eu3+ and ce3+ co-doped aluminosilicate glasses and...

6
Eu 3þ and Ce 3þ co-doped aluminosilicate glasses and transparent glass-ceramics containing gahnite nanocrystals M. Reza Dousti a, b , Atiar R. Molla c, d , Ana Candida M. Rodrigues c , Andrea S.S. de Camargo a, * a S~ ao Carlos Physics Institute, University of S~ ao Paulo, S~ ao Carlos, SP, 13566-590, Brazil b Physics Institute, Federal University of Alagoas, Macei o, AL, 57072-900, Brazil c Department of Materials Engineering, Federal University of S~ ao Carlos, S~ ao Carlos, SP, 13565-905, Brazil d CSIR e Central Glass & Ceramic Research Institute, Kolkata, 700032, India article info Article history: Received 6 March 2017 Received in revised form 7 April 2017 Accepted 23 April 2017 Keywords: Aluminosilicate glass and glass-ceramics Eu 3þ Ce 3þ Gahnite nanocrystals abstract Transparent zinc-aluminosilicate glass-ceramics containing cerium and europium ions were prepared by controlled thermal heating of parent glasses. Addition of CeO 2 resulted in the improved transparency of the glasses in the visible spectral region. X-ray diffraction patterns of the glass-ceramics indicate the formation of the crystalline gahnite phase and there are evidences of Ce 3þ and Eu 3þ occupancies in this phase, as well as co-existence in the amorphous phase. Ce 3þ emission corresponding to the allowed f- d transitions is identied, as well as an anomalous emission in the red-infrared region. The characteristic luminescence of Eu 3þ in the red is observed. Judd-Ofelt analysis of Eu 3þ -doped samples reveals enhanced site asymmetry around the ion and a high branching ratio for the 612 nm emission after ceramization. Due to enhanced Ce 3þ emission and Eu 3þ emission quenching, color tunability is possible in the red to blue spectral region, depending on the heat treatment duration. © 2017 Elsevier B.V. All rights reserved. 1. Introduction Glasses and glass-ceramics containing rare earth ions are widely studied for their ability to host optically active trivalent lanthanide ions with various luminescent and optical applications. Besides optimized spectroscopic characteristics, other pursued features in these materials are high chemical durability and thermal stability, large rare earth solubility, etc. Particularly, transparent glass- ceramics offer the perspective of improved mechanical properties and higher absorption and emission cross sections, over a large spectral range, provided that the rare earth dopant occupies crys- talline sites, while still maintaining the transparency required for applications as in solid state lasers, ampliers, sensors, etc. [1]. Aluminosilicate glass and glass-ceramics are among the most investigated systems due to their high chemical durability and thermal stability, as well as very low thermal expansion coefcient [2,3]. Besides technological applications, aluminosilicate glass- ceramics having additional components are also commercialized in the form of cooktop panels, cookware, tile glazes, etc [3e5]. Despite the fact that silicate glasses present relatively high maximum phonon energy (~1000 cm 1 [6]), and low rare earth solubility [7], these glasses still occupy the top of the list in appli- cations due to cost-effective production and wide transparency in the UV-VIS-NIR spectral range. Addition of ZnO to aluminosilicate glasses can lower the nal product cost while the addition of controlled amount of TiO 2 induces the desired bulk crystallization [8]. In the preparation of glass-ceramics containing rare earth ions, the main challenge lies in the control of nucleation and crystal growth mechanisms so that the transparency of the parent glass is preserved. As discussed in Rayleigh-Gans theory [1,9], the attenu- ation of light by the crystallites can be minimized by making sure that their average size is smaller than the wavelength of incident light. Therefore, crystals with average size 15 nm and refractive index difference between the crystalline and glassy phases of about 0.1, are suggested in order to obtain better transparency. Beall and Pinckney [10] reported that crystal sizes of about 30 nm and a refractive index difference of about 0.3 could also be suitable if the distance between crystals remains lower than six times their average size (d < 180 nm). Transparent aluminosilicate glass- * Corresponding author. E-mail address: [email protected] (A.S.S. de Camargo). Contents lists available at ScienceDirect Optical Materials journal homepage: www.elsevier.com/locate/optmat http://dx.doi.org/10.1016/j.optmat.2017.04.054 0925-3467/© 2017 Elsevier B.V. All rights reserved. Optical Materials 69 (2017) 372e377

Upload: others

Post on 26-Jan-2021

15 views

Category:

Documents


0 download

TRANSCRIPT

  • lable at ScienceDirect

    Optical Materials 69 (2017) 372e377

    Contents lists avai

    Optical Materials

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

    Eu3þ and Ce3þ co-doped aluminosilicate glasses and transparentglass-ceramics containing gahnite nanocrystals

    M. Reza Dousti a, b, Atiar R. Molla c, d, Ana Candida M. Rodrigues c,Andrea S.S. de Camargo a, *

    a S~ao Carlos Physics Institute, University of S~ao Paulo, S~ao Carlos, SP, 13566-590, Brazilb Physics Institute, Federal University of Alagoas, Macei�o, AL, 57072-900, Brazilc Department of Materials Engineering, Federal University of S~ao Carlos, S~ao Carlos, SP, 13565-905, Brazild CSIR e Central Glass & Ceramic Research Institute, Kolkata, 700032, India

    a r t i c l e i n f o

    Article history:Received 6 March 2017Received in revised form7 April 2017Accepted 23 April 2017

