• Aucun résultat trouvé

Spectroscopic Study of a Single Crystal of SrAl

N/A
N/A
Protected

Academic year: 2022

Partager "Spectroscopic Study of a Single Crystal of SrAl"

Copied!
12
0
0

Texte intégral

(1)

Article

Reference

Spectroscopic Study of a Single Crystal of SrAl

2

O

4

:Eu

2+

:Dy

3+

DELGADO PEREZ, Teresa, AFSHANI AGHAJARI, Jafar, HAGEMANN, Hans-Rudolf

Abstract

The extraordinary long phosphorescence of SrAl2O4:Eu2+:Dy3+ has been widely studied in powder samples because of its broad range of applications in this form and despite the fact that the bulk material shows higher intensity emission and longer afterglow. However, the investigation of SrAl2O4:Eu2+:Dy3+ crystals that, unlike the powder, do not contain surface defects, would allow a better insight into the mechanism that governs the long-lasting phosphorescence of this co-doped material. Thus, a SrAl2O4:Eu2+:Dy3+ single crystal was studied in detail by absorption spectroscopy and photoluminescence, including a novel estimation of its extinction coefficient. In addition, thermoluminescence measurements and wavelength-dependent quantum efficiency measurements have been performed to improve the understanding of the role of both europium and dysprosium ions in the corresponding persistent phosphorescence mechanism.

DELGADO PEREZ, Teresa, AFSHANI AGHAJARI, Jafar, HAGEMANN, Hans-Rudolf.

Spectroscopic Study of a Single Crystal of SrAl2O4:Eu2+:Dy3+. Journal of Physical Chemistry.

C, 2019, vol. 123, no. 14, p. 8607-8613

DOI : 10.1021/acs.jpcc.8b12568

Available at:

http://archive-ouverte.unige.ch/unige:147125

Disclaimer: layout of this document may differ from the published version.

1 / 1

(2)

1 Accepted paper

This paper was published in the Journal of Physical Chemistry C 2019, 123, 14, 8607–8613 https://doi.org/10.1021/acs.jpcc.8b12568

Spectroscopic study of a single crystal of SrAl

2

O

4

:Eu

2+

:Dy

3+

Teresa Delgadoa, Jafar Afshania, Hans Hagemanna

aDépartement de Chimie Physique, Université de Genève, Quai Ernest-Ansermet 30, Genève 1211, Switzerland.

Abstract :

The extraordinary long phosphorescence of SrAl2O4:Eu2+:Dy3+ has been widely studied in powder samples due to its broad range of applications in this form and despite the fact that the bulk material shows higher intensity emission and longer afterglow. However, the investigation of SrAl2O4:Eu2+:Dy3+

crystals that, unlike the powder do not contain surface defects, would allow a better insight into the mechanism that governs the long-lasting phosphorescence of this co-doped material. Thus, a SrAl2O4:Eu2+:Dy3+ single crystal was studied in detail by absorption spectroscopy and photoluminescence, including a novel estimation of its extinction coefficient. In addition, thermoluminescence measurements and wavelength dependent quantum efficiencies measurements have been performed to improve the understanding of the role of both europium and dysprosium ions in the corresponding persistent phosphorescence mechanism.

INTRODUCTION

The alkaline earth aluminate SrAl2O4:Eu2+:Dy3+ exhibits one of the longest afterglows among the broad range of inorganic persistent phosphors,1 which makes it an excellent candidate for a wide range of applications such as safety emergency signs,2 photocatalysis,3 solar cells,4 etc. The host structure of SrAl2O4:Eu2+:Dy3+ is well characterised. The SrAl2O4 crystalizes in a monoclinic system and consist of rings formed by six corner sharing oxide aluminate tetrahedra which form channels within which the Sr2+ ions are situated. There are two non-equivalent Sr2+ sites with the same coordination number and Sr-O distances but they differ by a slight distortion of their square planes.5 In consequence, when the Sr2+ ions (r = 126 pm in VIII coordination ) are replaced by Eu2+ (r = 125 pm)6, the luminescence properties of the two different Eu2+sites slightly differ too and two different emissions are observed at around 450 nm and 520 nm at low temperatures.7-9 Contrary, at ambient temperature the blue emission is thermally quenched. In addition, energy transfer of the excited Eu2+ from the blue site to the lower energy Eu2+ green site takes place.7,10 It is worthy to mention that the SrAl2O4:Eu2+ system itself shows a significant long phosphorescence. However, when another co-dopant such as Dy3+(r = 97 pm) is incorporated into the host by replacing the Sr2+ ions a longer-lasting afterglow is observed which lasts up to several hours after the excitation source is removed.11-13

Several mechanisms have been proposed in order to explain the later phenomena. In the very well- known “hole trapping mechanism” elucidated by Matsuzawa et al.,14 after excitation, the Eu2+ ion takes an electron from the valence band (VB) of the host and the Dy3+ traps the hole left in the VB, being reduced to Dy4+. The delayed radiative emission is produced when by thermal activation one electron is promoted from the top of the VB to the Dy4+ ion which is reduced back to Dy3+, and the trapped hole returns to the Eu1+ site. However, the low probability of the “hole trapping process” given the instability

