Next Article in Journal
The Effects of Hot-Pack Coating Materials on the Pack Rolling Process and Microstructural Characteristics during Ti-46Al-8Nb Sheet Fabrication
Next Article in Special Issue
Van der Waals Epitaxy of III-Nitrides and Its Applications
Previous Article in Journal
Harnessing Multi-Photon Absorption to Produce Three-Dimensional Magnetic Structures at the Nanoscale
Previous Article in Special Issue
Optical Properties of Red-Emitting Rb2Bi(PO4)(MoO4):Eu3+ Powders and Ceramics with High Quantum Efficiency for White LEDs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Temperature-Dependent Luminescence of Red-Emitting Ba2Y5B5O17: Eu3+ Phosphors with Efficiencies Close to Unity for Near-UV LEDs

by
Egle Ezerskyte
1,
Julija Grigorjevaite
1,
Agne Minderyte
1,
Sebastien Saitzek
2 and
Arturas Katelnikovas
1,*
1
Institute of Chemistry, Faculty of Chemistry and Geoscience, Vilnius University, Naugarduko 24, LT-03225 Vilnius, Lithuania
2
Université d’ARTOIS, CNRS, Centrale Lille, ENSCL, Université de Lille, UMR 8181, Unité de Catalyse et Chimie du Solide (UCCS), F-62300 Lens, France
*
Author to whom correspondence should be addressed.
Materials 2020, 13(3), 763; https://doi.org/10.3390/ma13030763
Submission received: 19 December 2019 / Revised: 30 January 2020 / Accepted: 5 February 2020 / Published: 7 February 2020
(This article belongs to the Special Issue Optical Materials for White Light Emitting Diodes (WLEDs))

Abstract

:
Solid state white light sources based on a near-UV LED chip are gaining more and more attention. This is due to the increasing efficiency of near-UV-emitting LED chips and wider phosphors selection if compared to devices based on blue LED chips. Here, a brief overview is given of the concepts of generating white light employing near-UV LED and some optical properties of the available phosphors are discussed. Finally, the synthesis and optical properties of very efficient red-emitting Ba2Y5B5O17:Eu3+ phosphor powder and ceramics is reported and discussed in terms of possible application as a red component in near-UV LED-based white light sources.

Graphical Abstract

1. Introduction

During the past decade, phosphor-converted white light-emitting diodes (pc-WLEDs) were the research interest for many scientists and engineers because of their outstanding brightness and luminous efficiency, low power consumption, long device lifetime, reliability and eco-friendly characteristics compared to other lighting sources, for instance, incandescent and halogen light bulbs, and fluorescent lamps. To date, white LEDs are the most efficient lighting source [1,2,3]. There are several ways to produce white light by employing a blue or near-UV-emitting LED chip. The first white LEDs were prepared by putting the yellow-emitting phosphor on the blue-emitting LED chip; however, such approach yielded cool white light with low colour rendering index (CRI) thus unsuitable for indoor lighting. In order to solve the deficiency of intensity in the red spectral region, the red-emitting phosphor was added together with yellow-emitting phosphor and the CRI and correlated colour temperatures (CCT) of obtained white light sources were significantly improved. At the very beginning of the solid state lighting (SSL) research the efficiency of near-UV-emitting LED chips was considerably lower if compared to their blue-emitting counterparts; therefore, they were not considered for application in SSL sources. However, with constant development of LED chips, the efficiency of near-UV-emitting LED chips was significantly improved and this opened a new possibility to generate white light. The graphical representation of several ways to produce white light employing the near-UV LED is given in Figure 1. The first option would be putting a blend of blue, green, and red broadband-emitting phosphors on near-UV LED. There are many blue, green, and red broadband-emitting phosphors doped with Eu2+ or Ce3+ ions reported in the literature and probably the most well-known are blue-emitting BaMgAl10O17:Eu2+ (BAM) [4], green-emitting Lu3Al5O12:Ce3+ (LuAG:Ce) [5], and red-emitting CaAlSiN3:Eu2+ [6] and (Ca,Sr,Ba)2Si5N8:Eu2+ [7]. Due to presence of the broad emission band of the red-emitting nitride phosphor, part of it extends to wavelengths longer than 650 nm, i.e., to the spectral region where the human eye sensitivity is very low. This reduces the luminous efficacy of the whole light source, thus this part of the spectrum can be considered as a waste [8].
From the practical point of view, the white light source containing a red line emitting phosphor at ca. 610–615 nm is the best compromise between luminous efficacy and colour rendering [9]. Such concept for generating white light is shown in Figure 1b. For such a concept, Eu3+-doped phosphors seem to be the perfect candidates to replace the broadband red-emitting phosphors, since they exhibit strong luminescence in the range of 600–625 nm originating from intraconfigurational 5D07F2 transitions of Eu3+ ions. Moreover, Eu3+-doped phosphors also possess high photostability, high luminous efficacy and high quantum efficiency [9].
However, the main disadvantage of Eu3+-doped phosphors is weak absorption of Eu3+ ion in the blue or even in the near-UV. This is especially true for the excitation in the blue where only one 7F05D2 absorption transition is observed for Eu3+ ions. A bit more generous situation is for excitation in the near-UV. In this range the absorption transitions of Eu3+ ions are more abundant, for instance 7F05D4; 5GJ; 5LJ. Furthermore, in host matrices with low-lying charge transfer (CT) bands, these absorption transitions become relatively strong. Such matrices include molybdates [10], tungstates [11], vanadates [12], niobates [13], tantalates [14], etc.
Since europium and most of the other lanthanides are rather expensive, significant research is ongoing in finding lanthanide-free luminescent materials, especially emitting in the red spectral region. The main focus is placed on Mn4+ ions doped into inorganic matrices, which exhibit strong red luminescence. The concept of generating white light with elimination of Eu3+-doped phosphor, which is replaced by red-emitting K2SiF6:Mn4+ phosphor is shown in Figure 1c. Mn4+-doped inorganic materials show efficient and sharp line red luminescence in the 600–750 nm spectral range. However, only phosphors based on fluoride host matrices (for instance, K2SiF6:Mn4+, BaTiF6:Mn4+) possess suitable emission for white LEDs, i.e., below 650 nm [3,15,16].
The last concept to produce white light by employing near-UV-emitting LED chip is using a single-phase white-light-emitting phosphor (see Figure 1d). The advantage of such an approach is that the white light is produced within one phosphor; therefore, this eliminates several problems such as thorough blending of different colour-emitting phosphors particles, reabsorption, and so on. However, there are also some problems relating these phosphors, such as colour shift at different temperatures and so on. Most of the white-light-emitting phosphors are based on energy transfer between Ce3+ → Mn2+, Eu2+ → Mn2+, Ce3+ → Tb3+, Tb3+ → Eu3+ ion pairs or even more complicated systems [3,17].
In this contribution, the red-emitting Ba2Y5B5O17:Eu3+ phosphors were prepared by high temperature solid-state reaction. The optical properties of obtained luminescent materials were analysed with respect to excitation wavelength, Eu3+ concentration, and temperature. The ceramic samples from powders were also prepared in order to investigate possible application as a remote red-emitting phosphor.