    Keywords:Aluminosilicate glass and glass-ceramicsEu3þ

    Ce3þ

    Gahnite nanocrystals

    * Corresponding author.E-mail address: [email protected] (A.S.S. de Ca

    http://dx.doi.org/10.1016/j.optmat.2017.04.0540925-3467/© 2017 Elsevier B.V. All rights reserved.

    a b s t r a c t

    Transparent zinc-aluminosilicate glass-ceramics containing cerium and europium ions were prepared bycontrolled thermal heating of parent glasses. Addition of CeO2 resulted in the improved transparency ofthe glasses in the visible spectral region. X-ray diffraction patterns of the glass-ceramics indicate theformation of the crystalline gahnite phase and there are evidences of Ce3þ and Eu3þ occupancies in thisphase, as well as co-existence in the amorphous phase. Ce3þ emission corresponding to the allowed f-d transitions is identified, as well as an anomalous emission in the red-infrared region. The characteristicluminescence of Eu3þ in the red is observed. Judd-Ofelt analysis of Eu3þ-doped samples reveals enhancedsite asymmetry around the ion and a high branching ratio for the 612 nm emission after ceramization.Due to enhanced Ce3þ emission and Eu3þ emission quenching, color tunability is possible in the red toblue spectral region, depending on the heat treatment duration.

    © 2017 Elsevier B.V. All rights reserved.

    1. Introduction

    Glasses and glass-ceramics containing rare earth ions arewidelystudied for their ability to host optically active trivalent lanthanideions with various luminescent and optical applications. Besidesoptimized spectroscopic characteristics, other pursued features inthese materials are high chemical durability and thermal stability,large rare earth solubility, etc. Particularly, transparent glass-ceramics offer the perspective of improved mechanical propertiesand higher absorption and emission cross sections, over a largespectral range, provided that the rare earth dopant occupies crys-talline sites, while still maintaining the transparency required forapplications as in solid state lasers, amplifiers, sensors, etc. [1].Aluminosilicate glass and glass-ceramics are among the mostinvestigated systems due to their high chemical durability andthermal stability, as well as very low thermal expansion coefficient[2,3]. Besides technological applications, aluminosilicate glass-ceramics having additional components are also commercialized

    margo).

    in the form of cooktop panels, cookware, tile glazes, etc [3e5].Despite the fact that silicate glasses present relatively highmaximum phonon energy (~1000 cm�1 [6]), and low rare earthsolubility [7], these glasses still occupy the top of the list in appli-cations due to cost-effective production and wide transparency inthe UV-VIS-NIR spectral range. Addition of ZnO to aluminosilicateglasses can lower the final product cost while the addition ofcontrolled amount of TiO2 induces the desired bulk crystallization[8].

    In the preparation of glass-ceramics containing rare earth ions,the main challenge lies in the control of nucleation and crystalgrowth mechanisms so that the transparency of the parent glass ispreserved. As discussed in Rayleigh-Gans theory [1,9], the attenu-ation of light by the crystallites can be minimized by making surethat their average size is smaller than the wavelength of incidentlight. Therefore, crystals with average size 15 nm and refractiveindex difference between the crystalline and glassy phases of about0.1, are suggested in order to obtain better transparency. Beall andPinckney [10] reported that crystal sizes of about 30 nm and arefractive index difference of about 0.3 could also be suitable if thedistance between crystals remains lower than six times theiraverage size (d < 180 nm). Transparent aluminosilicate glass-

    mailto:[email protected]://crossmark.crossref.org/dialog/?doi=10.1016/j.optmat.2017.04.054&domain=pdfwww.sciencedirect.com/science/journal/09253467www.elsevier.com/locate/optmathttp://dx.doi.org/10.1016/j.optmat.2017.04.054http://dx.doi.org/10.1016/j.optmat.2017.04.054http://dx.doi.org/10.1016/j.optmat.2017.04.054

  • M.R. Dousti et al. / Optical Materials 69 (2017) 372e377 373

    ceramics with composition SiO2eAl2O3eMgOeZnOeZrO2 con-taining b-quartz crystals of about 10 mm have also been reported[10], and most recently we reported glass and transparent glass-ceramics in the compositional system SiO2eAl2O3eZnOeTiO2eK2OeCeO2 [11]. In this case, CeO2 was used toeliminate the dark coloration introduced by Ti3þ ions in the glasses.Upon heating, the CeO2 dopant is reduced to the hexagonal ses-quioxide Ce2O3 and the residual oxygen participates in the oxida-tion of Ti3þ to Ti4þ yielding pale yellow colored glasses. Taking afurther step, controlled crystallization of the glasses resulted intransparent glass-ceramics, which could be interesting host for rareearth ions.

    In the light of possible optical applications, we have preparedand fully characterized the spectroscopic properties of threedifferent series of glasses and glass ceramics, one of which dopedwith Eu3þ ions. In our investigations, we firstly tackle the effect ofCeO2 agent to increase the transparency of the parent glass, andthen study the luminescence dynamics of different species in rareearth doped glass and glass-ceramics.