(3)

2

of Dy4+ and Eu1+ led other researchers to propose new alternatives. Dorenbos et al.15 raised the so- called “electron-trapping mechanism” based on the excitation of one electron of Eu2+ into the conduction band (CB) of the host to generate Eu3+ and the trapping of this electron by the Dy3+ with its consequent reduction to Dy2+. In this case the electron trap can be thermally released and the recombination with the Eu3+ gives rise to the delayed luminescence. Nevertheless, none of these mechanisms explained the intrinsic phosphorescence of the Dy3+ free material SrAl2O4:Eu2+. Thus, other models based on oxygen vacancies have been proposed by Clabau et al.,16 in which the electron that is excited from the Eu2+ into the CB is trapped by an oxygen vacancy instead. In addition, if there is residual Eu3+ not reduced during the synthesis of the doped aluminate, under irradiation, one electron from the VB can be promoted to the Eu3+ cations and the subsequent recombination with the VB would give rise to the luminescence at 450 nm. Recently, ab initio calculations have been used to assign the green luminescence of Eu2+ to the Sr2 site17 and calculations based on density functional theory have underlined the importance of the intrinsic defects of the host in the long luminescence mechanism.18 Nonetheless, although persistent phosphorescence is known from centuries ago,19 and despite the effort to prove experimentally the above described mechanisms, the existence of traps with diverse origins and the presence of traps distributions rather than single traps enhance the difficulty to expose a single valid mechanism.

Herein, the properties of a single crystal of SrAl2O4:Eu2+:Dy3+ have been studied in order to contribute to the development of a mechanism that could explain the role of co-dopans in the long-lasting phosphorescence of this kind of materials. SrAl2O4:Eu2+:Dy3+ single crystal fibers were synthesized for the first time by Jia et al. by laser heated pedestral growth taking advantage of the low vapor preassure of the alkaline earth aluminates at their melting point.20,21 At that time, in 1998, just after the Matsuzawa mechanism was raised, the new crystal fibers were preliminary characterised by photo and thermoluminescence. Since then few studies have been focused on the investigation of SrAl2O4:Eu2+:Dy3+ crystals22,23 but rather powders despite their lower intensity and shorter afterglow, since they can be more easily integrated in pigments or surfaces for applications,2 especially in the case of nanoparticles.24 However, the use of nano or microcrystalline powder implies that not only bulk defects are present in the material but also surface defects. Thus, in order to study the mechanism responsible for the long phosphorescence, only thermodynamically relatively stable defects should be present. This is the case of a single crystal whose purity is ensured by the slow growth process used for its synthesis.

Our SrAl2O4:Eu2+:Dy3+ single crystal characterization includes the estimation of its extinction coefficient by analysing its optical properties and the real concentration of Eu2+ within it. The value of the extinction coefficient is up to our knowledge the first one proposed for this kind of material and could serve as a starting point to calculate the real concentration of Eu2+ in any sample of the same composition. In addition, thermoluminescence analysis have been performed in order to study the effect of the crystallinity of these samples in the quality of the traps. Besides, wavelength dependence quantum efficiency measurements have been performed on the SrAl2O4:Eu2+:Dy3+ crystal and compared with a Dy3+ free sample so that the contribution of both Dy3+ and Eu2+ individually to the phosphorescence can be better understood.

EXPERIMENTAL

Sample preparation. A single crystal of SrAl2O4:Eu2+:Dy3+ has been prepared by travelling solvent floating zone. The nominal concentration of Eu2+ is 1%. The crystal has been polished until a thickness of 187±1 µm to avoid a saturated absorption. The polishing also led to a very flat surface avoiding internal reflexions (see image in Figure 4).

Absorption spectroscopy. Absorption spectra of the sample at 300 K were recorded with an optical spectrometer (Agilent Cary 5000). To do that, the crystal was mounted on a copper plate with a previously drilled hole of about 0.1 mm in diameter and glued with silver paste.

(4)

3

Photoluminescence spectroscopy. The room temperature excitation and luminescence spectra were recorded on a Fluorolog 3-22 (Horiba Jobin Yvon), equipped with a water-cooled photo multiplier tube (PMT). For the low temperature measurements, a closed-cycle cryostat (Janis-Sumitomo SHI-4.5) equipped with a programmable temperature controller (Lakeshore Model 331) was used. For the quantum yield measurements an integrated sphere was used (details are given in the supplementary material and Figure S1).

Thermoluminescence. Thermoluminescence (TL) measurements were performed with a home-built system.25 A cartridge (Omega, CSH-101100) was used for the heating and a cold nitrogen gas for the cooling. The excitation source is a LED-light at 445 nm (HL-3WHPL-Roy445, 1.5 W). The temperature was monitored with a temperature sensor (Omega, Type T) and controlled with a PID-controller (Omega, CN7523). A photon counting unit (Hamamatsu, C3863) is used to detect the TL. See details of the measurements in the supporting material.

