spectroscopic properties and energy level location of eu2

6
Spectroscopic properties and energy level location of Eu 2+ in Sr 2 Si 5 N 8 phosphor Agata Lazarowska a,, Sebastian Mahlik a , Marek Grinberg a , Chiao-Wen Yeh b , Ru-Shi Liu b,a Institute of Experimental Physics, University of Gdansk, Wita Stwosza 57, 80-952 Gdan ´sk, Poland b Department of Chemistry, National Taiwan University, Taipei 106, Taiwan article info Article history: Received 1 May 2014 Received in revised form 29 August 2014 Accepted 1 September 2014 Available online 3 October 2014 Keywords: Nitridosilcate Europium Luminescence Anharmonicity of potential Impurity trapped exciton (ITE) Nephelauxetic effect abstract In this contribution luminescence properties of Sr 2x Si 5 N 8 :Eu x 2+ (x = 0.02 and 0.1) at temperatures from 10 to 600 K are presented. Both materials exhibit broad band red emission assigned to the parity allowed 4f 6 5d 1 ? 4f 7 transitions in Eu 2+ ions occupying two different crystallographic sites labeled as Eu(Sr1) and Eu(Sr2). For low concentration of dopants only Sr1 site is occupied, whereas for higher concentration both Sr1 and Sr2 sites are occupied. Optical data are used to construct configurational coordination diagrams of the Eu(Sr1) and Eu(Sr2) centers in Sr 2 Si 5 N 8 lattice. Effects of temperature luminescence quenching and luminescence lifetime shortening were related to ionization of Eu 2+ ions and anharmonicity of adiabatic lattice potential. Effect of non-homogeneity of the Eu 2+ occupation between the Sr1 and Sr2 sites is dis- cussed by considering the energies of the ground states of Eu(Sr1) and Eu(Sr2) in respect to the valance and conduction bands. Ó 2014 Elsevier B.V. All rights reserved. 1. Introduction Recently, nitridosilicates doped with lanthanide ions have received interest due to high chemical stability, high hardness, excellent luminescence and thermo-mechanical properties. Among nitridosilicates, Sr 2 Si 5 N 8 :Eu 2+ is a promising red phosphor for light emitting diodes (LED’s) due to very efficient emission after UV or blue light excitation [1–5]. Moreover Sr 2 Si 5 N 8 :Eu 2+ can be synthe- sized under ambient pressure and therefore have the lower pro- duction cost in comparison to high pressure synthesized nitridosilicates [6]. Luminescence of Eu 2+ in Sr 2 Si 5 N 8 is related to the parity allowed 4f 6 5d ? 4f 7 transitions. The relatively long wavelength of the Eu 2+ emission is attributed to high covalence of the lattice bonds and a large crystal field splitting of the 4f 6 5d electronic configuration of the Eu 2+ coordinated by nitrogen ions [7]. The Sr 2 Si 5 N 8 crystallizes as orthorhombic lattice described by the space group of Pmn2 1 [6,8]. In the Sr 2 Si 5 N 8 lattice, Eu 2+ ions can replace Sr 2+ at two different crystallographic sites, tenfold coordinated site with mean distance of Sr1–N equal to 0.2928(7) nm and eightfold coordinated site with mean distance of Sr2–N equal to 0.2865(6) nm [3]. The luminescence of Sr 2 Si 5 N 8 : Eu 2+ varies from orange to red, depending on the Eu 2+ concentra- tion [3,9,10]. This effect has been explained by several mecha- nisms. Piao et al. [3] have argued that the red shift of emission band caused by increasing of Eu 2+ concentration is associated with the uneven distribution of the Eu 2+ in Sr1 and Sr2 positions. Sohn et al. [10] have assumed that Eu 2+ activators are evenly distributed to both Sr 2+ sites regardless of the Eu 2+ concentration and the red shift is associated with an energy transfer mechanism from Eu 2+ site with the higher emission energy to the site with the lower emission energy that increase for the higher Eu 2+ concentration. Additionally, decrease of unit cell volume of Sr 2 Si 5 N 8 due to increase of incorporation of Eu 2+ ions in the lattice is observed. Since Eu 2+ ions are smaller than Sr 2+ the size of Eu–N polyhedron is reduced in comparison to Sr–N polyhedron which increase the crystal field strength and eventually lead to the red shift of emis- sion. Lie et al. [9] have claim that with increase of Eu 2+ concentra- tion the reabsorption process starts to have impact, giving rise to the observed red shift of the emission. Numerous investigations have been dedicated to thermal quenching of Eu 2+ luminescence in Sr 2 Si 5 N 8 [3,6]. There are two proposal that have been used to explain thermal quenching of the Eu 2+ emission. The first is related to thermally activated electron in the excited state, that crosses the intersection of the lowest 4f 6 5d excited state and the 4f 7 ground state, and reaches the ground state nonradiatively [3,6,11]. The second assumes thermally activated ionization (autoionization) from the 4f 6 5d 1 excited state to the conduction band [12–14]. http://dx.doi.org/10.1016/j.optmat.2014.09.002 0925-3467/Ó 2014 Elsevier B.V. All rights reserved. Corresponding authors. E-mail addresses: [email protected] (A. Lazarowska), [email protected] (R.-S. Liu). Optical Materials 37 (2014) 734–739 Contents lists available at ScienceDirect Optical Materials journal homepage: www.elsevier.com/locate/optmat