2. Materials and Methods

Ba2Y5B5O17:Eu3+ phosphor powders were prepared by two-step solid state synthesis. Firstly, the stoichiometric amounts of raw materials (BaCO3 (99+%) and H3BO3 (99+%) from Acros Organics (Geel, Belgium), Y2O3 (99.99%) and Eu2O3 (99.99%) from Tailorlux (Münster, Germany)) were weighed and blended in an agate mortar with few millilitres of acetone to speed up the homogenization process. When all the acetone evaporated, the dry blend of starting materials was placed to the porcelain crucible and annealed at 450 °C for 4 h in air. Then the powder was reground and sintered at 1200 °C for 8 h in air. Finally, the sintered compounds were again ground to fine powder and used for further measurements. The Eu3+ concentration in given compounds was 0% (undoped sample), 1%, 5%, 10%, 25%, and 50%. Attempts to synthesize compounds with 75% and 100% Eu3+ were not successful.
Ba2Y5B5O17:50%Eu3+ phosphor powder was also pressed to ceramic disks with thicknesses of 0.73, 0.98, and 1.20 mm by applying 30 kN force for 5 min and, subsequently, annealing at 1200 °C for 5 h in air. The diameter of prepared ceramic disks was 8 mm.
X-ray diffraction (XRD) measurements were performed using Rigaku Ultima IV diffractometer (Rigaku, Tokyo, Japan) equipped with Cu anticathode, Soller slits to limit the divergence of X-ray beam and a nickel foil filter to attenuate the Cu Kβ line. XRD patterns were recorded in the range of 15°–80° (scan rate was 0.05°/min) using the Bragg–Brentano configuration.
Field-emission Hitachi SU-70 Scanning Electron Microscope (SEM) machine (Hitachi, Tokyo, Japan) was used to investigate phosphor particle size and morphological features.
IR spectra were recorded in the range of 3000–400 cm−1 using Bruker Alpha Fourier-Transform Infrared (FTIR) spectrometer (Bruker, Ettlingen, Germany) with 4 cm−1 resolution.
Reflection spectra, room temperature and temperature dependent (77–500 K range) excitation and emission spectra, as well as Photoluminescence (PL) decay curves were recorded using Edinburgh Instruments FLS980 spectrometer (Edinburgh Instruments, Livingston, UK). The detailed description of the entire experimental setup is given in our previous publication [10].
The quantum efficiencies (QE) of the synthesized phosphors were determined by integrating sphere method. The equation summarizing the experiment can be written as [10]:
Q E = I e m , s a m p l e I e m , B a S O 4 I r e f , B a S O 4 I r e f , s a m p l e × 100 % = N e m N a b s × 100 %
here I e m , s a m p l e and I e m , B a S O 4 denote the intensity of integrated emission of the sample and barium sulfate, respectively. I r e f , s a m p l e and I r e f , B a S O 4 stand for the integrated reflectance of the sample and barium sulfate BaSO4, respectively. Nem and Nabs are the number of emitted and absorbed photons, respectively. Each measurement was repeated five times in order to get some statistical data.