RESULTS

Optical characterization. The absorption spectrum of a 187 µm crystal of SrAl2O4:Eu2+:Dy3+ has been recorded at room temperature between 200 and 500 nm (Figure 1a). Two different broad bands at 254 nm and 330 nm are observed. These two bands are associated with the electric dipole allowed 4f–5d transitions. At 254 nm a maximum optical density of 4.07 is found whereas at 330 nm the maximum is slightly lower with a value of 3.68. The strong absorbance of the sample makes rather difficult to distinguish more individual bands in the spectrum. Since the UV-vis absorption spectrophotometer do not discriminate between the transmitted light from the probing wavelength after absorption and the intrinsic strong luminescence of the sample, the luminescence has been cut with different filters as explained in Figure S2.1 and no significant difference has been found in the absorption spectrum. The excitation spectra of the sample at room temperature and 6 K have been recorded between 250 and 500 nm (Figure 1b). The spectra are not exactly the same than the absorption one neither at room temperature nor at low temperature. It presents two broad bands due to the interconfigurational 4f7– 4f6 5d1 transitions between 250 and 460 nm with several maxima at 270 nm, 317 nm, 362 nm and 413 nm, in agreement with previous results observed in powders of the same composition.25 On the other hand the emission spectra of the sample at room temperature and 6 K have been recorded between 400 and 650 nm after excitation at 370 nm (Figure 1c). At room temperature the emission spectrum shows the characteristic broad band centred at 520 nm of Eu2+ corresponding to the electric dipole allowed 4f–5d transition. The broadening of this band is due to once again the interconfigurational transition from 4f6–5d1 to the ground state 8S7/2 since the electron in the 5d orbital couples strongly with the lattice resulting in a splitting of the energy levels by the crystal field. At 6 K the emission of the blue site appears centered at 455 nm. The latter band is shifted 10 nm upwards respect to the blue site emission in dysprosium free powder samples with the same concentration of europium (Figure S4b).10,26 A red shift of the Eu blue site emission was already observed by Jia et al.20 in single crystal fibers in the presence of dysprosium although no shift was observed in the presence of dysprosium in powder samples.8,25 Other studies relate the blue shift to the particle size of the phosphor.27 Furthermore, the presence of trivalent ions within the crystal was studied by selective irradiation at 472 nm (6H15/2 6H7/2 transition of Dy3+)28 and 522 nm ( 5D17F0 transition of Eu3+),29 where no absorption from the Eu2+ is expected. However, the strong emission due to the dipole allowed transitions from Eu2+ masks the weaker f-f magnetic (and slightly electric) dipole allowed emission bands from Dy+3 and Eu3+ (Figures S3a and S3b). It is not known whether the Eu3+ ions have been completely reduced to Eu2+ during the synthesis of the SrAl2O4:Eu2+:Dy3+ crystal. Alternatively, the presence of Dy has been confirmed by its characteristic near infrared absorption bands.30,31 The results are displayed in Figure S3c.

(5)

4

Figure 1. a) Room temperature absorption spectrum, b) room temperature and 6 K excitation spectra recorded at 520 nm, c) room temperature and 6 K emission spectra after excitation at 370 nm, d) room temperature excitation spectrum between 700 and 1000 nm recorded at 520 nm and emission spectrum after excitation at 875 nm (photostimulated luminescence).

Estimation of the extinction coefficient of the SrAl2O4:Eu2+:Dy3+ crystal. The actual concentration of europium in the doped crystal could be slightly different respect to the nominal one used for the synthesis (1% respect to strontium) due to the segregation that depends on the growth conditions.

Thus the real concentration of europium has been investigated as follows: The energy transfer enhancement from the blue site to the green site when increasing the nominal concentration of Eu2+

at 10 K has been proved for the SrAl2O4:Eu2+:Dy3+ powder samples.9 The dependence of the intensity ratio between the blue and the green emission with the concentration of Eu2+ follows a double exponential behaviour. The intensity ratio between the blue and green emission of the SrAl2O4:Eu2+:Dy3+ crystal at low temperature shown in Figure 1c has been used to extrapolate the value of the relative concentration of Eu2+ respect to the concentration of Sr2+ within the crystal. A relative concentration of 1 mol% has been found in agreement with the nominal concentration used for the synthesis (Figure S4). This value has been confirmed by Energy-dispersive X-ray spectroscopy (Figure S5). On the other hand, the molar concentration of Sr2+ within the crystal has been calculated by using equation 1:

(6)

5 [Sr2+](M) = ZNA

VC (L) = 4NA

383.8 ∙10−27(L) , (1)

where Z is the number of atoms per unit cell, NA is Avogadro constant and Vc is the unit cell volume calculated with the unit cell parameters from reference 5. A molar concentration of Sr2+ of 17.30 M is obtained. Therefore, a molar concentration of Eu2+ of 0.173 M is estimated. Thus, considering the later value, the thickness of the crystal and the optical density measured at 254 nm and 330 nm, the extinction coefficient of the SrAl2O4:Eu2+:Dy3+ crystal at these two wavelengths can be calculated by using equation 2:

ɛ(M−1cm−1) =[Eu2+](M)·thickness (cm) OD (2) ɛ = 254 nm= 4.07

0.173 M ·187·10−4 cm= 1258 M−1cm−1 ɛ = 330 nm= 3.68

0.173 M ·187·10−4 cm = 1137 M−1cm−1

An extinction coefficient value of 1258 M-1 cm-1 and 1137 M-1 cm-1 was found at 254 nm and 330 nm, respectively. Alternatively, a thinner crystal has been analysed and corresponding values of 1320 M-1 cm-1 (254 nm) and 1160 M-1 cm-1 (330 nm) have been found (Figure S2.2). These values are in good agreement with the results obtained for other Eu2+ compounds such as EuCl2 (ε = 1130 ± 56 M-1cm-1 ).32

Thermoluminescence measurements. Thermoluminescence (TL) measurements were performed on the crystal after irradiation at 223 K and 293 K with a 445 nm and 395 nm LED-light (see experimental section and Figure S6 for more details). In Figure 2a the results obtained by irradiating at 447 nm are shown. The shape of the curves is rather broad indicating a wide trap distribution. The full width at the half maximum (FWHM) is 113 K after excitation at 223 K and 85 K after excitation at 293 K, whereas the maxima of the afterglow peaks are centred at 328 K and 348 K after excitation at 223 K and 293 K, respectively. This indicates a thermal depletion of traps with increasing temperature in addition to identifying remarkably high ratio of the deep/shallow traps.

In Figure 2b the analogous results obtained by irradiating at 395 nm are shown. As observed before, the strong shape and intensity dependence of the TL curves with the irradiation wavelength is further evident after irradiation at low temperatures.25 Overall a significant decrease of the intensity is observed regardless the irradiation temperature. In addition, the curves are even broader with FWHM values of 143 K and 100 K after excitation at 223 K and 293 K, respectively. Contrary, the maxima of the afterglow peaks are not affected significantly and remain at slightly lower temperatures (320 K after excitation at 223 K, and 343 K after excitation at 293 K).

It is worth mentioning that besides the maxima of the peaks, a shoulder with relatively lower intensity appears at lower temperatures (250 K) when irradiating at 223 K, regardless the irradiation wavelength. In the case of irradiating at 293 K, a shoulder appears at a higher temperature (380 K) relative to the maxima of the peaks. This feature could be associated with two different kind of traps.

More importantly, the maxima observed at very high temperature is promising for a very long phosphorescence since they indicate the presence of very deep traps in the case of crystals. The depth of the traps was estimated with the initial rise analysis (Figure S7). In this method only the low temperature side of the TL curve is considered. Thus, one can assume that the concentration of trapped charge carriers is approximately constant.33 The analysis leads to a value of around 0.5 eV.

This value is in the optimal range for phosphorescence stablished by Matsuzawa et al.14 However, it must be taken into account that for a continuous distribution of traps, the depth is only due to the

(7)

6

shallowest occupied traps in the distribution since only the low-temperature region is analysed. The depth trap obtained for our SrAl2O4: Eu2+(1%):Dy3+ crystal is the same as the one recently obtained by Chithambo et al.34 In addition, the thermoluminescence curve of the SrAl2O4: Eu2+(1%):Dy3+ crystal has been measured after different irradiation times at room temperature and only a slight shift of 4 degrees to lower values of the temperature upon longer irradiation time has been found (Figure S8).

The shift of the maximum towards lower temperatures upon longer irradiation time has been attributed by Van den Eeckhout et al.33 in the case of CaAl2O4:Dy to a system more complex than a single trap obeying first order kinetics. For second order TL re-trapping of the excitations takes place.21

Figure 2. Thermoluminescence curve recorded for the SrAl2O4:Eu2+(1%):Dy3+ crystal after irradiation at a) 447 nm and b) 395 nm at 223 K and 293 K.

Quantum yield measurements. Quantum yield (QY) measurements on solid samples are usually difficult to analyse since the luminescence presents preferred directions instead of being distributed uniformly in the space. This is the case for instance of very thick samples such as powders or crystals where the excitation itself is not uniform. Because of that, integrated spheres which redistribute isotropically the light over its interior surface are used to calculate the QY.35 Thus, the absolute QY of the SrAl2O4:Eu2+(1%):Dy3+ crystal upon different irradiation wavelengths at room temperature has been measured with a quantum sphere and compared with the results obtained for a dysprosium free SrAl2O4:Eu2+(1%)powder sample and for a containing dysprosium SrAl2O4:Eu2+(1%):Dy3+ powder sample (Figure 3). As observed for the SrAl2O4:Eu2+(1%):Dy3+ crystal the irradiation at lower energy leads to higher values of the QY. Indeed, at lower energies the probability of an electron of the Eu2+ to be excited into the conduction band of the host is lower. Hence, the radiative recombination is favoured above 400 nm. Ueda et al. have reported the onset of photoconductivity in SrAl2O4:Eu2+:Dy3+ below 400nm.36 This observation can account for the reduced QY below 400nm, as well as the fact that the blue emitting Eu2+ which can transfer its energy10 to the green emitting Eu2+ is quenched at room temperature and does not contribute to the intensity of the emission at 520 nm. In order to measure the QY of the