    2. Experimental

    In previous work [11], we prepared glass and glass ceramicswith wt% compositions:

    Gn1: 50.6SiO2 þ 23.0Al2O3 þ 18.4ZnO þ 5TiO2 þ 3K2O, and

    Gn2: 50.6SiO2 þ 23.0Al2O3 þ 18.4ZnO þ 5TiO2 þ 3K2O þ 0.5 wt%CeO2 in excess

    The parent glasses were prepared by the conventional melt-quenching technique and the derived glass-ceramics were ob-tained by controlled thermal processing of the glasses to inducecrystal nucleation and growth. Having previously verified that Ce4þ

    could eliminate the brown color of Gn1 glass, for this work weprepared a glass co-doped with Ce3þ and Eu3þ (Gn3:[50.6SiO2 þ 23.0Al2O3 þ 18.4ZnO þ 5TiO2 þ 3K2O] þ 0.5CeO2 wt% þ 0.5 wt% Eu2O3, both in excess), aiming to investigate theluminescence properties of the rare-earth ions in the glass andglass-ceramic.

    The high purity precursors SiO2 (99%, Vitrovita, Brazil), Al2O3(99%, Almatis, USA), ZnO (99.7%, J. T. Baker, USA), TiO2 (99.5%, RiedelDe Haen, Germany), anhydrous K2CO3 (99%, Nuclear, Brazil), CeO2(99.0%, Vetec, Brazil) and Eu2O3 (99.99%, Aldrich, USA), were ho-mogeneously mixed and melted in a Pt crucible at 1620 �C for 4 h.The glass melts were quenched at room temperature, onto a pre-heated cast ironmold, and subsequently annealed at 600 �C for 12 hto relieve internal stresses. Differential scanning calorimetry (DSC)studies of the powdered and bulk samples were performed in orderto determine the characteristic glass transition temperature (Tg),crystallization on-set and peak temperature (Tp), using a calorim-eter model STA 449 F3 of NETZSCH-Ger€atebau GmbH, Selb, Ger-many. In order to obtain glass-ceramics, the glass samples weresubjected to thermal processing at 670 �C for 3 h to induce nucle-ation and at 880 �C for various periods of time (2, 25, 50 and 100 h)to promote the growth of the crystalline phase. Samples are namedfollowing the duration (in hours) of the final heat-treatment at880 �C, for example, Gn2_2, Gn2_25, etc.

    To identify the crystalline phases developed in the resultingglass-ceramics, X-ray diffraction (XRD) data were recorded in aRigaku Ultima IV X-ray diffractometer (Tokyo, Japan), equippedwith a 40 kV power supply and 20 mA output of CuKa radiation(1.5406 Å). Scanning was carried out in the 2q range 10 to 80�, witha step of 0.02� (2q). Transmission electron micrographs werecaptured using a Tecnai G2 (FEI Company, Hillsboro, OR) TEM

    machine, operating at 200 kV.The UVeVis spectra of the samples were measured using a

    Perkin-Elmer Lambda 25 spectrophotometer with 1 nm resolution.The photoluminescence spectrawere recorded using a Horiba JobinYvon spectrofluorimeter model Fluorolog FL-1057, equipped withCW and pulsed Xe lamps. The visible signals were collected by asingle photo-counting detector (PPD-850) at 60� with respect tothe excitation beam. The excitation spectra were corrected by thelamp profile and detector's response. Excited state lifetime valuesof Eu3þ 5D0 level were obtained from the fit of emission decaycurves at 612 nm (excited at 393 nm), using the decay-by-delayphosphorimetric analysis in the same fluorimeter. In the case ofCe3þ ions, the lifetime values were obtained by single exponentiallyfitting the emission decay curves at 360 nm measured using theTCSPC mode, with excitation at 289 nm by a Delta Diode.

    3. Results and discussion

    The optimization of the nucleation and crystallization proced-ures, by adjustment of temperature and heat treatment time is ofchief importance to obtain glass-ceramics with controlled volumefraction of crystalline phases and crystallite sizes. In order tonucleate the crystal centers and to allow consequent growth, it isnecessary to know the glass characteristics temperatures, Tg, Tp, andcrystallization behavior of the parent glass. The crystallizationbehavior of Gn1 and Gn2 glasses, were previously discussed [11].The DSC (10 K/min) glass transition temperature of Gn1 and Gn2glass samples are 710 and 680 �C, respectively while the crystalli-zation onset temperatures (Tx) are at about 833 and 813 �C, and thecrystallization peaks (TP) are around 868 and 850 �C, respectively.We argue that the decrease in the characteristic temperatures fromGn1 to Gn2 should be related to the incorporation of Ce3þ or to thereduction of Ce4þ to Ce3þ ions in the Gn2 glass, which can interferewith the glass structure and facilitate devitrification. By comparingthe DSC curves of Gn1 bulk and powder samples, it was observed[11] that bulk crystallization occurs.