3. Results and Discussion

The Ba2Y5B5O17 crystal structure was first reported by Hermus et al. in 2016 [18]. The compound crystallizes in primitive orthorhombic crystal structure and adopts Pbcn (#60) space group. There are four formula units per unit cell (Z = 4) and lattice parameters are: a = 17.38257 Å, b = 6.65299 Å, and c = 13.03055 Å.
Powder XRD patterns of synthesized Ba2Y5B5O17:50%Eu3+ and undoped Ba2Y5B5O17 are given in Figure 2a,b, respectively. The broad background is due to glass sample holder. The reference pattern is shown in Figure 2c. Single phase compounds isostructural with Ba2Y5B5O17 were synthesized when Eu3+ concentration has not exceeded 50%. The increase of Eu3+ concentration to 75% resulted in appearing of additional diffraction peaks indicating the presence of a mixture of phases. Furthermore, for 100% substituted compound the phase with Pbcn space group is lost in favour to a mixture of phases, which are difficult to identify. Nevertheless, some diffraction peaks could be attributable to an isostructural structure (with Pnma space group) to Ba3Pr2B4O12 compound [19].
SEM technique was employed to determine the particle size and form of the synthesized materials (see Figure 2d–g). The is virtually no changes in particle size and morphology with increasing Eu3+ concentration. The obtained particles are of irregular shape and formed from agglomerated crystallites.
The body colour of all synthesized Ba2Y5B5O17:Eu3+ was white regardless the Eu3+ concentration. The digital photograph of Ba2Y5B5O17:50%Eu3+ taken under daylight is depicted in Figure 2h. Nevertheless, this sample excited at both 254 and 365 nm radiation showed bright red luminescence as shown in Figure 2i,j, respectively.
In Figure S1, which shows a zoom on the 18°–32° angular range, we can observe a displacement of the diffraction peaks towards the smaller angles indicating an increase of the lattice parameters with the insertion of Eu3+ ions. This evolution is not surprising because the ionic radius of Eu3+ (r(Eu3+)VII = 1.01 Å) is larger than of Y3+ (r(Y3+)VII = 0.96 Å) [20]. The VII coordination is used for the occupied site because Hermus et al. showed that the crystallographic sites occupied by Y3+ mainly have this coordination and the site with X coordination is mainly occupied by Ba2+ [18]. This was confirmed by calculated lattice parameters from XRD data employing the Le Bail fit, which showed that lattice parameters increase linearly with increasing Eu3+ concentration, which is also consistent with Vegard’s law [21] and confirms successful Eu3+ incorporation in this compound. The obtained linear evolution of lattice parameters together with graphical representation of Le Bail fit for Ba2Y5B5O17 compound is given in Figure 3a.
Figure 3b shows the unit cell of Ba2Y5B5O17 along c-axis together with coordination polyhedrons of sites that Eu3+/Y3+ can occupy. There are two sites that Y3+ shares with Ba2+ ions. The first is 10-coordinated and is mostly occupied by Ba2+ ions (Ba1/Y1). The second is mostly occupied by Y3+ ions and is seven-coordinated, forming a distorted pentagonal bipyramid (Ba2/Y2). There are also two independent sites that are occupied solely by Y3+ ions. One of them is six-coordinated forming distorted octahedron (Y3), whereas the other one is seven-coordinated and forms distorted capped trigonal prism (Y4). Boron atoms are three-coordinated and form trigonal planar units that are slightly distorted [18].
The IR spectra of Ba2Y5B5O17:Eu3+ compounds as a function of Eu3+ concentration are given in Figure S2. The spectra up to 75% Eu3+ look very similar. There are several strong absorption band in the range of 1400–400 cm−1. The most intensive absorption bands lie at 1195, 738, 620, 520, and 435 cm−1. Nevertheless, the IR spectra are relatively simple, since there are only BO3 units in the structure. The lines at 1400–1100 cm−1 are assigned to asymmetric stretching of BO3 groups whereas those in the range of 750–400 cm−1 to the ring bending [22,23].
The excitation spectra of Ba2Y5B5O17:Eu3+ compounds doped with 1%, 5%, and 50% Eu3+ ions are given in Figure 4a. There is a broad excitation band ranging from 250 to 320 nm, which can be attributed to charge transfer transition (CT). At longer wavelengths, the typical Eu3+ excitation lines are visible. The excitation originates from the 7F0 and/or 7F1 ground state levels and ends up at 5FJ (ca. 320 nm), 5HJ (ca. 330 nm), 5D4 (ca. 363 nm), 5L7,8; 5GJ (ca. 370–390 nm), 5L6 (ca. 390–405 nm), 5D3 (ca. 410–420 nm), 5D2 (ca. 455–480 nm), 5D1 (ca. 515–545 nm), and 5D0 (ca. 570–600 nm) terminal levels [24,25]. The interesting feature of the given excitation spectra are the abundance of excitation lines, especially for the 7F05D2 transition. This originates from the fact that Eu3+ can occupy four lattice sites with each giving lines at slightly different wavelengths due to the different crystal field generated. This is the opposite to phosphors where only one lattice site is available for Eu3+ ions and there are very few excitation lines which, in turn, are also very narrow [13,26,27,28,29]. The absorption strength of the synthesized phosphors increased with increasing Eu3+ concentration. Since there are more excitation lines in excitation spectra, each of them could be used to excite the Ba2Y5B5O17:Eu3+ phosphor what is beneficial for practical application.
Figure 4b shows the emission spectra of Ba2Y5B5O17:Eu3+ samples doped with 1%, 5%, and 50% of Eu3+ ions for 394 nm excitation. There are five sets of lines visible in the given emission spectra. These emission lines originate from transitions starting form 5D0 excited state to 7F0 (ca. 577–581nm), 7F1 (ca. 582–600 nm), 7F2 (ca. 600–635 nm), 7F3 (ca. 655), and 7F4 (ca. 680–715 nm) terminal levels. It also evident the profile of emission spectra for 394 nm excitation is the same regardless the Eu3+ concentration. The strongest emission intensity was observed for Ba2Y5B5O17:50%Eu3+ sample as shown in the inset of Figure 4b.
Figure 4c–d shows PL decay curves (5D07F2 transition, λem = 615 nm) of Ba2Y5B5O17:Eu3+ samples doped with 1%, 25%, and 50% Eu3+ under excitation with 280, 394, and, 465 nm radiation, respectively. The PL of all samples decays bi-exponentially (except for 280 nm excitation where mono-exponential PL decay was observed until 25% Eu3+ concentration). The PL lifetime values for all samples were calculated according the following equation:
I ( t ) = A + B 1 e t τ 1 + B 2 e t τ 2
here I(t), A, B1 and B2, τ1 and τ2 stand for PL intensity at a given time t, background, constants, and PL lifetime values, respectively. The calculated PL lifetime values together with standard deviations, relative percentage and calculated average PL lifetime values for 280, 394, and 465 nm excitation are tabulated in Tables S1–S3, respectively. Under 280 nm excitation, the average PL lifetime values gradually decrease from 2400 to 1440 μs when changing Eu3+ concentration from 1% to 50%. Slightly different PL lifetime values were obtained under 394 and 465 nm excitation. Here average PL lifetime values increased from ca. 1600 μs (1% Eu3+-doped sample) to ca. 1760 μs (25% Eu3+-doped sample) and then again decreased to ca. 1390 μs when Eu3+ concentration reached 50%.