(8)

7

powder samples some considerations to check the reproducibility of the results have been done (see supplementary material and Figure S9). Then, the QY at different irradiation wavelengths has been measured by using 0.5 g of powder for both SrAl2O4:Eu2+(1%):Dy3+ and SrAl2O4:Eu2+(1%). The results are shown in Figure 3. As for the SrAl2O4:Eu2+(1%):Dy3+ crystal, the lower the irradiation energy the higher the QY value for the powder samples. However, in the case of the SrAl2O4:Eu2+(1%) sample the values are higher indicating that somehow the presence of dysprosium in SrAl2O4:Eu2+(1%):Dy3+ decreases the QY, significantly. This is in agreement with the longer afterglow time observed for the Eu-Dy co-doped samples due to the dysprosium traps and the lower concentration of Eu2+ available for radiative recombination.13 In addition, the non-radiative release of the dysprosium traps may contribute to the lowering of the QY. In the case of the SrAl2O4:Eu2+(1%):Dy3+ powder the values are even lower than in the crystal. The later observations reveal that the presence of dysprosium has a strong effect not only in the duration of the phosphorescence but also in the QY of the material.

Very recently, Van der Heggen et al37 have reported that the “trapped charge carriers can be optically detrapped by the excitation light”, which implies that these trapped charge carriers equally absorb very strongly around 350-450 nm. Such influence can be only understood considering a high proportion of europium and dysprosium sites participating in the long persistent phosphorescence mechanism, that is to say, a high number of europium sites being excited. However, the absolute storage capacity of SrAl2O4:Eu2+:Dy3+ was estimated by Van der Heggen et al.38 by concluding that only at around 1.6 % of the Eu ions participate in the phosphorescence mechanism. On the other hand, it is worthy to say that the lowering of the QY values in the presence of dysprosium is more obvious at lower irradiation wavelengths in the case of the crystal. Under higher irradiation wavelengths the QY values of the SrAl2O4:Eu2+(1%):Dy3+ crystal and SrAl2O4:Eu2+(1%) start to approach being the same under irradiation at 425 nm. That indicates that the dysprosium traps must be closer to the Eu2+ blue site since only when this site is excited the dysprosium traps show an effect on the QY.

The highest QY values of the same compound have been previously stablished as 55 %,39 although higher values have been recently observed in the case of cage-like microspheres of SrAl2O4:Eu2+ up to 95% and 94% after excitation at 365 nm and 460 nm respectively,40 and 28% and 71% in the case of SrAl2O4:Eu2+:Dy3+ powder samples after excitation at 375 nm and 445 nm respectively at very low excitation intensities.37

In order to verify the reliability of the values obtained with the integrating sphere, the internal QY of the SrAl2O4: Eu2+(1%) powder samples has been estimated through lifetimes measurements of the green site emission at 10 K and 300 K after irradiation at 370 nm. From reference 9 lifetimes of 1190 ns and 408 ns were obtained for the mentioned samples at 3 K and 300 K, respectively. Hence, the QY of the sample has been estimated as the ratio between both values as shown in equation 3.

Φint =k kr

r+ knr = 1/τ1/τn= ττ

n = τ ( 300 K)τ ( 3 K) (3)

, where τn is the natural radiative lifetime at low temperature in the absence of thermal quenching and τrepresents the total radiative and non-radiative lifetimes taking place due to thermal activation at high temperatures.

A value of 34% has been obtained for SrAl2O4:Eu2+(1%), in a very good agreement with the value obtained by the integrating sphere (35%). Consequently, the wavelength dependence study carried out through this method represents a reliable and fast technique to evaluate the differences between all the samples under study here.

(9)

8

Figure. 3. Quantum yield measurements performed in the SrAl2O4:Eu2+(1%):Dy3+ crystal, and SrAl2O4:Eu2+(1%) and SrAl2O4:Eu2+(1%):Dy3+ powder samples at different wavelengths.

CONCLUSIONS

In Figure 4, a summary of the optical properties of the SrAl2O4:Eu2+:Dy3+ single crystal and a recapitulation of the information obtained in this study are shown. A very thin and high-quality single crystal of SrAl2O4: Eu2+(1%): Dy3+ has been optically characterised. The absorption and excitation spectra show the broad bands associated with the electric dipole allowed 4f–5d transitions. The emission spectrum at room temperature shows the characteristic broad band centred at 520 nm. At 6 K the emission of the blue site appears centred at 455 nm. The latter band is shifted 10 nm upwards respect to the blue site emission in dysprosium free powder samples. Besides, photoinduced detrapping of the deeper traps by irradiation in the near infrared region has been proven.