    The X-ray diffraction patterns of the Gn2 parent glass and itsderived glass-ceramics are given in Ref. [11]. It is observed thatthere is an increase in crystallization up to 25 h of heat treatment,but no further development for longer heat treatment times. Still,the broad humps in the XRD patterns in 2q range of 10e40� evi-dence the existing glassy phase even after 100 h of ceramization.The XRD scattering peaks of Gn2_x, and Gn3_x (x ¼ 2, 25, 50 and100 h) samples confirm the formation of crystalline cubic ZnAl2O4spinel phase (PDF file no. 71-968), known as gahnite, with no traceof any other phase. The average crystal size of gahnite graduallyincreases from 4.1 to 5.8, 6.6 and 6.8 nm, for 2, 25, 50 and 100 h ofheat-treatments, as previously reported [11]. Representative TEMmicrographs of samples Gn1_x heat treated for 2 and 100 h at880 �C after nucleation at 670 �C for 3 h, are given in Fig. 1.

    As previously discussed, the addition of CeO2 to the Gn1 glasssamples results in more transparent yellowish glasses (Gn2) due tothe oxidation of Ti3þ to Ti4þ induced by the reduction of Ce4þ ionsto Ce3þ. Fig. 2 shows the optical transmission spectra of the Gn1,Gn2 and Gn3 glasses and glass-ceramics, in the 250e1100 nmspectral region. For the Gn3_100 sample, which were nucleated at670 �C for 3 h and then heat treated at 880 �C for 100 h, the ab-sorption bands of Eu3þ ions are not observable due to the lowdopant concentration and large host absorption in the UVeVisrange. The inset of Fig. 2 shows the photographs of the samples. Asit can be seen by the photographs and spectra, the Gn3 glass is moretransparent than Gn1 and Gn2 glasses and glass-ceramics. How-ever, ceramization further decreases the transparency of both Gn2and Gn3 series. The fundamental absorption edge of the Gn1 glassat about 3.70 eV shifts to 3.64 eV and 3.60 eV after incorporation of

  • Fig. 1. TEM micrographs of (a) Gn1_2 and (b) Gn1_100. Samples were heat-treated,respectively for 2 and 100 h at 880 �C, after being nucleated at 670 �C for 3 h.

    Fig. 2. UVeViseNIR transmission spectra of parent glasses and glass-ceramics whichwere normalized by sample thickness. The inset shows the photographs of the Gn1,Gn2 and Gn3 samples (from top to bottom) untreated and heat-treated at 880 �C for 2,25, 50 and 100 h (after previous nucleation treatment at 670 �C for 3 h) (from left toright).

    Fig. 3. Normalized luminescence spectra of glassy samples Gn1 and Gn2 and of Gn2glass ceramics excited at 314 nm (a) and at 393 nm (b). (c) Representative excitationspectra for emissions centered at 350 and 700 nm.

    M.R. Dousti et al. / Optical Materials 69 (2017) 372e377374

    CeO2 and further ceramization up to 100 h, respectively.Fig. 3(a) illustrates the luminescence spectra of the Gn1 and Gn2

    glasses and Gn2 glass-ceramics, excited at 314 nm. The spectrum ofGn1 glass consists of a broadband at around 550 nm which isattributed to the emission from Ti3þ ions [11]. The emission spectraof the Gn2 glass and the derived glass-ceramics shows a majorbroadband component at 350 nm while Ti3þ band is present inmuch lower intensity, confirming the oxidation mechanism pro-posed in Ref. [11]. The broadband around 350 nm is attributed tothe characteristic radiative d-f transitions of Ce3þ. It is interesting tonotice that in going from Gn2 glass to Gn2 glass ceramics a pro-gressive red shift is observed, indicating a change in site symmetryaround the dopant, because the 5d excited state of Ce3þ is highlylocalized and strongly influenced by the crystal field imposed bythe first coordination sphere ligands. A similar red spectral shift has

    also been reported in YAG:Ce nanoparticles containing acetatehydrate due to the increase in Gd3þ concentration and crystal size[12]. In our case, it is likely that the red shift for the Gn2 glass

  • M.R. Dousti et al. / Optical Materials 69 (2017) 372e377 375

    ceramics is associated to the incorporation of Ce3þ in the gahnitecrystalline phase. As a control experiment, for posterior compari-son with Eu3þ-doped samples, we have also measured lumines-cence with excitation at 393 nm as it is seen in Fig. 3(b). Theresulting spectra consist of a band in the range 400e600 nm and amore structured band around 700 nm, which is not observed forthe glassy Gn2 sample. In fact, by normalizing the spectra to thislatter band, one can see that the intensity of the higher energy bandprogressively decreases with increasing heat treatment time. Thisobservation gives strong evidence of Ce3þ ions in two differentenvironments (glassy and crystalline). It is worth mentioning thatfor both excitations at 314 and 393 nm, the spectral features wereconfirmed by measurements using a long pass filter at 500 nm.