The reflection spectra of undoped Ba2Y5B5O17 and Ba2Y5B5O17:50%Eu3+ are depicted in Figure 4f. The body colour of Ba2Y5B5O17 powder was white suggesting that this material does not absorb visible light. This goes hand in hand with the respective reflection spectrum where slight absorption starts at wavelengths shorter than 380 nm. The reflection spectrum of Ba2Y5B5O17:50%Eu3+ specimen, in turn, contains several sets of absorption lines in the visible light range and broad absorption band starting at ca. 340 nm and strongly increasing going to shorter wavelengths. This spectrum is very similar to excitation spectra of the given compounds; therefore, the assigned transitions are also the same for both.
The excitation spectra (for 615 nm emission) of Ba2Y5B5O17:50%Eu3+ sample were also measured at 77 and 500 K temperatures. These spectra are depicted in Figure 5a. Despite the similar profile of both spectra there are also some significant changes going on when temperature increases from 77 to 500 K. First of all, at low temperatures the thermal population of 7F1 is very suppressed; therefore, mostly the lines originating from the 7F0 ground state transitions to various excited states are visible. This is best observed for 7F15D3 (ca. 415 nm), 7F15D1 (ca. 535 nm) and 7F15D0 (ca. 590 nm) transitions, which do not occur at 77 K but are relatively strong at 500 K temperature. Moreover, the increase of the width of charge transfer state is also evident at elevated temperatures. This phenomenon was also observed by other researchers for different types of compounds doped with Eu3+ ions [30,31]. The broadening of the CT band can be explained as follows; at low temperatures transition to the CT state starts from one vibrational level of the ground state (ν1); however, at elevated temperatures this transition starts at the higher vibrational levels of the ground state. Therefore, the absorption transition energy decreases leading to the red shift of CT band [31].
Figure 5b shows normalized emission spectra (λex = 394 nm) of Ba2Y5B5O17:50%Eu3+ sample recorded at 77 and 500 K temperatures. Both spectra are nearly identical and the only observable difference is broader emission lines at elevated temperature. The lattice vibrations increase with increasing temperature; thus, the local surrounding of Eu3+ ions changes. This yields slight changes in crystal field strength and emission of Eu3+ ions at somewhat different wavelengths what eventually leads to emission line broadening. Inset in Figure 5b demonstrates temperature dependent PL decay curves (λex = 394 nm, λem = 615 nm) of Ba2Y5B5O17:50%Eu3+ specimen. The exact calculated PL lifetime values are summarized in Table S4. The PL decay curves become steeper at temperatures higher than 250 K showing decreasing PL lifetime values. This can be better appreciated from Figure 5c where PL lifetime values are plotted as a function of temperature. Severe decrease in PL lifetime values is observed when temperature reaches 450 K, indicating the increase of rates of nonradiative transitions.
Temperature dependent emission integrals can be used to evaluate phosphor performance at elevated temperatures. This is very important parameter since modern high power LEDs during operation can heat up to temperatures as high as 400 K and even above. The temperature dependent normalized emission integral values as a function of temperature are given in Figure 5c. These values were also used to determine the TQ1/2 value (temperature, when phosphor emission decreases twice) employing Boltzmann fit [32]. The extracted TQ1/2 value for Ba2Y5B5O17:50%Eu3+ phosphor sample was 516 ± 21 K. Moreover, at 450 K ca. 65% of initial emission still remains indicating relatively high emission stability with increasing temperature.
The quantum efficiency (QE) is also very important parameter of the phosphor if it is considered for practical application. The quantum efficiency values of Ba2Y5B5O17:Eu3+ phosphors as a function of Eu3+ concentration and excitation wavelength are shown in Figure 6a (taking the whole emission integral from 500 to 800 nm). Excitation at 280 nm yielded low quantum efficiencies ca. 30% for samples doped with 5% and 10% Eu3+. Further increase of Eu3+ concentration resulted in gradual decrease of QE to ca. 15% for 50% Eu3+-doped specimen. The low quantum efficiency for 280 nm (CT band) can be explained by the nonradiative relaxation of the excited electron through the charge transfer state [24,33]. Much higher QE values were obtained if 394 and 465 nm radiation was used for sample excitation. In this case QE vales as high as ca. 100% were obtained for samples doped with 10% and 25% Eu3+. The increase of Eu3+ concentration to 50% led to decrease of QE to ca. 80% for 394 nm excitation and ca. 72% for 465 nm excitation. Moreover, since the human eye is very insensitive to wavelengths above 650 nm [8], we have also calculated the “effective” quantum efficiency where we discarded the emission above 650 nm. The obtained values are given in Figure 6b. QE values calculated in this way are, of course, lower than those for which the whole emission was taken into account. However, the overall trend remains the same. The highest QE values are obtained for samples doped with 5–25% Eu3+. For 280 nm excitation the highest “effective” QE was ca. 20% (5% and 10% Eu3+-doped samples); for 395 nm excitation ca. 80% (5% and 10% Eu3+-doped samples); for 465 nm excitation ca. 65% (5–25% Eu3+-doped samples). These results show that the synthesized phosphors have potential to be used in both near-UV and blue InAlGaN chip-driven white LEDs.
We have also recorded the emission spectra (see Figure 6b–d) of Ba2Y5B5O17:Eu3+ samples under 280 nm (CT transition) and 465 nm (7F05D2 transition) excitation and observed some changes in emission spectra profile. Excitation with both 394 nm (Figure 6d) and 465 nm (Figure 6e) wavelength radiation yields virtually identical emission spectra. However, this is not the case for 280 nm excitation as shown in Figure 6c. Here, only the sample with 50% of Eu3+ gives emission spectra similar to those recorded under 394 and 465 nm excitation. Samples doped with low Eu3+ concentrations (1% and 5% in the given case), in turn, yield much different emission spectra, whereas lines originating from 5D07F0, 5D07F1, 5D07F3, and 5D07F4 transitions are much more pronounced. The explanation of such different emission spectra profile could be that, regardless the obtained single phase materials from the XRD data, there is still some small amount of impurity phases not detectable with XRD. These impurity phases would give their own emission spectra, which, of course, are different than for Ba2Y5B5O17:Eu3+. Thus, at low Eu3+ concentrations the emission intensity from these impurity phases would be rather comparable to Ba2Y5B5O17:Eu3+ emission. With increasing Eu3+ concentration, the concentration of impurities would remain the same; therefore, the emission from Ba2Y5B5O17:Eu3+ compounds would start to dominate the emission spectra. Nevertheless, the given two explanations are just assumptions and further spectroscopic investigation is needed in order to clarify this phenomenon.