Upload: others

Post on 09-Apr-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Optical Materials 37 (2014) 734–739

Contents lists available at ScienceDirect

Optical Materials

journal homepage: www.elsevier .com/locate /optmat

Spectroscopic properties and energy level location of Eu2+ in Sr2Si5N8

phosphor

http://dx.doi.org/10.1016/j.optmat.2014.09.0020925-3467/� 2014 Elsevier B.V. All rights reserved.

⇑ Corresponding authors.E-mail addresses: [email protected] (A. Lazarowska), [email protected]

(R.-S. Liu).

Agata Lazarowska a,⇑, Sebastian Mahlik a, Marek Grinberg a, Chiao-Wen Yeh b, Ru-Shi Liu b,⇑a Institute of Experimental Physics, University of Gdansk, Wita Stwosza 57, 80-952 Gdansk, Polandb Department of Chemistry, National Taiwan University, Taipei 106, Taiwan

a r t i c l e i n f o a b s t r a c t

Article history:Received 1 May 2014Received in revised form 29 August 2014Accepted 1 September 2014Available online 3 October 2014

Keywords:NitridosilcateEuropiumLuminescenceAnharmonicity of potentialImpurity trapped exciton (ITE)Nephelauxetic effect

In this contribution luminescence properties of Sr2�xSi5N8:Eux2+ (x = 0.02 and 0.1) at temperatures from

10 to 600 K are presented. Both materials exhibit broad band red emission assigned to the parity allowed4f65d1 ? 4f7 transitions in Eu2+ ions occupying two different crystallographic sites labeled as Eu(Sr1) andEu(Sr2). For low concentration of dopants only Sr1 site is occupied, whereas for higher concentration bothSr1 and Sr2 sites are occupied. Optical data are used to construct configurational coordination diagramsof the Eu(Sr1) and Eu(Sr2) centers in Sr2Si5N8 lattice. Effects of temperature luminescence quenching andluminescence lifetime shortening were related to ionization of Eu2+ ions and anharmonicity of adiabaticlattice potential. Effect of non-homogeneity of the Eu2+ occupation between the Sr1 and Sr2 sites is dis-cussed by considering the energies of the ground states of Eu(Sr1) and Eu(Sr2) in respect to the valanceand conduction bands.

� 2014 Elsevier B.V. All rights reserved.

1. Introduction

Recently, nitridosilicates doped with lanthanide ions havereceived interest due to high chemical stability, high hardness,excellent luminescence and thermo-mechanical properties. Amongnitridosilicates, Sr2Si5N8:Eu2+ is a promising red phosphor for lightemitting diodes (LED’s) due to very efficient emission after UV orblue light excitation [1–5]. Moreover Sr2Si5N8:Eu2+ can be synthe-sized under ambient pressure and therefore have the lower pro-duction cost in comparison to high pressure synthesizednitridosilicates [6]. Luminescence of Eu2+ in Sr2Si5N8 is related tothe parity allowed 4f65d ? 4f7 transitions. The relatively longwavelength of the Eu2+ emission is attributed to high covalenceof the lattice bonds and a large crystal field splitting of the 4f65delectronic configuration of the Eu2+ coordinated by nitrogen ions[7].

The Sr2Si5N8 crystallizes as orthorhombic lattice described bythe space group of Pmn21 [6,8]. In the Sr2Si5N8 lattice, Eu2+ ionscan replace Sr2+ at two different crystallographic sites, tenfoldcoordinated site with mean distance of Sr1–N equal to0.2928(7) nm and eightfold coordinated site with mean distanceof Sr2–N equal to 0.2865(6) nm [3]. The luminescence of Sr2Si5N8:

Eu2+ varies from orange to red, depending on the Eu2+ concentra-tion [3,9,10]. This effect has been explained by several mecha-nisms. Piao et al. [3] have argued that the red shift of emissionband caused by increasing of Eu2+ concentration is associated withthe uneven distribution of the Eu2+ in Sr1 and Sr2 positions. Sohnet al. [10] have assumed that Eu2+ activators are evenly distributedto both Sr2+ sites regardless of the Eu2+ concentration and the redshift is associated with an energy transfer mechanism from Eu2+

site with the higher emission energy to the site with the loweremission energy that increase for the higher Eu2+ concentration.Additionally, decrease of unit cell volume of Sr2Si5N8 due toincrease of incorporation of Eu2+ ions in the lattice is observed.Since Eu2+ ions are smaller than Sr2+ the size of Eu–N polyhedronis reduced in comparison to Sr–N polyhedron which increase thecrystal field strength and eventually lead to the red shift of emis-sion. Lie et al. [9] have claim that with increase of Eu2+ concentra-tion the reabsorption process starts to have impact, giving rise tothe observed red shift of the emission. Numerous investigationshave been dedicated to thermal quenching of Eu2+ luminescencein Sr2Si5N8 [3,6]. There are two proposal that have been used toexplain thermal quenching of the Eu2+ emission. The first is relatedto thermally activated electron in the excited state, that crosses theintersection of the lowest 4f65d excited state and the 4f7 groundstate, and reaches the ground state nonradiatively [3,6,11]. Thesecond assumes thermally activated ionization (autoionization)from the 4f65d1 excited state to the conduction band [12–14].

A. Lazarowska et al. / Optical Materials 37 (2014) 734–739 735

Autoionization of the Eu2+ creates Eu3+ and an electron in the con-duction band. However the mechanism of the non-radiativerecombination process of thermally excited ecb electrons in theconduction band at Eu3+ ions was not discussed in the literature.

Time-resolved spectroscopy is essential method to distinguishdifferent optical active centers, especially when the large spectraloverlap between excitation and emission spectra of these centersis observed. In this paper we performed the systematic study oftemperature dependence of the Eu2+ luminescence, for the rangefrom 10 K to 600 K. We have measured time-resolved lumines-cence of the Sr2Si5N8:Eu2+ phosphor with two concentration ofdopants (2 at.%, 10 at.%). Time evolution of the luminescence widthand maximum as well as luminescence decay at different part ofthe spectrum were examined. Using these data we discussed themechanisms of thermal luminescence quenching.

Fig. 1. Emission and excitation spectra of Sr1.98Si5N8:Eu2+0.02 (a) and Sr1.9Si5N8:Eu2+

0.1

(b). Excitation spectra of the emission (dashed curves) was recorded monitoredluminescence at 15,649 cm�1 (639 nm) for Sr1.9Si5N8:Eu2+

0.1 and at 16,260 cm�1

(615 nm) for Sr1.98Si5N8:Eu2+0.02. The emission spectra (solid curves) was collected at

RT under excitation at 460 nm for both samples.

2. Experiment

Sr2�xSi5N8:Eux2+ (x = 0.02, 0.1) that corresponds to 2 at.% and

10 at.% were synthesized by the solid-state reaction from Sr3N2,Si3N4 and EuN. The mixture of substrates were fired at 1400 �Cfor 16 h under flowing 90% N2 – 10% H2 atmosphere in a tube fur-nace. The sintered products were ground again, yielding crystallinepowder.

The obtained compounds have a orthorhombic crystal structurewith a space group Pmn21. The unit cell parameters area = 5.7138(4) Å, b = 6.8211(1) Å, c = 9.3363(4) Å for Sr1.98Si5N8:Eu2+

0.02 and a = 5.7110(3) Å, b = 6.8299(4) Å, c = 9.3298(4) forSr1.9Si5N8:Eu2+

0.1. More details on preparation and structural investi-gation are described in the paper Yeh et al. [6].

Room temperature luminescence excitation spectra were mea-sured using a spectrofluorimeter FluoroMax-4P TCSPC produced byHoriba. The excitation source in this system was a 150 W Xenonlamp emitting in the 220–850 nm range. Fluorescence intensitywas measured using a monochromator and a photomultiplierR928 Side-on. The experimental setup for luminescence kineticsand time resolved emission spectra consists of YAG:Nd laser ofPL 2143 A/SS type and the parametric optical generator PG 401/SH (OPG) [15]. The laser generated pulses of the 355 nm wave-length with frequency 10 Hz and pump PG generator which couldproduce light pulses with wavelength ranging from 220 nm to2200 nm. The halfwidth of the time profile of the light pulse pro-duced by an excitation source under experimental conditions is44 ps and includes the half time of the light pulse duration pro-duced by the OPG coupled to the YAG:Nd laser and the timeresponse of the detector. Emission signal was analyzed by thespectrometer 2501S (Bruker Optics) and a Hamamatsu Streak Cam-era model C4334-01 with a final spectral resolution of about0.47 nm. Time resolved luminescence spectra were obtained byintegration of streak camera images over time intervals, whereasluminescence decays were obtained by the integration of streakcamera images over the wavelength intervals. Sample were cooledby an APD Cryogenics closed-cycle DE-202 optical cryostat, whichallows to vary temperature between 10 K and 600 K.

