, J. Lumin., 2013, 133, 21–24 CrossRef CAS PubMed.
  32. X. Xu, Y. Wang, X. Yu, Y. Li and Y. Gong, J. Am. Ceram. Soc., 2011, 94, 160 CrossRef CAS.
  33. N. Suriyamurthy and B. S. Panigrahi, J. Lumin., 2007, 127, 483–488 CrossRef CAS PubMed.
  34. C. Guo, J. Yu, X. Ding, M. Li, Z. Ren and J. Bai, J. Electrochem. Soc., 2011, 158, J42–J46 CrossRef CAS PubMed.
  35. K. H. Kwon, W. B. Im, H. S. Jang, H. S. Yoo and D. Y. Jeon, Inorg. Chem., 2009, 48, 11525–11532 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.