In inorganic matrices the absorption of rare earth ions is relatively weak because the transitions within 4f orbitals are forbidden [34]. As was already mentioned, there are several classes of inorganic compounds, for instance, molybdates, tungstates, vanadates, etc., with low-lying charge transfer states, which spectrally overlap with energy levels of rare earth ions resulting in stronger emission due to energy transfer [35]. Preparation of ceramic discs from phosphor powder can also be one of the ways to increase the absorption efficiency due to increased penetration depth of the incident light. This approach also reduces the temperature of the phosphor since the ceramic layer is deposited further from hot InAlGaN chip [36]. To test this approach, we have prepared three ceramic disks (the thickness of disks was 0.73, 0.98, and 1.20 mm) from Ba2Y5B5O17:50%Eu3+ sample, placed them on 375, 400, and 455 nm emitting LEDs and measured emission spectra of the obtained light sources.
The digital image of 0.73-mm-thick ceramic disk is given in Figure S3a. Here we need to note, that the prepared ceramics are not transparent, but translucent instead. The image of ceramic disks with all three thicknesses under 365 nm excitation is shown in Figure S3b. Finally, the image of the 1.20-mm-thick Ba2Y5B5O17:50%Eu3+ ceramic disk illuminated with the 400 nm emitting LED from below is given in Figure S3c. The pinkish red colour was obtained due to mixing of the red light emitted by the ceramic disk and passed through unabsorbed light form 400 nm emitting LED.
The emission spectra of sole LEDs are given in Figure 7a, whereas Figure 7b shows the emission spectra of these LEDs with the thickest Ba2Y5B5O17:50%Eu3+ ceramic disk on top. The absorption of the LED emitted light increased with increasing ceramic thickness; therefore, only 1.2 mm ceramic disk properties will be analyzed further. It is evident that 1.2-mm-thick ceramic disk is still not capable to absorb all the incident light of either investigated LED. However, this is not necessarily a drawback, since the unabsorbed incident light can be used for excitation of phosphors emitting other colours. It is also clear that the ceramic disks absorb light emitted by 400 nm most efficiently. This, however, is not surprising, since there are a lot of absorption lines of Eu3+ ions in this spectral region. The light emitted by 375 nm LED is also efficiently absorbed. In both cases the emission spectrum of the ceramic disk is identical. The worst absorption was observed of 455 nm LED emitted light. In this area there is only one absorption line originating from the 7F05D2 transition; therefore, most of light emitted by LED passes through the ceramic disk unabsorbed. However, the absorption of 455 nm emitting LED radiation is still stronger if compared to ceramic disks prepared from other materials [10,32,37]. This arises from the fact that, as was already discussed above, there are four different lattice in Ba2Y5B5O17 compound that Eu3+ can occupy; thus, simply, there are more lines originating from the same transition.
The CIE 1931 colour coordinates and luminous efficacy (LE) values were calculated from the respective emission spectra. The fragments of CIE 1931 colour space with colour coordinates of Ba2Y5B5O17:Eu3+ as a function of Eu3+ concentration and temperature dependent colour coordinates of Ba2Y5B5O17:50%Eu3+ sample are given in Figure 8a,b, respectively. These colour coordinates are located directly on the edge of the CIE 1931 colour space diagram, indicating that the emission spectra are perceived as a monochromatic light by human eye. Colour coordinates also tend to move to more orange spectral region when Eu3+ concentration or phosphor temperature is increased. On the other hand, the shift is very small; thus colour coordinates can be considered as stable. The exact calculated values together with LE values as a function of Eu3+ concentration and excitation wavelength are summarized in Table S5. The LE values of all the samples are relatively the same and fluctuate around 240 lm/Wopt. This could be expected, since emission spectra do not change much with increasing Eu3+ concentration and excitation wavelength. The largest change in LE values was observed if samples were excited with 280 nm. In this case, LE decreased from 250 lm/Wopt (for 1% Eu3+-doped sample) to 243 lm/Wopt (for 50% Eu3+-doped sample). This is in good agreement with emission spectra shown in Figure 6b. With increasing Eu3+ concentration, the intensity of lines (5D07F1 transition) in the orange spectral region decreases; thus the LE values decrease, because human eye is more sensitive to the orange light [38]. Moreover, the LE values of the synthesized Ba2Y5B5O17:50%Eu3+ phosphors are higher or very similar to the ones reported for some well-established red-emitting Eu2+-doped phosphors, namely, Sr2Si5N8:Eu2+em = 620 nm; LE = 240 lm/Wopt), CaAlSiN3:Eu2+em = 650 nm; LE = 150 lm/Wopt), and CaS:Eu2+em = 650 nm; LE = 85 lm/Wopt) [9]. However, they are slightly lower if compared to other Eu3+ phosphors possessing lower intensity of 5D07F4 transition ca. 700 nm, for instance, Li3Ba2Eu3(MoO4)8em = 615 nm; LE = 312 lm/Wopt) [39] or LiLa(MoO4)2:Eu3+em = 616 nm; LE = 280 lm/Wopt) [40].
The CIE 1931 chromaticity coordinates for light sources obtained by combining 375, 400, and 455 nm emitting LEDs with various thicknesses of Ba2Y5B5O17:50%Eu3+ ceramic disks were also calculated and presented in Figure 8c–e, respectively. When 375 nm LED is used for excitation, the red light source is obtained with colour coordinates located close to the edge of the chromaticity diagram. However, this is not the case for other two LEDs used for excitation. Combination of 400 nm emitting LED with Ba2Y5B5O17:50%Eu3+ ceramic disks yielded light sources emitting in the purple region due to presence of red light emitted by ceramics and unabsorbed violet light emitted by LED. The colour coordinates shifted towards red region if the thickness of ceramics was increased. Similarly, only the tints of blue light were obtained if 455 nm LED was used to excite the ceramic disks, because only very small fraction of 455 nm LED emission was absorbed by ceramic disks regardless their thickness. This clearly shows, that 455 nm emitting LED is not suitable for excitation of these materials. The exact calculated CIE 1931 colour coordinates together with LE values are given in Table S7. The calculated LE values are relatively the same for the light sources obtained by combining Ba2Y5B5O17:50%Eu3+ ceramic disks with 375 and 400 nm emitting LEDs. The LE values increase from ca. 150 to ca. 185 lm/Wopt for the thinnest and the thickest ceramics, respectively. This is related to decreasing emission fraction of LED to which human eye is extremely insensitive. Furthermore, even lower LE values (ca. 70 lm/Wopt) were obtained if 455 nm emitting LED was used to excite ceramic disks. As mentioned above, only a small fraction of this LED is absorbed by the ceramic disk. This results in a situation when emission spectra of the light source contain the strongest emission in the blue and red spectral regions. Human eye is not very sensitive in either of these regions and thus the LE values of such light source are extremely low. Similar results were also reported earlier for ceramic disks prepared from other compounds [10,32,37].