3. Results and discussion

The luminescence and luminescence excitation spectra of theSr2�xSi5N8:Eux

2+ (x = 0.02, 0.1) obtained at room temperature arepresented in Fig. 1a and b. The emission spectra obtained underblue excitation at 460 nm consist of broad band attributed to the4f65d1 ? 4f7 transitions of Eu2+ ions, with maximum at 615 nmfor Sr1.98Si5N8:Eu2+

0.02 and at 639 nm for Sr1.9Si5N8:Eu2+0.1. For both

samples the luminescence excitation spectra consist of several over-lapping broad bands that are almost independent of the monitored

wavelength and Eu2+ concentration. The emission band with higherenergy is expected for the Eu2+ in ten-fold coordinated Sr1 site withmean distance Sr1–N of 2.928(7) Å and the emission band withlower energy is expected for the Eu2+ in the eight-fold coordinatedSr2 site with mean distance Sr2–N 2.865(6) Å [3]. When Eu2+ con-centration increases from 2 at.% to 10 at.% the strong red shift ofthe emission band (equal to 658 cm�1) and broadening of the band-width (equal to 292 cm�1) are observed. These effects can beexplained assuming uneven occupation of the Eu2+ in Sr1 and Sr2positions. The Eu2+ ions occupy mainly Sr1 sites in the sample with2 at.% Eu2+ concentration (Fig. 1a), while for the larger concentrationof Eu2+ (10 at.%), dopant ions occupy both: Sr1 and Sr2 sites (Fig. 1b).The emission spectrum of Sr1.9Si5N8:Eu2+

0.1 (Fig. 1b) was decomposedinto two Gaussian bands – the first with maximum at 15,905 cm�1

(629 nm) corresponding to Eu2+ located in the Sr1 position and thesecond with maximum at 14,959 cm�1 (669 nm) corresponding toEu2+ in the Sr2 site. One notices that maximum of the Eu(Sr1) lumi-nescence is slightly shifted to the red (by 355 cm�1) in the samplewith 10 at.% of Eu2+ compared to the sample with 2 at.% of Eu2+.The Eu2+ ion is slightly smaller than Sr2+ ion and thus the increaseof concentration of Eu2+ ions in Sr2Si5N8 reduces the Eu–N averagedistance. This causes the increase of crystal field strength and split-ting of the 4f65d1 electronic manifold.

For more detailed consideration of two different sites inSr2Si5N8:Eu2+ the time-resolved luminescence spectra were mea-sured. The main advantage of time resolved technique is the possi-bility to observe changes of the spectrum line-shape in time. Fig 2aand b present time resolved luminescence spectra obtained under460 nm pulse excitation at 10 K temperature for Sr1.98Si5N8:Eu2+

0.02

and Sr1.9Si5N8:Eu2+0.1, respectively. Shape and maximum of the

Sr1.98Si5N8:Eu2+0.02 emission (Fig. 2a) does not change over the time,

which suggests that only one the Eu(Sr1) luminescence center con-tributes to this emission. The time evolution of the Sr1.9Si5N8:Eu2+

0.1

luminescence is different, namely the emission band shifts towardthe longer wavelengths with the time after excitation. The lumines-cence maximum observed at 630 nm for acquisition time 0–0.5 ls isshifted to 670 nm for acquisition time 6–10 ls. This suggests thattwo luminescence centers with different decay times contribute tothe emission spectrum of Sr2Si5N8:Eu2+

0.1 and the decay of theshort-wavelength component of the spectrum is faster than thelong-wavelength one.

Fig. 2c and d shows the decays profiles of luminescencemeasured at 10 K, monitored at 580 nm and 660 nm for

Fig. 2. Luminescence spectra of (a) Sr1.98Si5N8:Eu2+0.02 and (b) Sr1.9Si5N8:Eu2+

0.1 obtained under 460 nm pulse laser excitation and for different time acquisition. Luminescencedecays that correspond to the 4f65d1 ? 4f7 transition of (c) Sr1.98Si5N8:Eu2+

0.02 monitored luminescence at 580 nm and 660 nm and (d) Sr1.9Si5N8:Eu2+0.1 monitored luminescence at

600 nm and 710 nm.

736 A. Lazarowska et al. / Optical Materials 37 (2014) 734–739

Sr1.98Si5N8:Eu2+0.02 and monitored at 600 nm and 710 nm for

Sr1.9Si5N8:Eu2+0.1, respectively. Table 1 assembles calculated lumines-

cence decay times for both samples. The decays of the Sr1.98Si5N8:Eu2+

0.02 luminescence are single exponential with a time constantequal to 1.19 ls for the short-wavelength part of spectra and equalto 1.39 ls for the long-wavelength part. The value of 1.19 ls isregarded as a radiative decay time of the Eu2+ ion occupying Sr1position. The slight elongation of the decay time in long-wavelengthpart of the spectrum can be caused by presence of a small amount ofthe Eu2+ ions occupying the Sr2 sites. For the sample with 10 at.% ofEu2+, the Sr2 sites become more likely to be occupied by Eu2+ ions.This causes the red shift of the emission spectrum and elongationof the luminescence decay in the red spectral region. The decay ofthe short wavelength part of the Eu2+ emission in Sr1.9Si5N8:Eu2+

0.1

(corresponding to the Eu2+ occupying the Sr1 position) is non-expo-nential. After decomposition of the luminescence decay into twoexponents, one obtained the faster component equal to 0.3 ls andthe slower component equal to 1.19 ls (the same as in theSr1.98Si5N8:Eu2+

0.02 system). The existence of the fast component canbe caused by nonradiative energy transfer between Eu(Sr1) sitesand from the Eu(Sr1) to Eu(Sr2) sites. The luminescence decay ofthe long wavelength part of the emission spectrum of Sr2Si5N8:Eu2+

0.1

is single exponential and is characterized by the decay time equal to1,8 ls. This value can be regarded as a radiative decay time of theEu2+ ions in Sr2 position. The single exponential decay of the Eu(Sr2)

Table 1Luminescence decay time of Eu2+ in Sr2Si5N8.

Eu2+

concentrationShort-wavelength observationEu(Sr1)

Long-wavelengthobservation Eu(Sr2)

s (ls) s (ls)

2 at.% 1.19 1.3910 at.% 1.19 1.8

0.3

emission indicates lack of energy transfer from Eu(Sr2) to Eu(Sr1) at10 K.

To obtain temperature dependence of full width at half maxi-mum (FWHM) of the Eu(Sr1) and Eu(Sr2) emissions, the Sr1.9Si5N8:Eu2+

0.1 luminescence have been decomposed into two Gaussian bandsand presented in Fig. 1b. The energies of the maxima positions werefixed and the FWHM-s of the bands were free parameters in the fit-ting procedure. In Fig. 3 calculated bandwidths of the luminescencerelated to the Eu(Sr1) and Eu(Sr2) as a function of temperature arepresented. One notices that FWHM of the Eu(Sr1) emission of theSr1.9Si5N8:Eu2+

0.1 is in good agreement with Eu(Sr1) bandwidthobtained from the Sr1.98Si5N8:Eu2+

0.02 emission. We have used hyper-bolic cotangent law describing the temperature dependence of theFWHM of the emission band given by [16]:

FWHMðTÞ ¼ 2:36�hxffiffiffiSp

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffifficoth

�hx2kT

� �sð1Þ

where �hx is the effective phonon frequency and S is Huang–Rhysfactor. Quantity S�hx is the energy of the lattice relaxation.

The temperature dependence of the FWHM-s presented in Fig. 3has been fitted to the relation (1). The best fits (represented bysolid curves) were obtained for �hx = 454 cm�1 and S = 3 for Eu(Sr1)in the Sr1.98Si5N8:Eu2+

0.02, while for Eu(Sr1) in the Sr1.9Si5N8:Eu2+0.1 for

S = 3 and ⁄x = 459 cm�1. The Best fit for Eu(Sr2) was obtained forS = 5 and ⁄x = 476 cm�1. Obtained parameters allow to constructone dimensional configurational coordinate diagrams representingthe ground state and the excited 4f65d states of the Eu(Sr1) andEu(Sr2).

One noticed that luminescence of the Eu(Sr2) is shifted to thelower energy with respect to the Eu(Sr1) emission by quantity of946 cm�1. Since this difference is approximately equal to the dif-ference between electron lattice coupling of the 4f65d1 state ofEu(Sr2) (which is equal to 2380 cm�1) and Eu(Sr1) (which is equal

Fig. 3. Dependence of the emission bands width (FWHM) on temperature forSr1.98Si5N8:Eu2+

0.02 (blue squares) and Sr1.9Si5N8:Eu2+0.1 (red and orange squares). (For

interpretation of the references to color in this figure legend, the reader is referred tothe web version of this article.)

A. Lazarowska et al. / Optical Materials 37 (2014) 734–739 737

to 1377 cm�1) the minimum energy of the lowest 4f65d1 state, con-sidered as the energy of the zero phonon line, is approximately thesame in Eu(Sr2) and Eu(Sr1) sites. It means that the sum of the cen-troid shift and the crystal field splitting of the 4f65d1 state isapproximately the same in both sites.

The calculated Eu(Sr1) and Eu(Sr2) luminescence decay times atdifferent temperatures are presented in Fig. 4. Decay times are con-stant below 400 K and 200 K, for Eu(Sr1) and Eu(Sr2) respectively,while for higher temperatures the decay times become shorterwith increasing temperature. Luminescence decays of Eu(Sr1)observed for 10 at.% Eu2+ sample are non-exponential and are fas-ter than the decays of Eu(Sr1) luminescence observed for 2 at.%Eu2+ sample. The luminescence decays of Eu(Sr1) observed for10 at.% Eu2+ were considered as a superposition of two exponentsand temperature dependence of the longer one is presented inFig. 4.

In order to analyze the temperature shortening of the emissiondecay time s for both sites we discussed the data presented inFig. 4 considering formula [17]:

s ¼1þ

Pi exp Ei

kT

� �1srþP

iPi exp EikT

� � ð2Þ

in which sr is photoluminescence radiative lifetime, Ei are the acti-vation energies and Pi are the probabilities of the non-radiativedepopulations for i-th nonradiative pathway.

Fig. 4. Calculated Eu(Sr1) and Eu(Sr2) luminescence decay times at differenttemperatures from 10 K to 600 K range.

Fitting relation (2) to the experimental data we have estimatedsingle nonradiative pathway for Eu(Sr1) center in Sr1.98Si5N8:Eu2+

0.02,characterized by activation energy Ea1 = 3388 cm�1 and probabilityP1 = 3 ⁄ 109 s�1. For Eu(Sr2) in Sr1.9Si5N8:Eu2+

0.1 we had to use fittingdouble nonradiative pathways: the first characterized by probabilityP2 = 3 ⁄ 109 s�1 and energy Ea2 = 2739 cm�1 and the second charac-terized by the energy Ea = 507 cm�1 and probability P = 8 ⁄ 105 s�1.

The pathways of the Eu(Sr1) and Eu(Sr2) luminescence quench-ing, characterized by probability P1 = P2 = 3 ⁄ 109 s�1 and activationenergies equal to 3388 cm�1 and 2739 cm�1, respectively, can berelated to autoionization of the Eu2+ excited to the 4f65d1 state,considered as transition from the 4f65d1 to the conduction band.Autoionization of the Eu2+ creates Eu3+ and an electron in the con-duction band that can be described as Eu2+ ? Eu3+ + ecb transition.Since there is a Coulomb attraction between Eu3+ and ecb the sys-tem can exist in the bound state – Europium Trapped Exciton(ETE) state [18,19] which is energetically is located below the con-duction band. Therefore finally one can consider autoionizationalso as the transition from the 4f65d1 to ETE state.

The second nonradiative pathway describing the quenching ofthe Eu(Sr2) emission, characterized by energy Ea = 507 cm�1 andprobability P = 8 ⁄ 105 s�1 can be attributed to the nonradiativeenergy transfer from the Eu(Sr2) to Eu(Sr1) centers. Probability ofsuch transfer P is almost equal to probability of radiative transitionin the Eu(Sr1) center. Transfer of energy from the Eu(Sr2) toEu(Sr1) needs the activation energy Ea since this energy is neces-sary to overlap between the Eu(Sr2) emission and the Eu(Sr1)absorption.

For more detailed analysis of the non-radiative intersystemcrossing we have constructed configurational coordinate diagramsdescribing the luminescence properties of the Eu(Sr1) and Eu(Sr2)sites (Fig. 5a and b). In the first approximation we have assumedthat all electronic configurations can be represented by harmonicpotentials that yield parabolas in configurational coordinate dia-grams. Using the data obtained from the luminescence spectraand calculated quantity of S⁄x the diagrams representing theground states 4f7(8S7/2) and the first excited states 4f65d1 in bothsites have been constructed and presented in Fig. 5a and b. Onenotices that in both cases energy of crossing of the 4f7(8S7/2) and4f65d1 states are too high to allow the temperature activated theintersystem crossing 4f7(8S7/2) – 4f65d1. As the next step to explainthe thermal quenching of Eu2+ luminescence in Sr2Si5N8 we consid-ered the existence of auto-ionization. From the analysis presentedin Ref. [12] it is known that energy of the lowest 4f65d1 state ofEu(Sr1) is about 4000 cm�1 below the bottom of the conductionband in Sr2Si5N8. We assumed that this distance is similar in thecase of Eu(Sr2) system. Using these data we have presented inFig. 5a and b the curves labeled as Eu3+ + ecb representing the ‘‘con-duction band’’ (actually representing the ionized Eu2+ (Eu3+) andelectron in the conduction band). We considered further that theETE state exists with energy higher than the 4f65d1 and lower thanenergy of the Eu3+ + ecb system. To not influence the luminescence,the minimum energy of the ETE electronic manifold should beplaced significantly above the minimum energy of the 4f65d1 elec-tronic manifold. Since we have not the certain information aboutthis energy we made arbitrary but reasonably assumption thatETE is placed not higher than 1000 cm�1 above the minimumenergy of the 4f65d electronic manifold. This assumption yieldsthat the bounding energy of the ETE is at most 3000 cm�1. Thusthe ETE is placed more than 3000 cm�1 below the conduction band(Eu3+ + ecb) state.

The (Eu3+ + ecb) state and ETE state are characterized by muchlarger coupling with lattice than the 4f65d1 state thus electron–lat-tice coupling energy is estimated to be S⁄x = 5000–7000 cm�1. Weconsidered that respective S⁄x is equal to 6400 cm�1 and is thesame for ETE state and for (Eu3+ + ecb) state.

Fig. 5. Configurational coordinate diagrams representing the Eu(Sr1) (a) and Eu(Sr2) (b) in Sr2Si5N8: Eu2+0.02. The individual curves representing Eu2+ ions in the ground state

8S7/2, excited state 4f65d, ionization state Eu3+ + ecb and ETE state are indicated by respective labels.

738 A. Lazarowska et al. / Optical Materials 37 (2014) 734–739

Considering above assumptions we present the ETE states onconfiguration coordination diagrams by dashed parabolas labeledas ETE in Fig. 5(a) and (b). One can consider that after autoioniza-tion the relaxation takes place first to the ETE state and then finallyto the ground state, due to the intersystem crossing between ETEand 4f7(8S7/2). The activation energies for the intersystem4f7(8S7/2) – ETE crossing for Eu(Sr1) and Eu(Sr2) obtain using dia-gram are marked as E1 in Fig. 5(a) and E2 in Fig. 5(b). One noticesthat still these activation energies are much larger then calculatedfrom formula (2).

The cases when temperature luminescence quenching yieldsthe activation energy much smaller than one can obtained fromconfigurational coordinate diagrams has been already studied[17,20]. Proposed solution is based on the fact that ions motionin solids is confined. As the result, the electronic energy increasefaster than in harmonic approximation. Specifically this effectinfluences of the ground state. To visualize the problem and to per-form qualitative calculations the model potential has been used:

VðQÞ ¼ 0:5 aða� 1Þ½ �1=2 tan2fQ ½aða� 1Þ��1=4g ð3Þ

where coordinate Q is related to ions displacement R,Q ¼ R p

A

� �aða� 1Þ½ �1=4. Coefficients a and A is the strength constant

and size of confinement, respectively. Coefficients a and A aremutually dependent,

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiaða� 1Þ

p¼ kA2

p2�hxð4Þ

where �hx and k are phonon energy and elastic constant of thematrix, respectively.

Potential (3) yields the discrete energy spectrum [17,20]:

En ¼ a a� 1ð Þ½ ��12

n2

2þ a nþ 1

2

� � ð5Þ

Assuming that displacement cannot be greater than ion-ion dis-tance (�5 Å) we end up with estimation of a � 102. For these quan-tities energies of a few lowest vibronic states are almost the sameas those of the harmonic potential Vh ¼ 1

2 Q . However for quantumnumber n larger than 20, the energy spectrum differs from the har-monic one. The distance between neighboring vibronic levelsincreases with increasing n.

In Fig. 5(a) and (b) anharmonic potentials representing theground electronic manifolds of Eu(Sr1) and Eu(Sr2) sites are repre-sented by dotted curves. One notices that anharmonicity of the

system diminishes significantly the activation energy for intersys-tem ETE – 4f7(8S7/2) crossing, and thus explains the observed tem-perature luminescence quenching. These energies are shown inFig. 5(a) and in Fig. 5(b) as Ea1 and Ea2, respectively.

Considering spectroscopic data of Sr2Si5N8:Eu2+ we can alsomake prediction concerning the energies of the ground states ofEu2+ ions (in both sites) in relation to the bands edges. As ithas been mentioned Sr–N average distance in the Sr2 site issmaller than respective distance in the Sr1 site. According tothe existing models [19,21] energy of the ground state of theEu2+ should be higher when the distance between Eu2+ andligands is smaller. Therefore the ground state energy of theEu2+ in Sr2 site should be higher than the energy of the groundstate of Eu2+ in the Sr1 site. This conclusion is in full accordancewith the results of this paper. We noticed that the Sr2 site is notoccupied for low Eu2+ concentration and become occupied whenEu2+ concentration is sufficiently large. Such effect can take placewhen energy of the Eu(Sr2) system is higher than energy of theEu(Sr1) system.

4. Conclusions

The Eu2+ ions in Sr2Si5N8 can occupy the two available Sr2+ sites:tenfold coordinated Sr1 sites and eightfold coordinated Sr2 sites.Since Eu2+ in Sr1 site is tenfold coordinated thus it is characterizedby lower crystal field splitting and electron–lattice couplingstrength than Eu2+ in eight fold coordinated Sr2 site thus theenergy of the emission related to the 4f65d1 ? 4f7 transition ishigher in the case of Eu(Sr1) site.

Considering the spectroscopic data we concluded that theground state energy of the Eu2+ in Sr2 site should be higher thanthe ground state energy of Eu2+ in the Sr1 site. This explainsuneven distribution of Eu2+ ions in Sr1 and Sr2 positions at differ-ent level of Eu2+ doping in Sr2Si5N8 lattice. For low concentration ofEu2+ only Sr1 site is occupied, whereas for higher concentrationboth Sr1 and Sr2 sites are occupied. The experiments performedat temperatures above 10 K show that the luminescence intensityand decay time decreases with increasing temperatures, howeverthe decrease of the Eu(Sr2) luminescence decay and intensity takesplace at lower temperatures than Eu(Sr1) luminescence.

To describe the thermal luminescence quenching and lifetimediminishing we considered thermally activated ionization of Eu2+

ions. After autoionization the system can exist in the ETE state

A. Lazarowska et al. / Optical Materials 37 (2014) 734–739 739

which energetically is located between the lowest 4f65d1 and theconduction band. Therefore one can consider that the relaxationafter autoionization takes place first to the ETE state and thenfinally to the ground state, due to the intersystem crossingbetween ETE and 4f7(8S7/2). If one uses the standard model ofharmonic potential with linear electron–lattice coupling parame-terized to be consistent with the spectroscopic data, one cannotreproduce the experimental temperature dependence of theemission decay time. We have considered anharmonic modelpotential which assume that the ion motion in a crystal is con-fined. Since the confined potential is steeper, it results in thedecrease of the energy barrier of nonradiative process and thusexplains the observed temperature luminescence quenching.

Acknowledgements

This study has been funded by the European Social Found andby POIG.01.01.02-02-006/09 project ‘‘New efficient phosphors forlighting and solar concentrators’’. The contribution of A.L. wassupported within the International PhD Project ‘‘Physics of futurequantum-based information technologies’’: grant MPD/2009-3/4from Foundation for Polish Science and by the project ‘‘InnoDok-torant – Scholarships for PhD students, VIth edition’’ co-financedby the European Union in the frame of the European Social Fund.R.S. Liu would like to thank the support from National ScienceCouncil with the contract number of NSC101-2113-M-002-014-MY3.

References

[1] R.J. Xie, N. Hirosaki, Silicon Sci. Technol. Adv. Mater. 8 (2007) 588.[2] H.T. Hintzen, J.W.H. Van Krevel, G. Botty, Patent EP110479920010606.[3] X. Piao, T. Horikawa, H. Hanzawa, K. Machida, Appl. Phys. Lett. 88 (2006)

161908.[4] R.J. Xie, N. Hirosaki, S. Takayuki, F.F. Xu, M. Mitomo, Chem. Mater. 18 (2006)

5578.[5] V. Bachmann, C. Ronda, O. Oeckler, W. Schnick, A. Meijerink, Chem. Mater. 21

(2009) 316.[6] C.W. Yeh, W.T. Chen, R.S. Liu, S.F. Hu, H.S. Sheu, J.M. Chen, H.T. Hintzen, J. Am.

Chem. Soc. 134 (2012) 14108.[7] Y.Q. Li, G. De With, H.T. Hintzen, J. State Chem. 181 (2008) 515.[8] T. Schlieper, W. Milius, W. Schnick, Z. Anorg. Allg. Chem. 621 (1995) 1380.[9] Y.Q. Li, J.E.J. van Steen, J.W.H. van Krevel, G. Botty, A.C.A. Delsing, F.J. DiSavlo, G.

de With, H.T. Hintzen, J. Alloys Compd. 417 (2006) 273.[10] K.S. Sohn, S. Lee, R.J. Xie, N. Hirosaki, Appl. Phys. Lett. 95 (2009) 121903.[11] X. Piao, T. Horikawa, H. Hanzawa, K. Machida, J. Electochem. Soc. 153 (2006)

H232.[12] O.M. ten Kate, Z. Zhang, P. Dorenbos, H.T. Hintzen, E. van der Kolk, J. Solid State

Chem. 197 (2013) 209.[13] E. van der Kolk, T. Dorenbos, J.T.M. de Haas, C.W.E. van Eijk, Phys. Rev. B 71

(2005) 045121.[14] V. Bachmann, T. Justel, A. Meijerink, C. Ronda, P.J. Schmidt, J. Lumin. 121

(2006) 441.[15] A.A. Kubicki, P. Bojarski, M. Grinberg, M. Sadownik, B. Kuklinski, Opt. Commun.

269 (2006) 275.[16] B. Henderson, G.F. Imbush, Optical Spectroscopy of Inorganic Solids, Clarendon

Press, Oxford, 1989.[17] M. Grinberg, W. Jaskólski, Phys. Rev. B 55 (1997) 5581.[18] M. Grinberg, J. Lumin. 131 (2011) 433.[19] M. Grinberg, S. Mahlik, J. Non Cryst. Solids 354 (2008) 4163.[20] M. Grinberg, W. Jaskólski, C. Koepke, J. Planelles, M. Janowicz, Phys. Rev. B 50

(1994) 6504.[21] P. Dorenbos, Phys. Rev. B 85 (2012) 165107.