Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The...

14
Optical Properties of Transparent Glass–Ceramics Containing Er 3+ -Doped Sodium Lutetium Fluoride Nanocrystals Maria J. Pascual, * Cristian Garrido, and Alicia Dur an Institute of Cer amica y Vidrio (CSIC), C/Kelsen 5, Campus de Cantoblanco, Madrid 28049, Spain Adri an Miguel Department of F ısica Aplicada I, Escuela Superior de Ingenieros, Alda. Urquijo, Bilbao 48013, Spain Laura Pascual Institute of Cat alisis y Petroleoqu ımica (CSIC), C/Marie Curie 2, Campus de Cantoblanco, Madrid 28049, Spain Araceli de Pablos-Mart ın Fraunhofer Institute for Microstructure of Materials and Systems IMWS, Walter-Hulse-Strabe 1, 06120 Halle, Germany Joaqu ın Fern andez and Rolindes Balda Department of F ısica Aplicada I, Escuela Superior de Ingenieros, Alda. Urquijo, Bilbao, 48013, Spain Materials Physics Center CSIC-UPV/EHU, Paseo Manuel Lardizabal 4, Donostia-San Sebastian 20018, Spain Transparent glassceramics containing Er 3+ -doped sodium lutetium fluoride nanocrystals for photonic applications have been synthesized. Glass transition temperature, softening temperature, and crystallization temperature were estimated by dilatometry and differential thermal analysis. Proper heat treatments were selected to crystallize lutetium fluoride nanocrystals. X-ray diffraction analysis was carried out to identify the crystalline phase and the crystal size. HRTEM indicates that the base glass is phase separated in droplets enriched in Lu, Na, F, and also Er ions. The thermal treatment induces the crystallization *[email protected] © 2015 The American Ceramic Society and Wiley Periodicals, Inc International Journal of Applied Glass Science, 7 [1] 27–40 (2016) DOI:10.1111/ijag.12177

Transcript of Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The...

Page 1: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

Optical Properties of Transparent Glass–CeramicsContaining Er3+-Doped Sodium Lutetium FluorideNanocrystals

Maria J. Pascual,* Cristian Garrido, and Alicia Dur�an

Institute of Cer�amica y Vidrio (CSIC), C/Kelsen 5, Campus de Cantoblanco, Madrid 28049, Spain

Adri�an Miguel

Department of F�ısica Aplicada I, Escuela Superior de Ingenieros, Alda. Urquijo, Bilbao 48013, Spain

Laura Pascual

Institute of Cat�alisis y Petroleoqu�ımica (CSIC), C/Marie Curie 2, Campus de Cantoblanco, Madrid28049, Spain

Araceli de Pablos-Mart�ın

Fraunhofer Institute for Microstructure of Materials and Systems IMWS, Walter-H€ulse-Strabe 1,06120 Halle, Germany

Joaqu�ın Fern�andez and Rolindes Balda

Department of F�ısica Aplicada I, Escuela Superior de Ingenieros, Alda. Urquijo, Bilbao, 48013, Spain

Materials Physics Center CSIC-UPV/EHU, Paseo Manuel Lardizabal 4, Donostia-San Sebastian20018, Spain

Transparent glass–ceramics containing Er3+-doped sodium lutetium fluoride nanocrystals for photonic applications havebeen synthesized. Glass transition temperature, softening temperature, and crystallization temperature were estimated bydilatometry and differential thermal analysis. Proper heat treatments were selected to crystallize lutetium fluoride nanocrystals.

X-ray diffraction analysis was carried out to identify the crystalline phase and the crystal size. HRTEM indicates that the baseglass is phase separated in droplets enriched in Lu, Na, F, and also Er ions. The thermal treatment induces the crystallization

*[email protected]

© 2015 The American Ceramic Society and Wiley Periodicals, Inc

International Journal of Applied Glass Science, 7 [1] 27–40 (2016)DOI:10.1111/ijag.12177

Page 2: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

inside the droplets. The optical characterization, which includes absorption and steady-state and time-resolved emission spec-

troscopy under one- and two-photon excitation, shows the differences between the phase-separated base glass and its corre-sponding glass–ceramic. The reduction of the Judd–Ofelt parameter Ω2 together with the increase of the fluorescencelifetimes as compared to the glass sample confirms the presence of Er3+ ions in a crystalline environment in the glass–ceramicsamples. Moreover, an enhancement of the green and red up-converted emissions (as well as the weak blue emission) is

observed in the glass–ceramic, indicating the Er3+ incorporation into the nanocrystals. The possible excitation mechanismsresponsible for this up-conversion luminescence are discussed on the basis of lifetime measurement results.

Introduction

Transparent glass–ceramics are interesting materialsfor various photonic applications. They have a goodforming ability and isotropy, as well as high chemical,photochemical, and mechanical stability, and they pro-vide advantageous fluorescence properties, such as highfluorescence quantum efficiency, long fluorescence life-time, and good up-conversion (UC) properties.1,2

Transparency is achieved when crystal size is in thenanometric scale, usually below 40 nm, avoiding lightscattering. A strict control of the nucleation and crystalgrowth processes is therefore necessary, which requiresa deep knowledge of the crystallization mechanisms.

In particular, transparent oxyfluoride glass–ceramicscontaining rare earth (RE) ions combine the advantagesof superior chemical stability of oxide glasses and low-phonon-energy environment of fluoride nanocrystals.These materials have attracted increasing attention asthey have potential applications in many fields such assolid state lasers, optoelectronic communication devices,color displays, down- and up-conversion phosphors, andW-LED phosphors.3

A. de Pablos et al. have been working on this type ofmaterials, specifically within the SiO2-Al2O3-Na2O-K2O-LaF3/YF3 system, in which heat treatments at tempera-tures above the glass transition temperature induced thecrystallization of different crystalline phases, depending onthe composition: LaF3

4, NaLaF45, KLaF4

6, and NaYF4.7

The crystallization mechanism is shown to occur viaregions of La- and Si-phase separation in the glass, fromwhich the fluoride crystals develop during heat treatment.The interface between the glass matrix and the crystals inthe de-mixed areas is enriched in network formers, mainlySiO2, creating a viscous barrier, which inhibits furthercrystal growth and limits the crystal size to the nanometricrange. Some recent papers further analyze the influence ofphase separation on the nanocrystallization process.8