4. Conclusions

Phase pure Ba2Y5B5O17:Eu3+ red-emitting phosphors were prepared by high temperature solid state reaction method. The solubility of Eu3+ in the given host matrix was found to be 50% with respect to Y3+ ions. All samples exhibited bright red luminescence if excited with UV, near-UV and blue radiation. The emission spectra were dominated by 5D07F2 and 5D07F4 transitions of Eu3+ ions at ca. 615 and 705 nm, respectively. The synthesized phosphors possess high colour purity and luminous efficacy values (ca. 240 lm/Wopt). Moreover, the 50% Eu3+-doped sample lost half of the efficiency only at ca. 500 K, showing high thermal stability. The prepared Ba2Y5B5O17:50%Eu3+ ceramic disks showed increasing absorption of near-UV and blue radiation with increasing thickness of the ceramic disk. It was also observed that near-UV radiation is more efficiently absorbed by ceramic disks if compared to the blue radiation. This is due to more abundant absorption transitions of Eu3+ ions in this spectral area. Finally, samples doped with 5%, 10%, and 25% Eu3+ showed quantum efficiency close to 100% if excited with 394 nm radiation what is a huge benefit for practical application of the synthesized phosphors. The “effective” QE values, obtained by discarding emission above 650 nm, were as high as 80% for 394 nm excitation.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1944/13/3/763/s1, Figure S1: XRD patterns of Ba2Y5B5O17: Eu3+ phosphors, Figure S2: FTIR spectra of Ba2Y5B5O17: Eu3+ phosphors, Figure S3. Digital images of: 0.73-mm-thick Ba2Y5B5O17:50%Eu3+ ceramic disk under daylight (a); 0.73, 0.98, and 1.20-mm-thick (from left to right) Ba2Y5B5O17:50%Eu3+ ceramic disks under 365 nm excitation (b); 1.20-mm-thick on top of 400 nm emitting LED (c), Table S1: PL lifetime values of Ba2Y5B5O17:Eu3+ phosphors as a function of Eu3+ concentration (λex = 280 nm, λem = 615 nm), Table S2: PL lifetime values of Ba2Y5B5O17:Eu3+ phosphors as a function of Eu3+ concentration (λex = 394 nm, λem = 615 nm), Table S3: PL lifetime values of Ba2Y5B5O17:Eu3+ phosphors as a function of Eu3+ concentration (λex = 465 nm, λem = 615 nm), Table S4: PL lifetime values of Ba2Y5B5O17:50%Eu3+ as a function of temperature (λex = 394 nm, λem = 615 nm), Table S5: CIE 1931 colour coordinates and luminous efficacies (LE) of synthesized phosphors as a function of Eu3+ concentration and excitation wavelength, Table S6: CIE 1931 colour coordinates and luminous efficacies (LE) of Ba2Y5B5O17:50%Eu3+ as a function of temperature (λex = 394 nm), Table S7: CIE 1931 colour coordinates and luminous efficacies (LE) of different thicknesses Ba2Y5B5O17:50%Eu3+ ceramics mounted on 375, 400, and 455 nm LEDs.