    The excitation spectra measured by monitoring the lumines-cence in the range 350e400 nm, and at 700 nmwere investigated.The results obtained by monitoring the UV-blue band were prac-tically identical, so representative curves (lem¼ 350 nm) are shownin Fig. 3(c), along with spectra monitored at 700 nm for the glassyGn2 and representative glass ceramic Gn2_50. In both cases, thereis a characteristic excitation band at 314 nm, which corresponds tothe f/d transitions of Ce3þ ions. However, the excitation spectrafor the 700 nm emission of glass ceramics differs from that of Gn2glass. It presents two additional bands centered at around 400 and540 nm, which, in agreement with the emission spectra, might beassociated to Ce3þ ions in the gahnite phase. In addition, for both,Gn2 and Gn2_50 an excitation band below 280 nm is perceivedwhen the emission is monitored at 700 nm. Further investigation ofthe origin of this bands were done by measuring emission spectraat 400 nm and at 540 nm. As expected, the spectra in Fig. 3(b) werereproduced for lexc ¼ 400 nm but, interestingly, no emission at allwas observed for lexc ¼ 550 nm, and so far we have no explanationfor this observation.

    The luminescence and excitation spectra of Gn3 samples arepresented in Fig. 4. The characteristic emission lines of Eu3þ ionsexcited at 393 nm are observed at 577, 590, 612, 652, 700 and744 nm and are attributed to transitions from the 5D0 excited stateto the lower energy states 7FJ¼0,1,2,3,4,5, respectively. There is anoverall decrease of intensity of these lines when the Gn3 glass isheat treated for 2, 25, 50 and 100 h to yield glass ceramics, while,concomitantly, an emission band centered at 450 nm emerges. In

    Fig. 4. Excitation and emission spectra of the Eu3þ-doped Gn3 glass and glass-ceramics. Glass-ceramics were obtained by crystallization of the precursor Gn3 glassat 880 �C for 2, 25, 50 and 100 h (after previous nucleation treatment at 670 �C for 3 h).

    accordance with Fig. 3(b) this band is attributed to Ce3þ ion f/dtransition. The excitation spectra of the Eu3þ ions doped glass andglass-ceramics for the emission centered at about 612 nm is alsoshown in Fig. 4. Both spectra show identical profile of the Eu3þ

    excited state bands, displaying spectral lines that correspond to thetransitions 7F0/5D4 (360 nm), 7F0/5L7 (380 nm), 7F0/5L6(393 nm), 7F0,1 / 5D3 (413 nm), 7F0/5D2 (463 nm), 7F0/5D1(524 nm), 7F1/5D1 (531 nm), 7F0/5D0 (577 nm) and 7F1/5D0(585 nm). For wavelengths below 350 nm a broad band isperceived, which is likely to be due to the host matrix absorption.

    The Eu3þ luminescence decay curves monitored at 612 nmwithexcitation at 393 nm are presented in Fig. 5(a). The decays werefitted to single exponential functions to yield average lifetimevalues as presented in Table 1. As it can be seen, the lifetime ofEu3þ-doped Gn3 glass sample is the highest (1.77 ms) and it pro-gressively decreases to a maximum of ~20% of this value as the heattreatment time increases to 100 h. The change in lifetime indicatesa change in the Eu3þ environment upon crystallization, eventhough no significant spectral change has been observed. Thepossibility of Eu3þ to Ce3þ energy transfer is not discarded, but itmust be further investigated in order to be confirmed.

    The luminescence decay curves were also measured at 360 nm,where Ce3þ emits, upon 289 nm excitation. The choice of theexcitation wavelength was made based on the availability of a fastpulsing delta diode, which allows usage of the TCSPC e TimeCorrelated Single Photon Counting technique to acquire lifetimes inthe nanosecond range. Fig. 5(b) presents the representative curvesfor Gn3 glass and Gn3_50 glass-ceramic together with the refer-ence curve (scattered excitation light). The average lifetime valuesare also presented in Table 1 for Gn2 and Gn3 samples. Similarly tothe case of Eu3þ, crystallization yields a decrease in the lifetimevalues of Ce3þ to a maximum of 50% in both cases. The Gn3 seriespresents overall values that are twice as higher as those for the Gn2samples. The larger variance in lifetime values is understandableconsidering the f-d character of the transition much more suscep-tible to differences in chemical environment of the Ce3þ ion incomparison to the f-f transitions of Eu3þ.

    The intensity ratio of Eu3þ emission bands at 612 and 590 nm(I612/590) is often used as a parameter to evaluate the symmetry ofthe Eu3þ sites [13,14]. As measured, I612/590 increases from 6.25 to7.44 by increasing the heat-treatment time of Gn3 glass ceramicsup to 100 h. Therefore, we can conclude that the asymmetry aroundthe Eu3þ ions sites increases by increasing the crystallinity of thesample. Finally, we have used the Judd-Ofelt approach to analyzethe radiative properties of the samples before and after ceramiza-tion. Rather than using the absorption spectra (with only a fewlines), for the calculation of the phenomenological intensity pa-rametersUl for the Gn3 and Gn3100h samples, our calculations werebased on themethod that employs the luminescence spectra [15] ofthese samples. This strategy minimizes the errors involved in theleast square fitting of oscillator strength values for the case oflimited number of transitions.