Among all of the investigated UC host materials,hexagonal-phase sodium yttrium fluoride (b-NaYF4),

known as one of the most efficient host materials ever,has been widely studied recently. As another importantfluoride, NaLuF4 also has excellent UC behavior due tothe similar structure to NaYF4. Recently, it has beenreported that lanthanide-doped hexagonal b-NaLuF4nanocrystals are even more efficient for UC than NaYF4nanocrystals as demonstrated by sensitive in vivobioimaging.9,10 On the other hand, S. K€uck et al.11

reported the phase purity of Pr3+:Na7Lu13F46 by struc-ture analysis which is an ideal cascade emitting phosphorfor Xe discharge lamps. However, there are limitedreports on the study of RE-doped NaLuF4 or LiLuF4nanocrystals obtained by the glass–ceramics route. Onlya few number of papers report about the preparation oftransparent glass–ceramics with lutetium nanocrystals forphotonic applications based on the glass quenchingmethod and always a cubic phase has been obtained.12–14

The present work is focused on the synthesis andcharacterization of sodium lutetium fluoride nano glass–ceramics, in which Lu3+ can be substituted by Er3+

through doping. The developed glass composition is basedon the work of Herrmann et al. in which the precipitationof cubic or hexagonal NaGdF4 nanocrystals was studied.

15

A similar composition was also employed in.14 Amongrare earth ions, Er3+ ions have played an important role inthe development of broadband erbium-doped fiber ampli-fiers (EDFA) due to the 4I13/2?

4I15/2 transition around1.5 lm where standard optical communications fiberspresent low losses. Moreover, erbium ion has an energylevel structure with two transitions around 980 nm and800 nm which can be efficiently excited with commerciallaser diodes, yielding to blue, green, and red emissionsdepending on the host matrix. In the present work, theinfluence of the nanocrystals on the absorption, near-infrared emission, as well as visible emission obtainedunder visible and near-infrared excitation, of Er3+ ions insodium lutetium fluoride nano glass–ceramics is analyzed.The up-converted green and red emissions obtained undernear-infrared excitation in the 4I9/2 level have been investi-gated in the base glass and glass–ceramic samples using

28 International Journal of Applied Glass Science—Pascual, et al. Vol. 7, No. 1, 2016

Page 3: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

steady-state and time-resolved laser spectroscopy and com-pared with those obtained under one-photon excitation.The possible excitation mechanisms responsible for thisup-conversion luminescence are discussed on the basis oflifetime measurement results.

Experimental Procedure

The base glass composition is 70SiO2-5Al2O3-2AlF3-2Na2O-18NaF-3Lu2O3 (mol%) doped with1 mol% of ErF3. This composition is based on Her-rmann et al. composition.15 Raw materials used wereSiO2 sand (Saint-Gobain, 99.6%), Al2O3 (Panreac),Na2CO3 (Panreac, 99.5%), Li2CO3 (Panreac, 99.5%),Lu2O3 (Alfa Aesar, 99.9%), ErF3 (Aldrich, 99.99%),and AlF3 (Alfa Aesar, 99.9%).

The generated batch was stirred in a Turbula mixerfor an hour to homogenize the content. Then, it wasmoved into a covered platinum crucible and calcined inan electric furnace at 1250°C. The batch was thenmelted at 1650°C for one hour. The glass was meltedtwice to increase the homogenization. The melt was caston a brass mold and then cooled down to room tempera-ture in air in order to increase the thermal shock andavoid the glass de-vitrification during cooling.

The coefficient of thermal expansion and the glasstransition and softening temperatures were measuredusing a Dilatometer 402 EP Netzsch G€eratebau at10°C/min under air atmosphere. The used sample wasannealed at 580°C for 30 min.

The glass–ceramics were obtained by different ther-mal treatments of the base glass; temperatures from600 to 750°C for 20 h were used to achieve the crys-tallization of nanometric size crystals.

The samples were properly powdered for X-raydiffraction (XRD) analysis. The crystalline phases wereidentified using a Bruker TT with a lynx-eye detector X-ray diffractometer. Scanning was performed from 10 to70° (2h) at 0.02° and one second per step. This techniquewas used to achieve two objectives: (i) determination ofthe crystalline phase and (ii) estimation of the crystal size.Crystal sizes were resolved from Scherrer equation withthe corresponding errors. A slight modification was madein the equation denominator, including the width of thediffraction peak of NaF, employed as standard.

Differential thermal analysis (DTA) was carriedout on samples with particle sizes between 1.5 and1 mm in order to reproduce the crystallization behavior

of bulk samples. DTA curves were implemented at dif-ferent heating rates, from 30 to 70°C/min, from roomtemperature to 1000°C. From this technique, the fol-lowing information can be extracted: glass transitiontemperature (Tg) and start and maximum crystallizationtemperatures (Tx and Tp, respectively).

Samples of the glasses and glass–ceramics for high-resolution transmission electron microscopy (HRTEM)were prepared by cutting slices, plane parallel grinding,dimpling to a residual thickness of 10–15 lm, andion-beam thinning using Ar+ ions. HRTEM data,including scanning transmission microscopy high-angleannular dark field (STEM-HAADF), and X-ray energy-dispersive spectroscopy (XEDS), were recorded on aJEOL 2100 field emission gun transmission electronmicroscope operating at 200 kV and providing a pointresolution of 0.19 nm. The TEM is equipped with anEDXS energy-dispersive X-ray spectrometer (INCA x-sight, Oxford Instruments). XED analysis was per-formed in STEM mode, with a probe size of ca. 1 nm.

Samples were prepared for optical quality, with asize of 2 9 2 cm and ~3 mm of thickness. Conven-tional absorption spectra were performed with a Cary 5spectrophotometer. The steady-state emission and exci-tation measurements were made with a Ti:sapphire ringlaser (0.4 per cm linewidth) in the 760–940 nm spec-tral range. The fluorescence was analyzed with a 0.25monochromator, and the signal was detected by anextended IR Hamamatsu H10330A-75 photomultiplierand finally amplified by a standard lock-in technique.Visible emission under one-photon excitation was per-formed by exciting with an argon laser at 488 nm anddetected by a Hamamatsu R636 photomultiplier.

Lifetime measurements were obtained by excitingthe samples with a dye laser pumped by a pulsed nitro-gen laser and a Ti:sapphire laser pumped by a pulsedfrequency-doubled Nd:YAG laser (9 ns pulse width),and detecting the emission with Hamamatsu R636 andH10330A-75 photomultipliers. Data were processed bya Tektronix oscilloscope. All measurements were takenat room temperature.