On the other hand, the absorption spectrum of the crystal and the estimation of the real Eu2+

concentration within the crystal through the intensity ratio of blue and green emission at 10 K allowed to obtain reliable values of the extinction coefficient for any SrAl2O4: Eu2+: Dy3+ sample. The value of the extinction coefficient has been estimated as 1258 M-1 cm-1 and 1137 M-1 cm-1 at 254 nm and 330 nm, respectively.

In addition, the thermoluminescence curves of the SrAl2O4: Eu2+(1%): Dy3+ crystal reveal the presence of very deep traps, which is crucial to get a longer phosphorescence due to the higher thermal energy needed to release them.

Furthermore, the dependence of the quantum yield values of the SrAl2O4: Eu2+(1%): Dy3+ crystal with the excitation wavelength has been followed. An increase of the QY values has been obtained when increasing the irradiation wavelength since the radiative recombination is favoured when only the green europium site is excited. More importantly, when the dependence is compared with the one of a dysprosium free sample, the QY values differ more at lower irradiation wavelengths when the Eu blue site is excited indicating that the dysprosium traps are situated closer to the blue sites.

These results shed light upon the role of co-dopants in the long- phosphorescence mechanism. In this regard, the analysis of SrAl2O4: Eu2+: Dy3+ crystals that in addition contained further dopants such as barium or boron16,41,42 is needed to achieve this goal.

Moreover, our estimation of the extinction coefficient serves as an orientation to calculate the real concentration of europium in any kind of SrAl2O4 samples, which is crucial to rationalise the experimentally observed phosphorescence.

55 50 45 40 35

30 25 20

QY /%

420 410 400 390 380 370 360

exc /nm SrAl2O4:Eu2+(1%):Dy3+ crystal SrAl2O4:Eu2+(1%) powder SrAl2O4:Eu2+(1%):Dy3+ powder

(10)

9

Figure 4. Summary of optical properties of the SrAl2O4:Eu2+:Dy3+ single crystal and information obtained.

ASSOCIATED CONTENT

Supporting Information. This material is available free of charge via the Internet at http://pubs.acs.org.

Quantum yield measurements of SrAl2O4:Eu2+:Dy3+ single crystal and powder samples, Luminescence spectra of SrAl2O4:Eu2+:Dy3+ single crystal and powder samples at different excitations, UV Absorption spectroscopy measurements, Thermoluminescence measurements, Quantum yield measurements of SrAl2O4:Eu2+:Dy3+ powder samples

AUTHOR INFORMATION Corresponding author.

*Mariateresa.delgadoperez@unige.ch +41 22 37 96106

Author Contributions

All authors have given approval to the final version of the manuscript.

AKNOWLEDGEMENT

We are grateful to Dr. Enrico Giannini for the help during the synthesis of the crystal and to the Swiss National Science Foundation (Project number: 200021-169033), as well the KTI (Project number 25902.1 PFNM-NM) for the financial support.

(11)

10 REFERENCES

(1) Rojas-Hernandez, R. E.; Rubio-Marcos, F.; Rodriguez, M. Á.; Fernandez, J. F. Long lasting phosphors: SrAl2O4:Eu,Dy as the most studied material. Renewable and Sustainable Energy Reviews 2018, 81, 2759-2770.

(2) Li, Y.; Gecevicius, M.; Qiu, J. Long persistent phosphors-from fundamentals to applications.

Chem. Soc. Rev. 2016, 45, 2090-2136.

(3) Li, H.; Yin, S.; Wang, Y.; Sato, T. Persistent Fluorescence-Assisted TiO2-xNy-Based Photocatalyst for Gaseous Acetaldehyde Degradation. Environmental Science & Technology 2012, 46, 7741-7745.

(4) Chunbo, W.; Tongtong, X.; Jiaqing, L.; Huili, L.; Zhuo, S. Long Afterglow SrAl2O4:Eu2+,Dy3+

Phosphors as Luminescent Down-Shifting Layer for Crystalline Silicon Solar Cells. International Journal of Applied Ceramic Technology 2015, 12, 722-727.

(5) Schulze, A. R.; Buschbaum, H. M. Zur Verbindungsbildung von MeO: M2O3. IV. Zur Struktur von monoklinem SrAl2O4. Zeitschrift für anorganische und allgemeine Chemie 1981, 475, 205-210.

(6) http://abulafia.mt.ic.ac.uk/shannon/radius.php.

(7) Poort, S. H. M.; Blokpoel, W. P.; Blasse, G. Luminescence of Eu2+ in barium and strontium aluminate and gallate. Chemistry of Materials 1995, 7, 1547-1551.

(8) Botterman, J.; Joos, J. J.; Smet, P. F. Trapping and detrapping in SrAl2O4:Eu,Dy persistent phosphors: influence of excitation wavelength and temperature. Phys. Rev. B: Condens. Matter Mater.