    The calculated, average, refractive index of the samples is 1.81for both Gn3 and Gn3_100 samples. It is worth noting that therefractive index of the glass-ceramic is, in fact, the average refrac-tive index of the residual glass and the obtained crystal. However,for the transparent glass-ceramics, the refractive index differencebetween glassy and crystalline phases should be as small as 0.1e0.3[1,9,10]. The radiative emission probabilities, branching ratio andJudd-Ofelt intensity parameters for the two representative samplesare given in Tables 2 and 3. The branching ratio of the 612 nmemission is slightly higher in the case of Gn3_100 (~69%) than inGn3 (~64%). Furthermore, as compared to other glasses, the b612 nmvalues in our studied glass and glass-ceramics is comparable to orhigher than previously reported values for germanate (60% [16]),

  • Fig. 5. Excited state decay curves of (a) Eu3þ(lexc ¼ 393 nm and lem ¼ 612 nm), and (b) Ce3þions (lexc ¼ 289 nm and lem ¼ 360 nm).

    Table 1Lifetime values of Eu3þ and Ce3þ ions doped into alumina-silicate glasses and glass-ceramics of the Gn2 and Gn3 series.

    Sample t Eu3þ (ms)lexc ¼ 393 nmlem ¼ 612 nm

    t Ce3þ (ns)lexc ¼ 290 nmlem ¼ 360 nm

    Gn2 e 3.12Gn22h e 1.84Gn225h e 1.62Gn250h e 1.61Gn2100h e 1.61Gn3 1.77 6.77Gn32h 1.57 5.98Gn325h 1.45 4.48Gn350h 1.44 3.64Gn3100h 1.44 3.60

    Table 2Calculated radiative parameters of two representative Eu3þ doped samples.

    Transitions5D0/

    Wavelength ±1(nm)

    Gn3 Gn3_100

    AJ/J'±0.01(s�1)

    a ±0.01(%)

    AJ/J'±0.01(s�1)

    a ±0.01(%)

    7F0 577 38.47 6.81 17.88 2.597F1 590 71.96 12.74 71.49 10.377F2 612 359.6 63.65 481.92 68.907F3 652 21.18 3.75 27.10 3.937F4 700 70.10 12.40 84.60 12.277F5 744 3.33 0.59 5.76 0.847F6 800 0.37 0.06 0.71 0.10

    Table 3Judd-Ofelt intensity parameters of Eu3þ ions in the two representative glass Gn3 andglass-ceramic Gn3100h samples, in comparison to the literature.

    Glass U2(10�20 cm2)

    U4(10�20 cm2)

    U6(10�20 cm2)

    Gn3 6.19 2.51 0.23Gn3_100 8.36 3.06 0.44Fluorosilicate glass [20] 7.44 0.72 eBorophosphate glass [21] 2.98 0.42 eTelluro-fluoroborate glass [22] 6.92 3.43 0Tellurite ceramics [23] 2.95 0.39 eLead-zinc phosphate glass [24] 6.52 2.29 0.98Sodium phosphate glass [18] 4.12 4.69 1.83Pentaphosphate crystal [18] 3.66 1.43 3.07

    M.R. Dousti et al. / Optical Materials 69 (2017) 372e377376

    boro-tellurite (60% [16]), ZBLA (10% [17]), pentaphosphate (61%[18]]), zirconium fluoride (24.2% [18]) and tungsten-phosphate(69% [15]) glasses. As a general rule, the emission bands withbranching ratios higher than 50% are known to be suitable for laseraction [19]. The increase in the values of U2 and U6, when goingfrom Gn3 to Gn3_100 indicates the increase in site asymmetry atthe vicinity of the Eu3þ ions and increased rigidity of the sample,respectively [13,14]. The values of U6 are less commonly reported inthe literature but our result is comparable to other glasses and thevalue ofU2 for Gn3_100 is higher than that for several compositionsas seen in Table 3.

    The chromaticity diagram (CIE 1931) is a useful tool to define thecolor purity of luminescent materials through two calculated di-mensions (x,y) [25]. The spectra of the Eu3þ-Ce3þ co-doped Gn3glass and Gn3100h glass-ceramic, presented in Fig. 4, were employedto calculate the chromaticity coordinates. The values are given inTable S1 (Suppl. Materials). Fig. 6 presents the CIE-1931 chroma-ticity diagram. As seen, under 393 nm excitation, the glassy sampleGn3 presents red color, which progressively evolves into the blueregion by increase of the heat-treatment time and consequent in-duction of the crystalline phase in Gn3_x (x¼ 0, 2, 25, 50 and 100 h)glass ceramics. Thus, the blue emission associated to Ce3þ ionssignificantly contributes to the emission of the glass ceramics, andthe emission can be tuned in the blue-red spectral range by se-lection of the heat-treatment duration. This tunability offers theperspective of combining colors in samples co-doped with greenemitting Tb3þ ions to generate white light.

    4. Conclusions

    Zinc-aluminosilicate glasses were prepared by the conventionalmelt quenching technique and subjected to controlled thermalprocessing to successfully obtain transparent glass ceramics con-taining nanosized gahnite phase. The addition of CeO2 to the parentglass, results in the improved transparency of the glasses in thevisible region through reduction of Ce4þ to Ce3þ concomitantlywith Ti3þ to Ti4þ oxidation. Upon ceramization, the samples havetheir transparency decreased in both, the singly CeO2-doped

  • Fig. 6. CIE 1931 chromaticity diagram of the Eu3þ-Ce3þ co-doped glass and glass-ceramics for Gn3_x with x ¼ 0, 2, 25, 50 and 100 h.