Results and Discussion

The dilatometric curve of the glass sample (notshown) revealed the following dilatometric properties:glass transition temperature (Tg) 589 � 2°C, dilato-metric softening temperature (Td) 667 � 5°C, and

www.ceramics.org/IJAGS Optical Properties of Transparent Glass–Ceramics 29

Page 4: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

coefficient of thermal expansion (a) 8.5 9 10�6 � 0.5per K.

DTA

Figure 1 shows the differential thermal analysis(DTA) curves collected at temperatures from 30°C/minto 70°C/min. Table I shows the corresponding glasstransition temperature (Tg), temperature of beginningof crystallization (Tx), maximum temperature of crystal-lization (Tp), and melting temperature (Tm). Tg valuesare similar to those obtained by dilatometry. Neverthe-less, the differences are reasonable given the differenttechniques and the heating rates in each measurement.

As it can be observed, the crystallization peak isvery small, which can be related to the phase separationand incipient crystallization already present in the baseglass. The general trend observed in the DTA is theincreasing value of Tg and Tx with the heating rate.The Tp variations with the heating rate are insignificant

which is related to low activation energy of the crystal-lization process.

XRD Analysis

Figure 2 shows the XRD patterns of the glassesafter thermal treatment for 20 h at temperatures in therange 600–750°C. The crystal size was calculated fromthe diffraction peak at 2h = 28˚, corresponding to theplane (111).

Samples annealed at different temperatures showalmost identical patterns, and all observed peaks areattributable to a cubic phase. Na5Lu9F32 nanocrystals16

exhibit the same diffractogram (JPCDS No. 27-0725)which can be indexed on a cubic unit cell with parame-ter a = 5.463 A (space group Fm-3 m). A cubica-NaLuF4 powder pattern has not been reported, butthe a-NaYF4 pattern (JPCDS No. 77-2042) is also coin-cident with the XRD patterns obtained in this work. Anorthorhombic phase with stoichiometry Na7Lu13F46(JPCDS No. 48-0839) has also been reported in the lit-erature.17,18 We propose that the obtained nanocrystalsadopt a composition within a cubic solid solution of thetype NaxLu2x-1F7x-3. The peak positions are not changedby applying different annealing temperatures; however,with increasing annealing temperature, the XRD peaksbecome narrower and more intense as a result of greatercrystallization.

As it can be observed in Table II when tempera-ture is increased, crystal size also increases. This is due

Fig. 1. DTA curves for the studied glass at different heatingrates.

Table I. DTA Characteristic Temperatures

Heating rate (�C/min) 30 40 50 60 70

Tg � 7 (˚C) 535 548 554 550 559Tx � 9 (˚C) 635 637 641 651 650Tp � 9 (˚C) 718 718 717 720 724Tm � 9 (˚C) 831 832 830 831 831

Fig. 2. XRD patterns of the studied glass annealed at tempera-tures of 600, 650, 700, and 750°C for 20 h.

30 International Journal of Applied Glass Science—Pascual, et al. Vol. 7, No. 1, 2016

Page 5: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

to the lower viscosity, which favors the fluoridediffusion, and the formation of bigger size crystals.There are no appreciable differences until 750°C. Thistranslates into a fast and effective diffusion of the fluo-rides, so the crystals acquire their final size quick. Onthe other hand, it is expected to observe the formationof agglomerates at high temperatures as thermal treat-ment duration is increased, which is translated into aloss of the transparency in the glass–ceramics.

HRTEM Characterization

Figure 3a shows the TEM micrograph of the baseglass. The glass is clearly phase separated in Lu-enriched regions of two different diameters around 50and 5 nm (Fig. 3b). The HRTEM image of a dropletand its corresponding SAED (selected area electrondiffraction) pattern are shown in Fig. 3c. SAED patternshows very weak diffraction spots indicating an incipi-ent crystallization of the droplets. The compositionalprofile obtained from an EDXS line scan across severaldroplets (indicated with a yellow line in Fig. 3d,e)shows that the droplets are enriched in Lu, F, Na, Al,and also Er. Moreover, it is observed that the Si signalintensity decreases toward the center of the droplet.

Figure 4a shows a TEM micrograph of the glass-ceramic sample treated at 600°C for 20 h. The size ofthe droplets keeps quite the same in comparison withthe base glass. In contrast, in the case of the glass–cera-mic, a higher ordered electron diffraction pattern isobtained, pointing out a higher degree of crystallinity,as it is also directly observed in its correspondingHRTEM image (Fig. 4b). STEM/EDXS measurementsevidence that Al is forming a thin shell surrounding thecrystals (Fig. 4c,d). The same behavior was alsoreported in.19 The Er is incorporated only inside thebigger crystals. The incorporation of Er inside thenanocrystals is clearly observed in the EDXS mappingsin Fig. 5. The size of the crystals determined by XRD

(44 nm) is in good agreement with the size of thebigger droplets around 50 nm which could indicate theformation of a unique crystal inside each droplet. Theremaining volume inside the droplet is filled with Al.In fact, the Al line scan exhibits two well-defined peaksat the edges of the crystal in the line scan, constitutingan Al-enriched shell of 10–20 nm thickness in theperiphery of the crystal.

Optical Characterization

Figure 6 shows the base glass and glass–ceramictreated during 600°C for 20 h used for the opticalcharacterization.

Absorption Spectra: Absorption spectra for the Er3+-doped glass and its corresponding glass–ceramic are shownin Fig. 7. Nine absorption bands corresponding to thetransitions from the 4I15/2 ground state of Er3+ to the 4I13/2,

4I11/2,4I9/2,

4F9/2,4S3/2,

2H11/2,4F7/2,

4F5/2,4F3/2,

2H9/

2,2G11/2, and

2G7/2 excited levels are observed in bothsamples. As evidenced in the HRTEM analysis, the baseglass is already phase separated and there is Er enrichmentalready in the phase-separated droplets. As a result, bothspectra are quite similar. However, the analysis of theabsorption bands shows that the intensity of the hypersen-sitive transitions 4I15/2?

2H11/2 and4I15/2?

4G11/2 slightlydecreases in the heat-treated sample compared with thebase glass, which indicates the variation in the local struc-ture around Er3+ ions.