Phys. 2014, 90, 085147/085141-085147/085115, 085115 pp.

(9) Nazarov, M.; Brik, M. G.; Spassky, D.; Tsukerblat, B. Crystal field splitting of 5d states and luminescence mechanism in SrAl2O4:Eu2+ phosphor. Journal of Luminescence 2017, 182, 79-86.

(10) Bierwagen, J.; Yoon, S.; Gartmann, N.; Walfort, B.; Hagemann, H. Thermal and concentration dependent energy transfer of Eu2+ in SrAl2O4. Opt. Mater. Express 2016, 6, 793-803.

(11) Yamamoto, H.; Matsuzawa, T. Mechanism of long phosphorescence of SrAl2O4:Eu2+, Dy3+ and CaAl2O4:Eu2+, Nd3+. Journal of Luminescence 1997, 72-74, 287-289.

(12) Aitasalo, T.; Dereń, P.; Hölsä, J.; Jungner, H.; Krupa, J. C.; Lastusaari, M.; Legendziewicz, J.;

Niittykoski, J.; Stręk, W. Persistent luminescence phenomena in materials doped with rare earth ions.

Journal of Solid State Chemistry 2003, 171, 114-122.

(13) Nakazawa, E.; Mochida, T. Traps in SrAl2O4:Eu2+ phosphor with rare-earth ion doping. J.

Lumin. 1997, 72-74, 236-237.

(14) Matsuzawa, T.; Aoki, Y.; Takeuchi, N.; Murayama, Y. A new long phosphorescent phosphor with high brightness, SrAl2O4:Eu2+,Dy3+. J. Electrochem. Soc. 1996, 143, 2670-2673.

(15) Dorenbos, P. Mechanism of Persistent Luminescence in Eu2+ and Dy3+ Codoped Aluminate and Silicate Compounds. J. Electrochem. Soc. 2005, 152, H107-H110.

(16) Clabau, F.; Rocquefelte, X.; Jobic, S.; Deniard, P.; Whangbo, M. H.; Garcia, A.; Le Mercier, T.

Mechanism of Phosphorescence Appropriate for the Long-Lasting Phosphors Eu2+-Doped SrAl2O4 with Codopants Dy3+ and B3+. Chem. Mater. 2005, 17, 3904-3912.

(17) Ning, L.; Huang, X.; Huang, Y.; Tanner, P. A. Origin of the green persistent luminescence of Eu- doped SrAl2O4 from a multiconfigurational ab initio study of 4f7 → 4f65d1 transitions. Journal of Materials Chemistry C 2018, 6, 6637-6640.

(18) Finley, E.; Mansouri Tehrani, A.; Brgoch, J. Intrinsic Defects Drive Persistent Luminescence in Monoclinic SrAl2O4:Eu2+. The Journal of Physical Chemistry C 2018, 122, 16309-16314.

(19) Shionoya, S.; Yen, W. M.; Yamamoto, H.: Phosphor Handbook; CRC Press, 2006.

(20) Jia, W.; Yuan, H.; Lu, L.; Liu, H.; Yen, W. M. Phosphorescent dynamics in SrAl2O4:Eu2+, Dy3+ single crystal fibers. J. Lumin. 1998, 76&77, 424-428.

(21) Jia, W.; Yuan, H.; Lu, L.; Liu, H.; Yen, W. M. Crystal growth and characterization of Eu2+, Dy3+:SrAl2O4 and Eu2+, Nd3+:CaAl2O4 by the LHPG method. Journal of Crystal Growth 1999, 200, 179- 184.

(22) Katsumata, T.; Sakai, R.; Komuro, S.; Morikawa, T. Thermally Stimulated and Photostimulated Luminescence from Long Duration Phosphorescent SrAl2O4:Eu, Dy Crystals. J. Electrochem. Soc. 2003, 150, H111-H114.

(12)

11

(23) Ishibashi, T.; Furukawa, S.; Yamanaka, A. Crystal growth of long phosphorescent phosphor SrAl2O4:Eu,Dy. Kidorui 2013, 62, 12-13.

(24) Terraschke, H.; Suta, M.; Adlung, M.; Mammadova, S.; Musayeva, N.; Jabbarov, R.; Nazarov, M.; Wickleder, C. SrAl2O4:Eu2+(,Dy3+) Nanosized Particles: Synthesis and Interpretation of Temperature- Dependent Optical Properties. Journal of Spectroscopy 2015, 2015, 12.

(25) Hagemann, H.; Lovy, D.; Yoon, S.; Pokrant, S.; Gartmann, N.; Walfort, B.; Bierwagen, J.

Wavelength dependent loading of traps in the persistent phosphor SrAl2O4:Eu2+, Dy3+. J. Lumin. 2016, 170, 299-304.

(26) Dutczak, D.; Justel, T.; Ronda, C.; Meijerink, A. Eu2+ luminescence in strontium aluminates.

Physical Chemistry Chemical Physics 2015, 17, 15236-15249.