    M.R. Dousti et al. / Optical Materials 69 (2017) 372e377 377

    samples and CeO2/Eu2O3 co-doped samples. Emission of Ce3þ ionscorresponding to allowed f-d transition is identified for differentchemical environments (glassy and crystalline). An anomalous,additional emission is observed in the red-infrared region. Theluminescence of Eu3þ ions gradually quenches by increasing theheat-treatment times probably due to some interaction with Ce3þ

    ions - energy transfer has not been discarded. This hypothesis iscorroborated by decreased Eu3þ donor level lifetime values whilelarger values are measured for Ce3þ acceptor level, upon ceram-ization. The Judd-Ofelt analysis of Eu3þ-doped samples revealsenhanced site asymmetry around the Eu3þ ions and a highbranching ratio for the emission at 612 nm, after ceramization. Dueto enhanced Ce3þ emission and Eu3þ quenching, color tunability ispossible in the red to blue region, depending on heat treatmentduration.

    Acknowledgments

    Authors acknowledge the financial support of FAPESP throughCEPID process 2013/07793-6. ARM is especially thankful to thedirector of the CSIR-CGCRI, Kolkata for granting a leave of absenceto work at LaMaV, UFSCar, Brazil, and to FAPESP for the visitingprofessorship 2014/208659. MRD also thanks FAPESP for thepostdoctoral research grant 2013/24064-8.

    References

    [1] M. Clara Gonçalves, L.F. Santos, R.M. Almeida, Comptes Rendus Chim. 5 (2002)845e854.

    [2] H. Scheidler, E. Rodek, Am. Ce-Ram. Soc. Bull. 68 (1989) 1926e1930.[3] E.D. Zanotto, Am. Ceram. Soc. Bull. 89 (2010) 19e27.[4] C. Leonelli, T. Manfredini, G. Pellacani, J. AmorosAlbaro, J. Enrique Navarro,

    M. Jose Orts, S. Bruni, F. Cariati, J. Am. Ceram. Soc. 74 (1991) 983e987.[5] L.R. Pinckney, in: S.J. Schneider (Ed.), Eng. Mater. Handb., Jr, ASM International,

    Metals Park, OH, 1991, pp. 433e438.[6] C.B. Layne, W.H. Lowdermilk, M.J. Weber, Phys. Rev. B 16 (1977) 10e20.[7] A.J. Kenyon, Prog. Quantum Electron 26 (2002) 225e284.[8] L. Barbieri, C. Leonelli, T. Manfredini, C. Siligardi, A.B. Corradi, P. Mustarelli,

    C. Tomasi, J. Am. Ceram. Soc. 80 (2005) 3077e3083.[9] R.W. Hopper, J. Non. Cryst. Solids 70 (1985) 111e142.

    [10] G.H. Beall, L.R. Pinckney, J. Am. Ceram. Soc. 82 (2004) 5e16.[11] A.R. Molla, et al., Crystallization, mechanical and optical properties of trans-

    parent, nanocrystalline gahnite glass-ceramics, J. Am. Ceram. Soc. 00 (2017)1e13, http://dx.doi.org/10.1111/jace.14753.

    [12] B.A. Dong, J. Wang, J. Sun, S. Xu, X. Bai, Z.L. Jiang, L. Xia, L.K. Sun, H.W. Song,RSC Adv. 2 (2012) 3897e3905.

    [13] Proulx Capobianco, Negrisolo Bettinelli, Phys. Rev. 42 (1990) 5936e5944.[14] M. Nogami, N. Umehara, T. Hayakawa, Phys. Rev. B 58 (1998) 6166e6171.[15] M.R. Dousti, G.Y. Poirier, A.S.S. de Camargo, Opt. Mater 45 (2015) 185e190.[16] A. Jha, B. Richards, G. Jose, T. Teddy-Fernandez, P. Joshi, X. Jiang, J. Lousteau,

    Prog. Mater. Sci. 57 (2012) 1426e1491.[17] M. Dejneka, E. Snitzer, R. Riman, J. Lumin 65 (1995) 227e245.[18] B. Blanzat, L. Boehm, R. Reisfeld, N. Spector, J. Solid State Chem. 32 (1980)

    185e192.[19] R. Reisfeld, E. Greenberg, R.N. Brown, M.G. Drexhage, C.K. Jørgensen, Chem.