The absorption spectra of both samples have beenanalyzed in the framework of the Judd–Ofelt theory.Using this theory, the Judd–Ofelt (JO) parameters forthe base glass and glass–ceramic samples, Ωt, have beenobtained by fitting the electric dipole contributions ofthe experimental oscillator strengths to the calculatedones by a least-squares method.20,21 The experimentaloscillator strengths have been calculated from the bandsobserved in the absorption spectra using the expression:

fexp ¼ mec2

pe2N

Zað�mÞd�m ð1Þ

where N is the ion concentration, me and e are theelectron mass and the electron charge, respectively, c isthe velocity of light, and að�mÞ is the absorption coeffi-cient. All the transitions observed in the absorptionspectra have purely electric dipolar character, with theexception of the 4I15/2?

4I13/2 which contains electric

Table II. Crystal Size for the Different ThermalTreatments

Thermal treatment (T,t) Crystal size (nm)

600°C, 20 h 44 � 1650°C, 20 h 38 � 1700°C, 20 h 59 � 2750°C, 20 h 146 � 24

www.ceramics.org/IJAGS Optical Properties of Transparent Glass–Ceramics 31

Page 6: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

dipole and magnetic dipole contributions. The mag-netic dipole contribution to the experimental oscillatorstrength in the 4I15/2?

4I13/2 transition has been esti-mated from the expression fmd = nf

0, where n is the

refractive index of the studied samples and f0 is a physi-cal quantity reported by Carnall22, and then subtractedfrom the experimental one. The obtained Judd–Ofeltparameters of Er3+ for the base glass and the heat-trea-ted samples are shown in Table III.

The obtained parameters are similar to thosefound in other oxyfluoride silicate-based glasses.23,24

According to the Judd–Ofelt theory, the Ω2 parameteris very sensitive to the local symmetry around the rareearth ion sites as well as to the covalency of the chemi-cal bonding between the rare earth ions and theligands.25,26 Because the value of the Ω2 parameter isindicative of the amount of covalent bonding betweenRE ions and ligand anions, its value decreases with the

(a) (c)

(d)(b)

(e)

Fig. 3. (a) TEM micrograph of the base glass. (b) HRTEM micrograph and SAED of a droplet. (c) Size distributions of the droplets.(d and e) STEM image and EDX of a line scan crossing several droplets.

32 International Journal of Applied Glass Science—Pascual, et al. Vol. 7, No. 1, 2016

Page 7: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

composition of the host from oxides to fluoride, indi-cating a more ionic bonding. In this case, a value ofΩ2 = 3.96 9 10�20 is obtained for the glass sample,which is intermediate between the typical values forfluoride and for oxide matrices27,28, which indicatesthat the coordination anions of Er3+ ions are both F-

and O2- ions in this glass. Moreover, the Ω2 parameterdecreases for the heat-treated sample, which suggeststhat the heat treatment produces a decrease in thecovalency as well as in the asymmetry around the Er3+

ions which confirms the incorporation of Er3+ ionsinto the nanocrystals. In addition, the Ω2 parameter isclosely related to the hypersensitive transitions.29,30

The more intense the hypersensitive transition is, thelarger the value of Ω2 is. The lower value of thisparameter in the heat-treated sample is in agreementwith the decrease in the strength of the hypersensitivetransitions 4I15/2?

4G11/2 and 4I15/2?2H11/2 which

indicates the variation in the local structure around the

Er3+ ions and further confirms their incorporation intothe nanocrystals.

The probability of spontaneous emission, A [(S, L)J; S0, L0) J0], between the excited states, (S, L) J, andlower manifolds, (S0, L0) J0, is given in the Judd–Ofelttheory by:

A S ; Lð ÞJ ; S 0; L0ð ÞJ 0½ � ¼ Aed þ Amd

¼ 64p4

3hk3ð2J þ 1Þ ½ved Sed þ vmd Smd �

ð2Þwhere ved ¼ nðn2þ2Þ2

9 is the local field correction forelectric dipole transitions and vmd = n3 for magneticdipole transitions and Sed and Smd are the electric andmagnetic dipole line strengths, respectively.

The radiative lifetime of an excited state in termsof the total radiative transition probability can be calcu-lated as:

(a)

(b)

(c)

(d)

Fig. 4. (a) TEM micrograph of the glass–ceramic treated at 600°C for 20 h. (b) HRTEM micrograph and SAED of a nanocrystal. cand d) STEM image and EDX of a line scan crossing a nanocrystal.

www.ceramics.org/IJAGS Optical Properties of Transparent Glass–Ceramics 33

Page 8: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

sR ¼ 1PS 0;L0J 0 A½ S ; Lð ÞJ ; S 0; L0ð ÞJ 0� ð3Þ

The calculated radiative lifetimes of level 4I13/2 forthe glass and the heat-treated samples are 10.8 and11.5 ms, respectively.

NIR Emission Spectra: The near-infrared emissionspectra have been obtained for both samples by excit-ing at 800 nm in resonance with the 4I9/2 level. Afterexcitation of this level, the next lower levels are popu-

lated by multiphonon relaxation. As can be seen inFig. 8, the spectra show two bands corresponding tothe 4I11/2?

4I15/2 and 4I13/2?4I15/2 transitions. The

(a)

(b)

(c)

(d)

Fig. 5. (a) STEM-HAADF micrograph of the glass–ceramic and its corresponding EDXS mappings of (b) Lu (c) Er, (d) F.

Fig. 6. Transparent glass obtained (left) and it correspondingglass–ceramic treated at 600°C for 20 h (right).

Fig. 7. Absorption spectra at room temperature for the baseglass and corresponding glass–ceramic.

34 International Journal of Applied Glass Science—Pascual, et al. Vol. 7, No. 1, 2016

Page 9: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

spectral features are similar for the glass and glass–cera-mic samples; however, the intensity of the 4I11/2?

4I15/2 emission is higher in the glass–ceramic sample againstthe base glass. Moreover, the lifetime of this level islonger in the glass–ceramic sample (Fig. 9a). The life-time of this level has a strong dependence with thephonon energy of the matrix, as it is at ~3600 per cmfrom the first excited level 4I13/2. Multiphonon de-exci-tation is more likely to occur if the phonon energy ishigh which reduces the lifetime. Thus, the increase inthe lifetime from 1 ms in the base glass to 2.3 ms inthe glass–ceramic sample suggests a fluoride environ-ment for Er3+ ions according to their incorporation inthe nanocrystalline phase. A longer lifetime (6.8 ms) isalso observed for the 4I13/2 level in the case of theglass–ceramic sample (Fig. 9b).