(27) Sahu, I. P.; Bisen, D. P.; Sharma, R. UV excited green luminescence of SrAl2O4:Eu2+, Dy3+

nanophosphor. Research on Chemical Intermediates 2016, 42, 2791-2804.

(28) Liepina, V.; Millers, D.; Smits, K.; Zolotarjovs, A.; Bite, I. X-ray excited luminescence of SrAl2O4:Eu,Dy at low temperatures. J. Phys. Chem. Solids 2018, 115, 381-385.

(29) Binnemans, K. Interpretation of europium(III) spectra. Coordination Chemistry Reviews 2015, 295, 1-45.

(30) Lisiecki, R.; Solarz, P.; Niedźwiedzki, T.; Ryba-Romanowski, W.; Głowacki, M. Gd3Ga3Al2O12

single crystal doped with dysprosium: Spectroscopic properties and luminescence characteristics.

Journal of Alloys and Compounds 2016, 689, 733-739.

(31) Delgado, T.; Ajoubipour, S.; Afshani, J.; Yoon, S.; Walfort, B.; Hagemann, H. Spectroscopic properties of Dy3+- and Dy3+, B3+- doped SrAl2O4. Optical Materials 2019, 89, 268-275.

(32) Uehara, A.; Shirai, O.; Nagai, T.; Fujii, T.; Yamana, H. Spectroelectrochemistry and electrochemistry of europium ions in alkali chloride melts. Z. Naturforsch., A: Phys. Sci. 2007, 62, 191- 196.

(33) Van den Eeckhout, K.; Bos, A. J. J.; Poelman, D.; Smet, P. F. Revealing trap depth distributions in persistent phosphors. Physical Review B 2013, 87, 045126.

(34) Chithambo, M. L.; Wako, A. H.; Finch, A. A. Thermoluminescence of SrAl2O4:Eu2+, Dy3+: Kinetic analysis of a composite-peak. Radiat. Meas. 2017, 97, 1-13.

(35) Palmer, D. A. A System of Mesopic Photometry. Nature 1966, 209, 276.

(36) Ueda, J.; Nakanishi, T.; Katayama, Y.; Tanabe, S. Optical and optoelectronic analysis of persistent luminescence in Eu2+-Dy3+ codoped SrAl2O4 ceramic phosphor. physica status solidi c 2012, 9, 2322-2325.

(37) Van der Heggen, D.; Joos, J. J.; Smet, P. F. Importance of Evaluating the Intensity Dependency of the Quantum Efficiency: Impact on LEDs and Persistent Phosphors. ACS Photonics 2018, 5, 4529- 4537.

(38) Van der Heggen, D.; Joos, J.; Rodríguez Burbano, D.; Capobianco, J.; Smet, P. Counting the Photons: Determining the Absolute Storage Capacity of Persistent Phosphors. Materials 2017, 10, 867.

(39) Meister, F.; Batentschuk, M.; Dröscher, S.; Osvet, A.; Stiegelschmitt, A.; Weidner, M.;

Winnacker, A. Eu2+ luminescence in the EuAl2O4 concentrated phosphor. Radiation Measurements 2007, 42, 771-774.

(40) Wan, J.; Zhang, Y.; Wang, Y.; Ma, R.; Wu, Y.; Qiao, X.; Fan, X. Facile synthesis of monodisperse SrAl2O4:Eu2+ cage-like microspheres with an excellent luminescence quantum yield. Journal of Materials Chemistry C 2018, 6, 3346-3351.

(41) Van den Eeckhout, K.; Poelman, D.; Smet, P. F. Persistent Luminescence in Non-Eu(2+)-Doped Compounds: A Review. Materials 2013, 6, 2789-2818.

(42) Yoon, S.; Bierwagen, J.; Trottmann, M.; Walfort, B.; Gartmann, N.; Weidenkaff, A.; Hagemann, H.; Pokrant, S. The influence of boric acid on improved persistent luminescence and thermal oxidation resistance of SrAl2O4:Eu2+. Journal of Luminescence 2015, 167, 126-131.

Références

Documents relatifs

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

In the case of the pulse magnetron diode, filled with argon at low pressure, it has been found that the spectrum con- sists of the lines which correspond to tr- ansitions

Now that a relation between the threshold friction velocity and the diameter of the po- tentially mobilized particles is available, the friction velocity in a powder snow avalanche

The shapes obtained following the combination of the genetic algorithm reported in [20] and the topological optimization by the mixed relaxation/level set method are presented

TiO emission from oxygen-rich Mira-type S stars Out of a sample of 87 S-type AGB stars, 49 sources do not show significant dust emission features, and of these NP Aur is the only

Abstract 2014 The temperature variations of the order parameter invariants are measured in the uniaxial and biaxial nematic phases of different mixtures of potassium

I t may be seen in this expression that the stress- dependent change in the critical field depends upon the two-ion magnetostrictive constant Dl and D, and

- Zeeman spectra with the field along the ordering direction displayed splitting appropriate to the g = 19 ground state.. Zeeman spectra were also obtained with the