Author Contributions

Conceptualization, A.K.; investigation, E.E., J.G., A.M. and S.S.; resources, A.K.; writing—original draft preparation, E.E. and J.G.; writing—review and editing, A.K.; visualization, E.E., J.G. and S.S.; funding acquisition, A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant (No. S-MIP-17-48) from the Research Council of Lithuania.

Acknowledgments

The authors are indebted to Rokas Vargalis (Vilnius University) for taking SEM images.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhou, Q.; Dolgov, L.; Srivastava, A.M.; Zhou, L.; Wang, Z.; Shi, J.; Dramićanin, M.D.; Brik, M.G.; Wu, M. Mn2+ and Mn4+ red Phosphors: Synthesis, luminescence and applications in WLEDs. A review. J. Mater. Chem. C 2018, 6, 2652–2671. [Google Scholar] [CrossRef]
  2. Li, G.; Tian, Y.; Zhao, Y.; Lin, J. Recent progress in luminescence tuning of Ce3+ and Eu2+-activated phosphors for pc-WLEDs. Chem. Soc. Rev. 2015, 44, 8688–8713. [Google Scholar] [CrossRef] [PubMed]
  3. Nair, G.B.; Swart, H.C.; Dhoble, S.J. A review on the advancements in phosphor-converted light emitting diodes (pc-LEDs): Phosphor synthesis, device fabrication and characterization. Prog. Mater. Sci. 2019, 109, 100622. [Google Scholar] [CrossRef]
  4. Jeet, S.; Pandey, O.P. Template free synthesis route to monophasic BaMgAl10O17:Eu2+ with high luminescence efficiency. J. Alloys Compd. 2018, 750, 85–91. [Google Scholar] [CrossRef]
  5. Xia, Z.; Meijerink, A. Ce3+-Doped garnet phosphors: composition modification, luminescence properties and applications. Chem. Soc. Rev. 2017, 46, 275–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Piao, X.; Machida, K.; Horikawa, T.; Hanzawa, H.; Shimomura, Y.; Kijima, N. Preparation of CaAlSiN3:Eu2+ Phosphors by the Self-Propagating High-Temperature Synthesis and Their Luminescent Properties. Chem. Mater. 2007, 19, 4592–4599. [Google Scholar] [CrossRef]
  7. Li, Y.Q.; van Steen, J.E.J.; van Krevel, J.W.H.; Botty, G.; Delsing, A.C.A.; DiSalvo, F.J.; de With, G.; Hintzen, H.T. Luminescence properties of red-emitting M2Si5N8:Eu2+ (M=Ca, Sr, Ba) LED conversion phosphors. J. Alloys Compd. 2006, 417, 273–279. [Google Scholar] [CrossRef]
  8. Zukauskas, A.; Vaicekauskas, R.; Ivanauskas, F.; Vaitkevicius, H.; Shur, M.S. Spectral Optimization of Phosphor-Conversion Light-Emitting Diodes for Ultimate Color Rendering. Appl. Phys. Lett. 2008, 93, 051115. [Google Scholar] [CrossRef]
  9. Jüstel, T. Luminescent Materials for Phosphor-Converted LEDs. In Luminescence: from Theory to Applications; Ronda, C.R., Ed.; Wiley-VCH: Weinheim, Germany, 2008; p. 179. [Google Scholar]
  10. Grigorjevaite, J.; Ezerskyte, E.; Minderyte, A.; Stanionyte, S.; Juskenas, R.; Sakirzanovas, S.; Katelnikovas, A. Optical Properties of Red-Emitting Rb2Bi(PO4)(MoO4):Eu3+ Powders and Ceramics with High Quantum Efficiency for White LEDs. Materials 2019, 12, 3275. [Google Scholar] [CrossRef] [Green Version]
  11. Zhang, L.; Deng, B.; Shu, S.; Wang, Y.; Geng, H.; Yu, R. Preparation, characterization, and luminescence properties of BiLaWO6:Eu3+ red-emitting phosphors for w-LEDs. Spectrochim. Acta, Part A 2020, 224, 117454. [Google Scholar] [CrossRef]
  12. Colmont, M.; Saitzek, S.; Katelnikovas, A.; Kabbour, H.; Olchowka, J.; Roussel, P. Host-sensitized luminescence properties of KLa5O5(VO4)2:Eu3+ for solid-state lighting applications. J. Mater. Chem. C 2016, 4, 7277–7285. [Google Scholar] [CrossRef]
  13. Mackevic, I.; Grigorjevaite, J.; Janulevicius, M.; Linkeviciute, A.; Sakirzanovas, S.; Katelnikovas, A. Synthesis and optical properties of highly efficient red-emitting K2LaNb5O15:Eu3+ phosphors. Opt. Mater. 2019, 89, 25–33. [Google Scholar] [CrossRef]
  14. Zhong, Y.; Sun, P.; Gao, X.; Liu, Q.; Huang, S.; Liu, B.; Deng, B.; Yu, R. Synthesis and optical properties of new red-emitting SrBi2Ta2O9:Eu3+ phosphor application for w-LEDs commercially based on InGaN. J. Lumin. 2019, 212, 45–51. [Google Scholar] [CrossRef]
  15. Adachi, S. Review—Mn4+-Activated Red and Deep Red-Emitting Phosphors. ECS J. Solid State Sci. Technol. 2020, 9, 016001. [Google Scholar] [CrossRef]
  16. Li, Y.; Qi, S.; Li, P.; Wang, Z. Research progress of Mn doped phosphors. RSC Adv. 2017, 7, 38318–38334. [Google Scholar] [CrossRef] [Green Version]
  17. Ye, S.; Xiao, F.; Pan, Y.X.; Ma, Y.Y.; Zhang, Q.Y. Phosphors in phosphor-converted white light-emitting diodes Recent advances in materials, techniques and properties. Mater. Sci. Eng. R 2010, 71, 1–34. [Google Scholar] [CrossRef]
  18. Hermus, M.; Phan, P.-C.; Brgoch, J. Ab Initio Structure Determination and Photoluminescent Properties of an Efficient, Thermally Stable Blue Phosphor, Ba2Y5B5O17:Ce3+. Chem. Mater. 2016, 28, 1121–1127. [Google Scholar] [CrossRef]
  19. Khamaganova, T.N.; Trunov, V.K.; Dzhurinskii, B.F.; Efremov, V.A. The crystal structures of Ba3TR2(BO3)4 (TR= La, Pr). Kristallografiya 1990, 35, 856–860. [Google Scholar]
  20. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  21. Ropp, R.C. Design of Phosphors. Luminescence and the Solid State, 2nd ed.; Elsevier: Amsterdam, The Netherland, 2004; p. 447. [Google Scholar]
  22. Weir, C.E.; Schroeder, R.A. Infrared Spectra of the Crystalline Inorganic Borates. J. Res. Natl. Stand. Sec. A 1964, 68, 465–487. [Google Scholar] [CrossRef]
  23. Peak, D.; Luther, G.W.; Sparks, D.L. ATR-FTIR spectroscopic studies of boric acid adsorption on hydrous ferric oxide. Geochim. Cosmochim. Acta 2003, 67, 2551–2560. [Google Scholar] [CrossRef]
  24. Binnemans, K. Interpretation of Europium (III) Spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef] [Green Version]
  25. Carnall, W.T.; Crosswhite, H.; Crosswhite, H.M. Energy Level Structure and Transition Probabilities in the Spectra of the Trivalent Lanthanides in LaF3; ANL-78-XX-95; Argonne National Laboratory: Lemont, IL, USA, 1978. [Google Scholar]
  26. Kruopyte, A.; Giraitis, R.; Juskenas, R.; Enseling, D.; Jüstel, T.; Katelnikovas, A. Luminescence and luminescence quenching of efficient GdB5O9:Eu3+ red phosphors. J. Lumin. 2017, 192, 520–526. [Google Scholar] [CrossRef]
  27. Li, C.; Zhang, C.; Hou, Z.; Wang, L.; Quan, Z.; Lian, H.; Lin, J. β-NaYF4 and β-NaYF4:Eu3+ Microstructures: Morphology Control and Tunable Luminescence Properties. J. Phys. Chem. C 2009, 113, 2332–2339. [Google Scholar] [CrossRef]
  28. Yan, M.; Liu, G.; Wen, J.; Wang, Y. Blue-Light-Excited Eu3+/Sm3+ Co-Doped NaLa(MoO4)2 Phosphors: Synthesis, Characterizations and Red Emission Enhancement for WLEDs. Materials 2018, 11, 1090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Pawlik, N.; Szpikowska-Sroka, B.; Pisarska, J.; Goryczka, T.; Pisarski, W.A. Reddish-Orange Luminescence from BaF2:Eu3+ Fluoride Nanocrystals Dispersed in Sol-Gel Materials. Materials 2019, 12, 3735. [Google Scholar] [CrossRef] [Green Version]
  30. Baur, F.; Glocker, F.; Jüstel, T. Photoluminescence and Energy Transfer Rates and Efficiencies in Eu3+ Activated Tb2Mo3O12. J. Mater. Chem. C 2015, 3, 2054–2064. [Google Scholar] [CrossRef] [Green Version]
  31. Shi, L.; Zhang, H.; Li, C.; Su, Q. Eu3+ doped Sr2CeO4 phosphors for thermometry: Single-color or two-color fluorescence based temperature characterization. RSC Adv. 2011, 1, 298–304. [Google Scholar] [CrossRef]
  32. Grigorjevaite, J.; Katelnikovas, A. Luminescence and Luminescence Quenching of K2Bi(PO4)(MoO4):Eu3+ Phosphors with Efficiencies Close to Unity. ACS Appl. Mater. Interfaces 2016, 8, 31772–31782. [Google Scholar] [CrossRef]
  33. Struck, C.W.; Fonger, W.H. Role of the charge-transfer states in feeding and thermally emptying the 5D states of Eu+3 in yttrium and lanthanum oxysulfides. J. Lumin. 1970, 1, 456–469. [Google Scholar] [CrossRef]
  34. Yen, W.M.; Shionoya, S.; Yamamoto, H. Principal Phosphor Materials and Their Optical Properties. Fundamentals of Phosphors; CRC Press: Boca Raton, FL, USA, 2007; p. 335. [Google Scholar]