The stimulated emission cross section of the4I13/2?

4I15/2 transition has been determined from theemission spectra and the radiative transition probabil-ity calculated through the Judd–Ofelt parametersusing the expression:31

reff ¼ k4pArad

8pn2cDkeffð4Þ

where kp is the peak fluorescence wavelength, n is therefractive index, Dkeff is the effective linewidth of the4I13/2?

4I15/2 transition defined by Dkeff ¼ R I kð ÞdkImax

,

Table III. Er3+ Ions Concentration, Judd–Ofelt Intensity Parameters, and r.m.s. Deviation for the Base Glassand the Heat-Treated Sample

Sample N (ions/cm3) Ω2 3 10�20 Ω4 3 10�20 Ω6 3 10�20 r.m.s.

Glass 2.22 9 1020 3.96 1.19 0.88 1.448 9 10�7

600°C - 20 h 2.24 9 1020 3.50 1.07 0.81 1.243 9 10�7

Fig. 8. Room temperature near-infrared emission spectraobtained by exciting level 4I9/2 for the base glass and correspond-ing glass–ceramic.

(a)

(b)

Fig. 9. Semilogarithmic plot of the fluorescence decays of the(a) 4I11/2 and (b) 4I13/2 levels for the glass and glass–ceramicsamples obtained after excitation at 800 nm.

www.ceramics.org/IJAGS Optical Properties of Transparent Glass–Ceramics 35

Page 10: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

and Arad is the radiative transition probability for thistransition. The spectroscopic parameters and the stimu-lated emission cross section of the 4I13/2?

4I15/2 transi-tion as well as the refractive index for the glass and theheat-treated samples are shown in Table IV. The valuesof the emission cross section obtained in the studiedsamples are close to those found in other Er3+-dopedsilica-based glasses.32 The effective linewidth in bothsamples is relatively large compared to silica-basedglasses (�37 nm)33 which indicates the influence of flu-orine bonding on the environment inhomogeneity ofEr3+ ions. This broadband emission is useful to developtunable amplifiers for optical communications.

Direct and Up-Converted Visible Emission: Theluminescence spectra of the base glass and the heat-trea-ted sample obtained under visible excitation are shownin Fig. 10. Under excitation at 488 nm, the Er3+ ions

are promoted to the 4F7/2 level and then, the 2H11/2,4S3/2, and 4F9/2 levels are populated through multi-phonon relaxation. In both cases, the emission spectrashow green and red emissions which can be assigned tothe (2H11/2,

4S3/2)?4I15/2 and 4F9/2?

4I15/2 transitions,respectively. The green emission consists of two over-lapped bands because the population of the 2H11/2 and4S3/2 levels are in thermal equilibrium at room temper-ature. The intensity of the green emission remainsalmost unchanged after the heat treatment; however,the intensity of the red emission becomes slightlyhigher in the heat-treated sample. The rise in the inten-sity of the red emission may be due to the fact thatpart of Er3+ ions are forming clusters, which leads to areduction in the interionic Er3+-Er3+ distances. Thisfavors the energy transfer mechanism between Er3+ ions(4F7/2?

4I11/2); (4I11/2?4F9/2) populating the 4F9/2

level.Visible emission has also been obtained for both

samples under near-infrared (NIR) excitation at800 nm. The spectra consist of two main bands, corre-sponding to the emissions from the (2H11/2,

4S3/2) and4F9/2 levels to the ground state (4I15/2) and a weakband due to the 2H9/2?

4I15/2 transition. As it may beseen in Fig. 11, the red emission is strongly enhancedunder NIR excitation, which means that the 4F9/2 levelis not only populated through multiphonon relaxationfrom the immediately upper level 4S3/2, but other pro-cesses such as energy transfer are also present. Theincrease in the red up-conversion emission has beenpreviously observed in other glass–ceramics and attribu-

Table IV. Room Temperature Emission Propertiesof the 4I13/2?

4I15/2 Transition for the Base Glassand the Heat-Treated Sample

Sample n kp (nm)

Arad

(4I13/2)(per s)

Dkeff(nm)

reff(10�21

cm2)

Glass 1.458 1533.3 92.5 59.8 5.3600°C -20 h

1.453 1533.3 86.6 60.2 5.0

Fig. 10. Room temperature visible emission spectra of the baseglass and glass–ceramic samples obtained under excitation at488 nm.

Fig. 11. Room temperature up-conversion emission spectra ofthe base glass and glass–ceramic samples obtained under excita-tion at 800 nm.

36 International Journal of Applied Glass Science—Pascual, et al. Vol. 7, No. 1, 2016

Page 11: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

ted to the formation of “clusters” where the RE ionsare mainly incorporated surrounded by a fluorine envi-ronment.23,34 The short distances among the Er3+ ionsinto the formed clusters favor the interionic interactionsimproving the efficiency of the energy transfer up-con-version (ETU) processes. Furthermore, the up-con-verted luminescence of the heat-treated sample is threetimes more intense than in the base glass. The forma-tion of a fluorine-rich environment around the Er3+

ions in the heat-treated sample, with lower maximumphonon energy (~500 per cm) than in silicate glasses,leads to a reduction in the vibration energy of the pho-nons coupled to the Er3+ ions which benefits the up-converted luminescence. It is also remarkable the pres-ence of the weak up-converted blue emission due tothe 2H9/2?

4I15/2 transition in the present sampleswhich indicates that Er3+ ions are in a local environ-ment of low phonon energy.

The up-conversion emission can be due to excitedstate absorption (ESA) and/or energy transfer up-con-version (ETU) processes. The analysis of the temporaldecay of the up-converted luminescence under pulsedexcitation provides a way to distinguish between bothmechanisms. Thus, the luminescence decays from the4S3/2 and 4F9/2 levels have been recorded under 800-nm excitation and then compared with those obtainedunder direct excitation at 488 nm (Table V). In thecase of the green emission, whereas the luminescencedecays obtained under excitation at 488 nm show anearly single-exponential behavior, two different com-ponents can be distinguished under 800-nm excitationwith lifetimes much longer than under direct excitation(see Table V). This behavior suggests the presence ofETU processes to populate this level. On the otherhand, in the case of the red emission, the decays arewell described by single-exponential functions under488-nm and 800-nm excitation. As it may be seen in

Table V, the lifetimes obtained under infrared excita-tion, in both samples, are longer than under visibleexcitation. This lengthening of the lifetimes togetherwith the initial rise time observed in the temporaldecay of the red luminescence indicates the presence ofETU processes in order to populate the 4F9/2 level.Besides, the lifetimes of the 4S3/2 and 4F9/2 levels underone-photon excitation (488 nm) are slightly longer inthe heat-treated sample which is attributed to the exis-tence of more symmetric and lower phonon environ-ment around the rare earth ions. Figure 12 shows asemilogarithmic plot of the experimental decays of thegreen and red emissions for both samples under excita-tion at 800 nm.

An analysis of the lifetime values obtained under800-nm excitation shows that in the case of the greenemission from 4S3/2 level, the lifetime of the long com-ponent is roughly half the 4I11/2 lifetime (see Table V).This suggests that under excitation at 800 nm, thegreen emission is mainly due to an ETU mechanism inwhich two neighboring Er3+ ions in the 4I11/2 levelinteract, promoting one of them to the 4F7/2 levelwhereas the other one loses its energy and goes to theground state35. Then, the 4F7/2 relaxes nonradiativelyto the (2H11/2,

4S3/2) coupled levels which emit to theground state. Nevertheless, the absence of rise time sug-gests that ESA processes from the lower levels 4I11/2and 4I13/2 can also contribute to the green up-convertedluminescence. By pumping at 800 nm, the Er3+ ionsare excited to the 4I9/2 level, then multiphonon relax-ation occurs to level 4I11/2, and then the sequentialabsorption of a second NIR pump photon promotesthe electrons to the 4F3/2,5/2 levels, and finally, by non-radiative relaxation, 2H11/2 and 4S3/2 levels are reached.Another possibility is an ESA from level 4I13/2 popu-lated radiatively and nonradiatively from level 4I11/2.These mechanisms are displayed in Fig. 13.

Table V. Lifetimes of the 4S3/2 and 4F9/2 Levels Obtained under Excitation at 488 nm and 800 nm Togetherwith the Lifetime of the 4I11/2 Level

Samplekexc = 488 nm(4S3/2) (ls)

kexc = 800 nm(4S3/2) (ms)

kexc = 488 nm(4F9/2) (ls)

kexc = 800 nm(4F9/2) (ms)

kexc = 800 nm(4I11/2) (ms)

Glass 9 s1 = 36s2 = 431

160 0.97 1.0

600°C - 20 h 10 s1 = 128s2 = 1300

188 2.2 2.3

www.ceramics.org/IJAGS Optical Properties of Transparent Glass–Ceramics 37

Page 12: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

In the case of the 4F9/2 level, this level can be popu-lated by several ETU processes in which the 4I11/2 levelacts an intermediate state (Fig. 13).35 Firstly, one Er3+

ion in the 4I11/2 level can transfer its energy to anotherion in the 4I13/2, which is promoted to the 4F9/2 level,according to the nonresonant process described by thepair of transitions (4I11/2?

4I15/2);(4I13/2?

4F9/2)(DE�+1539 per cm). Besides, one Er3+ ion in the 4I9/2level can transfer part of its energy, going down to the4I13/2 level, to another ion in the 4I11/2 level, whichreaches the 4F9/2 level by the nonresonant mechanism(4I9/2?

4I13/2)?(4I11/2?4F9/2) (DE�+777 per cm).

Finally, an Er3+ ion in the 4F7/2 level can transfer part ofits energy to another ion in the 4I11/2 level, going both tothe 4F9/2 level by the nearly resonant ETU process (4F7/

2?4F9/2);(

4I11/2?4F9/2). As it can be seen in Table V,

the lifetime of the 4F9/2 level under 800-nm excitation ispractically the same as 4I11/2 lifetime. This together withthe values of the energy mismatch of the different pro-cesses suggests that (4I9/2?

4I13/2);(4I11/2?

4F9/2) and(4F7/2?

4F9/2);(4I11/2?

4F9/2) are the likeliest ones toexplain the population of the 4F9/2 level by ETU.However, we cannot disregard the contribution of the(4I11/2?

4I15/2);(4I13/2?

4F9/2) process.35

Finally, the weak blue emission from 2H9/2 levelshould be due to a three-photon up-conversion processsince although this level lies at twice the energy of the4I9/2, an ESA process from the 4I9/2 to the 2H9/2 levelseems to be unlikely because of the maximum phononenergy in these hosts. There are different up-conversionprocesses to explain the population of this level(Fig. 13). Firstly, a nonresonant ESA process from the4F9/2 level, previously populated through the ETUmechanisms described above, can occur promoting theEr3+ ions to the 4G9/2 (DE�+315 per cm) whichrelaxes nonradiatively to the 2H9/2 level. In addition,two Er3+ ions in the 4S3/2 level can interact through anearly resonant ETU process determined by the pair oftransitions (4S3/2?

4I9/2);(4S3/2?

2H9/2) (DE��219per cm).36 Another possibility is the quasi-resonantETU mechanism (4S3/2?

4I11/2);(4S3/2?

2G11/2)(DE�+158 per cm).37 Besides, the 2H9/2 can be popu-

Fig. 13. Energy level diagram with the observed luminescencetransitions and possible up-conversion mechanisms under excita-tion at 800 nm. Black lines indicate absorption of 800 nm IRphotons. Dashed lines correspond to ETU processes to populatelevels 4S3/2 (green lines), 4F9/2 (red lines), and 2H9/2 (blue lines).

(a)

(b)

Fig. 12. Experimental emission decay curves of the (a) greenand (b) red emission under excitation at 800 nm for the glassand glass–ceramic sample.

38 International Journal of Applied Glass Science—Pascual, et al. Vol. 7, No. 1, 2016

Page 13: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

lated through an ETU process in which the Er3+ ionsin the 4F9/2 level relax to the 4I13/2 level and can trans-fer part of their energy to those in the 4S3/2 level whichare promoted to the 4G9/2 according to the pair oftransitions (4F9/2?

4I13/2);(4S3/2?

4G9/2) (DE��345per cm).38 Then, the 2H9/2 level is reached by multi-phonon relaxation from the upper levels.

Conclusions

Transparent glass–ceramics containing Er3+-dopednanocrystals have been prepared from adequate thermaltreatment of an Er3+-doped SiO2-Al2O3-AlF3-Na2O-NaF-Lu2O3 glass. The parent glass is constituted by50- and 5-nm amorphous droplets containing Lu, F,Na, Al, and Er. The near surrounding of the dropletsis enriched in Al and Si. After the heat treatment ofthe parent glass, the droplets develop nanocrystals,which act as main host for the Er3+ ions. The obtainednanocrystals adopt a composition within a cubic solidsolution of the type NaxLu2x-1F7x-3. The Al-enrichedlayer around the crystals enables the nanocrystal size.

The analysis of the absorption and emission prop-erties suggests that Er3+ ions enter in the nanocrystalsin the heat-treated sample. The reduction in the Ω2 JOparameter and the intensity of the hypersensitive transi-tions in the heat-treated sample indicate that Er3+ ionsare in a crystalline environment. The longer lifetimes ofthe 4I11/2 and 4I13/2 levels together with the enhancedintensity of the 4I11/2?

4I15/2 emission in the glass–ceramic sample compared to the base glass suggest afluoride environment for Er3+ ions according to theirincorporation in the nanocrystalline phase. This is alsoconfirmed by HRTEM elemental analysis.

Finally, intense up-conversion emission due to(2H11/2,

4S3/2)?4I15/2 and 4F9/2?

4I15/2 transitionstogether with a weak blue emission due to 2H9/2?

4I15/2transition has been observed under excitation at 800 nmin the glass and glass–ceramic samples. The up-convertedluminescence of the heat-treated sample is three timesmore intense than in the base glass. In both samples,an enhancement of the red emission is observed under800-nm excitation, which suggests that ETU processespopulate level 4F9/2. The time evolution of the up-converted red emission suggests that ETU processes areresponsible for the increase of this emission in the glassand glass–ceramic samples. The temporal dynamic of thegreen emission from level 4S3/2 indicates that this level is

populated by an ETU process involving the interactionof two Er3+ ions in the 4I11/2 level. Nevertheless, theabsence of rise time suggests that ESA processes from thelower levels 4I11/2 and 4I13/2 can also contribute to thegreen up-converted luminescence.

These results indicate that transparent glass–cera-mics based on cubic sodium lutetium fluoridenanocrystals present advantageous properties such ashigh transparency, high chemical and mechanical stabil-ity, high fluorescence intensity and lifetime, and up-conversion effect. This can be useful for applications inthe field of high-power LED lightning, up-conversionmaterial in solar technology, or fluorescence standard inbiotechnology.

Acknowledgments

This work was supported by the Spanish Govern-ment under projects MAT2013-48246-C2-1-P andMAT2013-48246-C2-2-P, and Basque CountryGovernment IT-659-13. A. Miguel acknowledges finan-cial support from the University of the Basque CountryUPV-EHU. A. de Pablos-Mart�ın thanks FhG InternalPrograms under Grant No.Attract 692 280. Theauthors thank Dr. G.C.Mather for helpful discussion.

References

1. A. de Pablos-Mart�ın, A. Dur�an, and M. J. Pascual, “Nanocrystallisation inOxyfluoride Systems. Mechanisms of Crystallization and Photonic Proper-ties,” Int. Mater. Rev., 57 165–186 (2012).

2. A. De Pablos-Mart�ın, M. Ferrari, M. J. Pascual, and G. C. Righini,“Glass-Ceramics: A Class of Nanostructured Materials for Photonics,” RIV.Nuovo Cimento, 38 311–369 (2015).

3. D. Chen, Y. Yu, P. Huang, and Y. Wang, “Nanocrystallization of Lan-thanide Trifluoride in an Aluminosilicate Glass Matrix: Dimorphism andRare Earth Partition,” Cryst. Eng. Comm., 11 1686–1690 (2009).

4. A. de Pablos-Mart�ın et al. , “Crystallization Kinetics of LaF3 Nanocrystalsin an Oxyfluoride Glass,” J. Am. Ceram. Soc., 8 2420–2428 (2011).

5. A. de Pablos-Mart�ın, M. O. Ram�ırez, A. Dur�an, L. E. Baus�a, and M. J.Pascual, “Tm3+ Doped Oxy-Fluoride Glass-Ceramics Containing Nalaf4Nano-Crystals,” Opt. Mater., 33 180–185 (2010).

6. A. de Pablos-Mart�ın et al. , “KLaF4 Nanocrystallisation in OxyfluorideGlass-Ceramics,” Cryst. Eng. Comm., 15 10323–10332 (2013).

7. de Pablos-Mart�ın A., J. M�endez-Ramos, J. del-Castillo, A. Dur�an, V. D.Rodr�ıguez, and M. J. Pascual, “Crystallization and Up-Conversion Lumi-nescence Properties of Er3+/Yb3+-Doped Nayf4-Based Nano-Glass-Cera-mics,” J. Eur. Ceram. Soc., 35 1831–1840 (2015).

8. C. Lin, C. Bocker, and C. R€ussel, “Nanocrystallization in OxyfluorideGlasses Controlled by Amorphous Phase Separation,” Nano Lett., 156764–6769 (2015).

9. L. Wang et al. , “Enhanced Deep-Ultraviolet Upconversion Emission ofGd3+ Sensitized by Yb3+ and Ho3+ in Β-Naluf4 Microcrystals under980 nm Excitation,” J. Mater. Chem. C, 1 2485 (2013).

www.ceramics.org/IJAGS Optical Properties of Transparent Glass–Ceramics 39

Page 14: Optical Properties of Transparent Glass–Ceramics ...righini/TC20/PASCUAL_ijags_2016.pdf · The possible excitation mechanisms responsible for this up-conversion luminescence are

10. Q. Liu, Y. Sun, T. Yang, W. Feng, C. Li, and F. Li, “Sub-10 nm Hexago-nal Lanthanide-Doped Naluf4 Up-Conversion Nanocrystals for SensitiveBioimaging In Vivo,” J. Am. Chem. Soc., 133 17122–17125 (2011).

11. S. K€uck, I. Sok�olska, M. Henke, T. Scheffler, and E. Osiac, “ThermalCoupling of 1S0 and 4f5d Levels and Second Optical, Centre In Pr3+

Doped Na7Lu13F46,” Phys. Rev., B, 71 165112 (2005).12. Y. Le Wei, X. Liu, X. Chi, R. Wei, and H. Guo, “Intense Up-Conversion

in Novel Transparent Naluf4:Tb3+, Yb3+ Glass–Ceramics,” J. Alloys Comp.,

578 385–388 (2013).13. D. Chen et al. , “Enhanced Up-Conversion Luminescence in Phase-Separa-

tion-Controlled Crystallization Glass-Ceramics Containing Yb/Er(Tm):Naluf4 Nanocrystals,” J. Eur. Ceram. Soc., 35 2129–2137 (2015).

14. Z. Wan et al. , “Eu3+ and Er3+ Doped NaLu1�xYbxF4 (x = 0� 1) Solid-Solution Self-Crystallization Nano-Glass-Ceramics: Microstructure andOptical Spectroscopy,” J. Eur. Ceram. Soc., 35 3673–3679 (2015).

15. A. Herrmann, M. Tylkowski, C. Bocker, and C. R€ussel, “Cubic and Hexag-onal NaGdF4 Crystals Precipitated from an Aluminosilicate Glass: Prepara-tion and Luminescence Properties,” Chem. Mat., 25 2878–2884 (2013).

16. I. M. Shmyt0ko, and G. K. Strukova, “Fine Structure of Na5Lu9F32Nanocrystallites Formed at the Initial Stages of Crystallization,” Phys. SolidState, 51 1796–1800 (2009).

17. B. P. Sobolev, The Rare Earth Trifluorides, Institute of Crystallography,Moscow, Russia, 2000.

18. A. M. Golubev, P. P. Fedorov, O. S. Bondareva, B. P. Sobolev, and V. I.Simonov, “Model of the atomic structure Na7Lu13F46,” Institute of Crys-tallography, Russian Academy of Science (2000).

19. A. de Pablos-Martin, C. Patzig, T. H€oche, A. Dur�an, and M. J. Pascual,“Distribution of Thulium in Tm3+-Doped Oxyfluoride Glasses and Glass-Ceramics,” Cryst. Eng. Comm., 15 6979–6985 (2013).

20. B. R. Judd, “Optical Absorption Intensities of Rare-Earth Ions,” Phys.Rev., 127 750–761 (1962).

21. G. S. Ofelt, “Intensities of Crystal Spectra of Rare-Earth Ions,” J. Chem.Phys., 37 511–520 (1962).

22. W. T. Carnall, P. R. Fields, and B. G. Wybourne, “Spectral Intensities ofthe Trivalent Lanthanides and Actinides in Solution. I. Pr3+, Nd3+, Er3+,Tm3+, and Yb3+,” J. Chem. Phys., 42 3797–3806 (1965).

23. J. Mendez-Ramos et al. “Optical Properties of Er3+ Ions in TransparentGlass Ceramics,” J. All. Comp., 323-324 753–758 (2001).

24. L. Lin, G. Ren, M. Chen, Y. Liu, and Q. Yang, “Study of Fluorine Lossesand Spectroscopic Properties of Er3+ Doped Oxyfluoride Silicate Glassesand Glass Ceramics,” Opt. Mater., 31 1439–1442 (2009).

25. H. Ebendorff-Heidepriem, D. Ehrt, M. Bettinelli, and A. Speghini, “Effectof Glass Composition on Judd–Ofelt Parameters and Radiative DecayRates of Er3+ in Fluoride Phosphate and Phosphate Glasses,” J. Non-Cryst.Solids, 240 66–78 (1998).

26. C. K. Jorgensen, and R. Reisfeld, “Judd-Ofelt Parameters and ChemicalBonding,” J. Less-Common Met., 93 107–112 (1983).

27. X. Zou, and T. Izumitani, “Spectroscopic Properties and Mechanisms ofExcited State Absorption and Energy Transfer Upconversion for Er3+-Doped Glasses,” J. Non-Cryst. Solids, 162 68–80 (1993).

28. S. Tanabe, T. Ohyagi, S. Todoroki, T. Hanada, and N. Soga, “Relationbetween the Ω6 Intensity Parameter of Er3+ Ions and the 151Eu IsomerShift in Oxide Glasses,” J. Appl. Phys., 73 8451–8454 (1993).

29. S. F. Mason, R. D. Peacock, and B. Stewart, “Ligand-Polarization Contri-butions to Intensity of Hypersensitive Trivalent Lanthanide Transitions,”Mol. Phys., 30 1829–1841 (1975).

30. C. K. Jorgensen, and B. R. Judd, “Hypersensitive Pseudoquadrupole Tran-sitions in Lanthanides,” Mol. Phys., 8 281–290 (1964).

31. M. J. Weber, D. C. Ziegler, and C. A. Angell, “Tailoring StimulatedEmission cross Sections of Nd3+ Laser Glass: Observation of Large CrossSections for BiCl3 Glasses,” J. Appl. Phys., 53 4344–4350 (1982).

32. X. Qiao, X. Fan, M. Wang, and X. Zhang, “Up-Conversion Luminescenceand near Infrared Luminescence of Er3+ in Transparent Oxyfluoride Glass-Ceramics,” Opt. Mater., 27 597–603 (2004).

33. Y. Ding et al. , “Spectral Properties Of Erbium-Doped Lead HalotelluriteGlasses for 1.5 lm Broadband Amplification,” Opt. Mater., 15 123–130(2000).

34. A. Miguel et al. , “Structural, Optical, and Spectroscopic Properties ofEr3+-Doped TeO2-ZnO-ZnF2 Glass-Ceramics,” J. Eur. Ceram. Soc., 343959–3968 (2014).

35. R. Balda, A. J. Garc�ıa-Adeva, J. Fern�andez, and J. M. Fdez-Navarro,“Infrared-To-Visible Up-Conversion of Er3+ Ions in GeO2-PbO-Nb2O5

Glasses,” J. Opt. Soc. Am. B, 21 744–752 (2004).36. R. Balda, A. Garcia-Adeva, M. Voda, and J. Fernandez, “Up-Conversion

Processes in Er3+-Doped KPb2Cl5,” Phys. Rev. B, 69 205203 (2004).37. Z. Pan, S. H. Morgan, A. Loper, V. King, B. H. Long, and W. E.

Collins, “Infrared to Visible Upconversion in Er3+-Doped-Lead-Germa-nate Glass: Effects of Er3+ Ion Concentration,” J. Appl. Phys., 77 4688–4692 (1995).

38. T. Riedener, P. Egger, J. Hulliger, and H. U. Gudel, “UpconversionMechanisms in Er3+-Doped Ba2YCl7,” Phys. Rev. B, 56 1800–1808(1997).

40 International Journal of Applied Glass Science—Pascual, et al. Vol. 7, No. 1, 2016