    Phys. Lett. 95 (1983) 91e94.[20] D. Ramachari, L. Rama Moorthy, C.K.K. Jayasankar, J. Lumin 143 (2013)

    674e679.[21] N. Kiran, J. Mol. Struct. 1065e1066 (2014) 93e98.[22] R. Vijayakumar, K. Maheshvaran, V. Sudarsan, K. Marimuthu, J. Lumin 154

    (2014) 160e167.[23] B.J. Chen, E.Y.B. Pun, H. Lin, J. Alloys Compd. 479 (2009) 352e356.[24] G.H. Silva, V. Anjos, M.J.V. Bell, A.P. Carmo, A.S. Pinheiro, N.O. Dantas, J. Lumin

    154 (2014) 294e297.[25] T. Smith, J. Guild, Trans. Opt. Soc. 33 (1931) 73e134.

    http://refhub.elsevier.com/S0925-3467(17)30275-6/sref1http://refhub.elsevier.com/S0925-3467(17)30275-6/sref1http://refhub.elsevier.com/S0925-3467(17)30275-6/sref1http://refhub.elsevier.com/S0925-3467(17)30275-6/sref2http://refhub.elsevier.com/S0925-3467(17)30275-6/sref2http://refhub.elsevier.com/S0925-3467(17)30275-6/sref3http://refhub.elsevier.com/S0925-3467(17)30275-6/sref3http://refhub.elsevier.com/S0925-3467(17)30275-6/sref4http://refhub.elsevier.com/S0925-3467(17)30275-6/sref4http://refhub.elsevier.com/S0925-3467(17)30275-6/sref4http://refhub.elsevier.com/S0925-3467(17)30275-6/sref5http://refhub.elsevier.com/S0925-3467(17)30275-6/sref5http://refhub.elsevier.com/S0925-3467(17)30275-6/sref5http://refhub.elsevier.com/S0925-3467(17)30275-6/sref6http://refhub.elsevier.com/S0925-3467(17)30275-6/sref6http://refhub.elsevier.com/S0925-3467(17)30275-6/sref7http://refhub.elsevier.com/S0925-3467(17)30275-6/sref7http://refhub.elsevier.com/S0925-3467(17)30275-6/sref8http://refhub.elsevier.com/S0925-3467(17)30275-6/sref8http://refhub.elsevier.com/S0925-3467(17)30275-6/sref8http://refhub.elsevier.com/S0925-3467(17)30275-6/sref9http://refhub.elsevier.com/S0925-3467(17)30275-6/sref9http://refhub.elsevier.com/S0925-3467(17)30275-6/sref10http://refhub.elsevier.com/S0925-3467(17)30275-6/sref10http://dx.doi.org/10.1111/jace.14753http://refhub.elsevier.com/S0925-3467(17)30275-6/sref12http://refhub.elsevier.com/S0925-3467(17)30275-6/sref12http://refhub.elsevier.com/S0925-3467(17)30275-6/sref12http://refhub.elsevier.com/S0925-3467(17)30275-6/sref13http://refhub.elsevier.com/S0925-3467(17)30275-6/sref13http://refhub.elsevier.com/S0925-3467(17)30275-6/sref14http://refhub.elsevier.com/S0925-3467(17)30275-6/sref14http://refhub.elsevier.com/S0925-3467(17)30275-6/sref15http://refhub.elsevier.com/S0925-3467(17)30275-6/sref15http://refhub.elsevier.com/S0925-3467(17)30275-6/sref16http://refhub.elsevier.com/S0925-3467(17)30275-6/sref16http://refhub.elsevier.com/S0925-3467(17)30275-6/sref16http://refhub.elsevier.com/S0925-3467(17)30275-6/sref17http://refhub.elsevier.com/S0925-3467(17)30275-6/sref17http://refhub.elsevier.com/S0925-3467(17)30275-6/sref18http://refhub.elsevier.com/S0925-3467(17)30275-6/sref18http://refhub.elsevier.com/S0925-3467(17)30275-6/sref18http://refhub.elsevier.com/S0925-3467(17)30275-6/sref19http://refhub.elsevier.com/S0925-3467(17)30275-6/sref19http://refhub.elsevier.com/S0925-3467(17)30275-6/sref19http://refhub.elsevier.com/S0925-3467(17)30275-6/sref19http://refhub.elsevier.com/S0925-3467(17)30275-6/sref20http://refhub.elsevier.com/S0925-3467(17)30275-6/sref20http://refhub.elsevier.com/S0925-3467(17)30275-6/sref20http://refhub.elsevier.com/S0925-3467(17)30275-6/sref21http://refhub.elsevier.com/S0925-3467(17)30275-6/sref21http://refhub.elsevier.com/S0925-3467(17)30275-6/sref21http://refhub.elsevier.com/S0925-3467(17)30275-6/sref22http://refhub.elsevier.com/S0925-3467(17)30275-6/sref22http://refhub.elsevier.com/S0925-3467(17)30275-6/sref22http://refhub.elsevier.com/S0925-3467(17)30275-6/sref23http://refhub.elsevier.com/S0925-3467(17)30275-6/sref23http://refhub.elsevier.com/S0925-3467(17)30275-6/sref24http://refhub.elsevier.com/S0925-3467(17)30275-6/sref24http://refhub.elsevier.com/S0925-3467(17)30275-6/sref24http://refhub.elsevier.com/S0925-3467(17)30275-6/sref25http://refhub.elsevier.com/S0925-3467(17)30275-6/sref25

    Eu3+ and Ce3+ co-doped aluminosilicate glasses and transparent glass-ceramics containing gahnite nanocrystals1. Introduction2. Experimental3. Results and discussion4. ConclusionsAcknowledgmentsReferences