  35. Dexter, D.L. A Theory of Sensitized Luminescence in Solids. J. Chem. Phys. 1953, 21, 836–850. [Google Scholar] [CrossRef]
  36. Abe, S.; Joos, J.J.; Martin, L.I.D.J.; Hens, Z.; Smet, P.F. Hybrid remote quantum dot/powder phosphor designs for display backlights. Light Sci. Appl. 2017, 6, e16271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Janulevicius, M.; Marmokas, P.; Misevicius, M.; Grigorjevaite, J.; Mikoliunaite, L.; Sakirzanovas, S.; Katelnikovas, A. Luminescence and Luminescence Quenching of Highly Efficient Y2Mo4O15:Eu3+ Phosphors and Ceramics. Sci. Rep. 2016, 6, 26098. [Google Scholar] [CrossRef] [PubMed]
  38. Smet, P.F.; Parmentier, A.B.; Poelman, D. Selecting Conversion Phosphors for White Light-Emitting Diodes. J. Electrochem. Soc. 2011, 158, R37–R54. [Google Scholar] [CrossRef] [Green Version]
  39. Katelnikovas, A.; Plewa, J.; Sakirzanovas, S.; Dutczak, D.; Enseling, D.; Baur, F.; Winkler, H.; Kareiva, A.; Jüstel, T. Synthesis and Optical Properties of Li3Ba2La3(MoO4)8: Eu3+ Powders and Ceramics for pcLEDs. J. Mater. Chem. 2012, 22, 22126–22134. [Google Scholar] [CrossRef]
  40. Uhlich, D. Kristallographische und spektroskopische Untersuchungen an Eu3+-dotierten Molybdaten als potentielle Konverter für LEDs. Ph.D. Thesis, Universität Osnabrück, Osnabrück, Germany, 21 April 2009. [Google Scholar]
Figure 1. Normalized emission spectra of near-UV LED (λem = 380 nm) based white light sources with all broadband-emitting phosphors (a), line-emitting red phosphor (b,c) and single-phase white-light-emitting phosphor (d). Here blue phosphor is BaMgAl10O17:Eu2+; green—Ba2SiO4:Eu2+; red—Li3Ba2Eu3(MoO4)8 (b) and K2SiF6:Mn4+ (c); white—Ba3MgSi2O8:Eu2+, Mn2+ (d).
Figure 1. Normalized emission spectra of near-UV LED (λem = 380 nm) based white light sources with all broadband-emitting phosphors (a), line-emitting red phosphor (b,c) and single-phase white-light-emitting phosphor (d). Here blue phosphor is BaMgAl10O17:Eu2+; green—Ba2SiO4:Eu2+; red—Li3Ba2Eu3(MoO4)8 (b) and K2SiF6:Mn4+ (c); white—Ba3MgSi2O8:Eu2+, Mn2+ (d).
Materials 13 00763 g001
Figure 2. XRD patterns of Ba2Y5B5O17:50%Eu3+ (a), Ba2Y5B5O17 (b), and reference pattern of Ba2Y5B5O17 (c). SEM images of Ba2Y5B5O17:50%Eu3+ (d,e), and Ba2Y5B5O17 (f,g). Digital photographs of Ba2Y5B5O17:50%Eu3+ at daylight (h), and under 254 nm (i) and 365 nm (j) excitation.
Figure 2. XRD patterns of Ba2Y5B5O17:50%Eu3+ (a), Ba2Y5B5O17 (b), and reference pattern of Ba2Y5B5O17 (c). SEM images of Ba2Y5B5O17:50%Eu3+ (d,e), and Ba2Y5B5O17 (f,g). Digital photographs of Ba2Y5B5O17:50%Eu3+ at daylight (h), and under 254 nm (i) and 365 nm (j) excitation.
Materials 13 00763 g002
Figure 3. Unit cell parameters of Ba2Y5B5O17:Eu3+ as a function of Eu3+ concentration derived from the Le Bail fit and graphical representation of Ba2Y5B5O17 Le Bail fit (a). Unit cell of Ba2Y5B5O17 along the c-axis and coordination polyhedrons that Eu3+ ions could possibly occupy (b).
Figure 3. Unit cell parameters of Ba2Y5B5O17:Eu3+ as a function of Eu3+ concentration derived from the Le Bail fit and graphical representation of Ba2Y5B5O17 Le Bail fit (a). Unit cell of Ba2Y5B5O17 along the c-axis and coordination polyhedrons that Eu3+ ions could possibly occupy (b).
Materials 13 00763 g003
Figure 4. Eu3+ concentration dependent excitation (λem = 615 nm) (a) and emission (λex = 394 nm) (b) spectra of Ba2Y5B5O17:Eu3+ phosphors. Inset in (b) shows integrated emission intensity as a function of Eu3+ concentration. PL decay curves (λem = 615 nm) of samples doped with 1%, 25%, and 50% Eu3+ recorded under different excitation wavelengths: 280 nm (c), 394 nm (d), and 465 nm (e). Reflection spectra of undoped and 50% Eu3+-doped Ba2Y5B5O17 specimens (f).
Figure 4. Eu3+ concentration dependent excitation (λem = 615 nm) (a) and emission (λex = 394 nm) (b) spectra of Ba2Y5B5O17:Eu3+ phosphors. Inset in (b) shows integrated emission intensity as a function of Eu3+ concentration. PL decay curves (λem = 615 nm) of samples doped with 1%, 25%, and 50% Eu3+ recorded under different excitation wavelengths: 280 nm (c), 394 nm (d), and 465 nm (e). Reflection spectra of undoped and 50% Eu3+-doped Ba2Y5B5O17 specimens (f).
Materials 13 00763 g004
Figure 5. Excitation (λem = 615 nm) (a) and emission (λex = 394 nm) (b) spectra of Ba2Y5B5O17:50%Eu3+ at 77 and 500 K temperatures. Inset in (b) shows temperature dependent PL decay curves (λex = 394 nm, λem = 615 nm) of Ba2Y5B5O17:50%Eu3+. Temperature dependent emission integrals with Boltzmann fit and PL lifetime values (c).
Figure 5. Excitation (λem = 615 nm) (a) and emission (λex = 394 nm) (b) spectra of Ba2Y5B5O17:50%Eu3+ at 77 and 500 K temperatures. Inset in (b) shows temperature dependent PL decay curves (λex = 394 nm, λem = 615 nm) of Ba2Y5B5O17:50%Eu3+. Temperature dependent emission integrals with Boltzmann fit and PL lifetime values (c).
Materials 13 00763 g005
Figure 6. Quantum efficiencies of Ba2Y5B5O17:Eu3+ samples as a function of Eu3+ concentration and excitation wavelength for 500–800 nm range (a) and 500–650 nm range (b). Emission spectra of 1%, 5% and 50% Eu3+-doped specimens under 280 nm (c), 394 nm (d), and 465 nm (e) excitation wavelength.
Figure 6. Quantum efficiencies of Ba2Y5B5O17:Eu3+ samples as a function of Eu3+ concentration and excitation wavelength for 500–800 nm range (a) and 500–650 nm range (b). Emission spectra of 1%, 5% and 50% Eu3+-doped specimens under 280 nm (c), 394 nm (d), and 465 nm (e) excitation wavelength.
Materials 13 00763 g006
Figure 7. Emission spectra of 375, 400, and 455 nm emitting LEDs (a). Emission spectra of 1.20-mm-thick Ba2Y5B5O17:50%Eu3+ ceramic disk excited with 375, 400, and 455 nm emitting LEDs (b).
Figure 7. Emission spectra of 375, 400, and 455 nm emitting LEDs (a). Emission spectra of 1.20-mm-thick Ba2Y5B5O17:50%Eu3+ ceramic disk excited with 375, 400, and 455 nm emitting LEDs (b).
Materials 13 00763 g007
Figure 8. Magnified sections of CIE 1931 colour space diagram showing colour coordinates of Ba2Y5B5O17:Eu3+ samples as a function of Eu3+ concentration (λex = 394 nm) (a) and temperature (λex = 394 nm) (b) for Ba2Y5B5O17:50%Eu3+ specimen. Enlarged areas (c), (d), and (f) show colour coordinates of Ba2Y5B5O17:50%Eu3+ ceramic disks (0.73, 0.98, and 1.20 mm in thickness) excited with 375, 400, and 455 nm emitting LEDs, respectively.
Figure 8. Magnified sections of CIE 1931 colour space diagram showing colour coordinates of Ba2Y5B5O17:Eu3+ samples as a function of Eu3+ concentration (λex = 394 nm) (a) and temperature (λex = 394 nm) (b) for Ba2Y5B5O17:50%Eu3+ specimen. Enlarged areas (c), (d), and (f) show colour coordinates of Ba2Y5B5O17:50%Eu3+ ceramic disks (0.73, 0.98, and 1.20 mm in thickness) excited with 375, 400, and 455 nm emitting LEDs, respectively.
Materials 13 00763 g008

Share and Cite

MDPI and ACS Style

Ezerskyte, E.; Grigorjevaite, J.; Minderyte, A.; Saitzek, S.; Katelnikovas, A. Temperature-Dependent Luminescence of Red-Emitting Ba2Y5B5O17: Eu3+ Phosphors with Efficiencies Close to Unity for Near-UV LEDs. Materials 2020, 13, 763. https://doi.org/10.3390/ma13030763

AMA Style

Ezerskyte E, Grigorjevaite J, Minderyte A, Saitzek S, Katelnikovas A. Temperature-Dependent Luminescence of Red-Emitting Ba2Y5B5O17: Eu3+ Phosphors with Efficiencies Close to Unity for Near-UV LEDs. Materials. 2020; 13(3):763. https://doi.org/10.3390/ma13030763

Chicago/Turabian Style

Ezerskyte, Egle, Julija Grigorjevaite, Agne Minderyte, Sebastien Saitzek, and Arturas Katelnikovas. 2020. "Temperature-Dependent Luminescence of Red-Emitting Ba2Y5B5O17: Eu3+ Phosphors with Efficiencies Close to Unity for Near-UV LEDs" Materials 13, no. 3: 763. https://doi.org/10.3390/ma13030763

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop