Next Article in Journal
Thermal Stability and Flammability Behavior of Poly(3-hydroxybutyrate) (PHB) Based Composites
Next Article in Special Issue
A Terahertz (THz) Single-Polarization-Single-Mode (SPSM) Photonic Crystal Fiber (PCF)
Previous Article in Journal
Durability of Gadolinium Zirconate/YSZ Double-Layered Thermal Barrier Coatings under Different Thermal Cyclic Test Conditions
Previous Article in Special Issue
Efficient 1 µm Laser Emission of Czochralski-Grown Nd:LGSB Single Crystal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multicolor and Warm White Emissions with a High Color Rendering Index in a Tb3+/Eu3+-Codoped Phosphor Ceramic Plate

by
Haggeo Desirena
1,*,
Jorge Molina-González
1,
Octavio Meza
2,
Priscilla Castillo
1 and
Juan Bujdud-Pérez
1
1
Centro de Investigaciones en Óptica A.C., A.P. 1-948, León 37150, Mexico
2
Instituto de Física, “Ing. Luis Rivera Terrazas’’, Benemérita Universidad Autónoma de Puebla, Apartado Postal J-48, Puebla 72570, Mexico
*
Author to whom correspondence should be addressed.
Materials 2019, 12(14), 2240; https://doi.org/10.3390/ma12142240
Submission received: 12 April 2019 / Revised: 9 May 2019 / Accepted: 9 May 2019 / Published: 11 July 2019
(This article belongs to the Special Issue Photonic Materials and Devices)

Abstract

:
A series of Tb3+/Eu3+-codoped phosphor ceramic plates with a high color rendering index (CRI) for a near-ultraviolet light emitting diode (LED) were fabricated. Color emission can be tuned from green to reddish as a function of Eu3+ concentration. By doping only 0.15 mol% of Eu3+ concentration, a comfortable warm white emission is promoted as a result of simultaneous emissions of Tb3+ and Eu3+ ions. A theoretical model is proposed to calculate the contributions of the emitted color of the donor (Tb3+) and acceptor (Eu3+) ions in terms of europium concentration. The energy transfer from Tb3+ to Eu3+ ions is corroborated by the luminescence spectra and decay time of Tb3+, with a maximum energy transfer efficiency of 76% for 28 mol% of Tb3+ and 14 mol% of Eu3+. Warm white LEDs were constructed using a 380 nm UV chip and showed a CRI of 82.5, which was one of highest values reported for Tb3+/Eu3+-codoped samples. Color-correlated temperature (CCT), color coordinate (CC), and luminous efficacy (LE) were utilized to know the potentials as a phosphor converter in solid-state lighting.

1. Introduction

In the last decade, phosphor-converted light emitting diodes (pcLEDs) have been introduced into the market to conquer the deficiencies of fluorescent and incandescent lamps. Current pcLEDs are basically a mixture of a phosphor converter and a silicon binder placed in the top of a blue LED chip (InGaN), and the combination of both colors yields white light. However, silicon binders that support the phosphor still present innate issues, such yellowing problems when exposed to a high density of energy caused by their low thermal conductivities [1]. In that sense, luminescence glasses, glass ceramics, phosphor in glass, and ceramic phosphor plates have emerged as good candidates to replace the silicon binder in high-power LEDs [2,3,4,5]. Among these approaches, special attention has concentrated on ceramic phosphor plates because of their superior thermal stability over different approaches for solid-state lighting [6]. It exhibits robustness, a homogenous luminescence emission, chemical stability, ageing resistance, and good thermal conductivity. The last property is key for the development of stable high-power LEDs and can be controlled through network formers, network modifiers, and network intermediators [7,8]. Even though this system shows attractive thermal and optical properties, most of the high-power systems using one single phosphor do not meet the characteristics for satisfactory color rendering index (CRI) and low color-correlated temperature (CCT) for residential lighting applications [9,10]. In that sense, a second phosphor to compensate the weak red emission of YAG:Ce3+ has been introduced, and the combination of both phosphors leads to warm white devices with low CCT and improved CRI [11].
Terbium (Tb3+) and europium (Eu3+) ions show prominent emission peaks in green and red regions, respectively. Such ions have been often used as phosphors in cathode ray tubes, plasma display panels, X-ray scintillation screens, and fluorescent lamps [12]. Recently, Eu3+-activated phosphors have been proposed as a new red emitter to compensate for the weak red emission of YAG:Ce3+-based pcLEDs. Results show that by choosing a proper host matrix, Eu3+ presents superior photometric properties compared to commercial CASN:Eu2+ phosphors under blue and UV excitation [13,14]. The narrow reabsorption of Eu3+ is the main advantage over nitride phosphors.
Tb3+/Eu3+-codoped materials such as glass, glass ceramic, and phosphor powder have been demonstrated to produce warm and cold white emission through energy transfer from Tb3+ to Eu3+ ions. Indeed, color emission and full conversion from green to red can be achieved by changing the ratio of the Tb3+ and Eu3+ concentration [15,16,17,18,19,20,21]. However, energy transfer efficiency and radiative decay probability depend strongly on the phonon energy of the host. Therefore, different promising host materials, such as oxides, oxysulfides, oxynitrides, chloride, and fluoride, have been investigated to develop efficient phosphors [22,23,24,25,26]. In that sense, special attention should be given to inorganic fluorides because of their low phonon energies that promote a decrease in the nonradiative decay probability. Presently, an NaYF4 fluoride crystal is one of the most efficient materials to produce upconversion emission as well as good chemical stability. However, tetragonal-phase LiYF4 crystals exceed the optical properties of NaYF4 because they promote stronger upconversion emissions and generate UV emission lines [27]. The crystalline phases of several LiYF4 samples have been synthetized by thermal decomposition, sol-gel, and Bridgman methods [28,29,30]. However, with these approaches it is impossible to sinter a phosphor plate. In that sense, a melt-quenching method shows greater potential than the methods mentioned above to synthetize phosphor plates with the desired shape and crystalline phase. These properties make LiYF4 an excellent host for Tb3+/Eu3+ ions to be used in solid-state lighting applications.
In spite of several works that show that Tb3+/Eu3+-codoped materials are good candidates for solid-state lighting applications, most of the published reports only present spectroscopic and structural properties. However, luminous efficacy, lighting properties, and device prototypes based in Tb3+/Eu3+-codoped material for solid-state lighting have been seldom reported [31]. In this work, the color tuning of a Tb3+/Eu3+ ceramic phosphor plate as a function of Eu3+ concentration has been investigated. Color coordinate and color rendering indexes were measured to evaluate the potential of the phosphor converter in solid-state lighting.

2. Materials and Methods

Several LiY1−x−yF4:yTb3+/xEu3+ ceramic phosphor plates were prepared by a melt-quenching technique (where y = 28 mol% and x = 0.15, 0.3, 1.5, 8.5, and 14 mol%). In the composition, 5 mol% extra of TeO2 was used as a sintering additive to promote the formation of a ceramic plate. The total lanthanide concentration in the host matrix and other optical properties are show in Table 1. All reactants were 99.99% pure and used as received. The reactants were tellurium oxide (TeO2), lithium fluoride (LiF), yttrium fluoride (YF3), terbium fluoride (TbF3), and europium fluoride (EuF3). Calculated quantities of the chemicals were mixed in an agate mortar for 30 min and melted in a PDI electric furnace at 1100 °C for 1 h in alumina crucibles so that a homogeneously mixed melt was obtained. The reaction performance was 96%, indicating a low volatility of fluoride compounds. The samples were subsequently annealed at a temperature of 280 °C. The time to finish the annealing process took around 18 h. Samples were cut and then polished to 800 µm-thick slabs for different measurements. Photoluminescence characterizations were performed using a xenon lamp, monochromator 2300i from Acton Research (Trenton, NJ, USA) and R955 Hamamatsu photomultiplier tube (Hamamatsu, Japan). The decay profile (lifetime) corresponding to 545 nm and 613 nm was recorded using a pulsed UV LED (Opulent Americas, Raleigh, NC, USA) centered at 365 nm and a Teledyne LeCroy oscilloscope (HDO 5055, New York, NY, USA). Special care was taken to maintain the alignment of the set-up in order to compare the intensity of the visible signal between different characterized samples. The ceramic phosphor plate was placed on a commercial UV LED with an emission wavelength centered in 380 nm. An integrating sphere 1 m in diameter (Labsphere Co., North Sutton, NH, USA) was used to measure the CCT, CIE (Comission Internationale de l´Éclairage 1931) chromaticity diagram, color coordinates, and luminous efficacy (LE) with a bias current of 20 mA.
The crystalline structures of the samples were characterized using the X-ray diffraction (XRD) of Bruker instrument (D2 Phaser, Bruker, Billerica, MA, USA) equipment with Cu Kα radiation at 1.54184 Å. The recorded XRD diffractograms were obtained from 10 to 70° 2θ range with increments of 0.02° and a sweep time of 0.5 s. The SEM images were performed using a SEM of JEOL (JSM-7800F, Tokyo, Japan).

3. Results and Discussion

3.1. X-ray Diffraction (XRD) Characterization and Photoluminescence Properties

The composition and phase purity of Tb3+-doped and Tb3+/Eu3+-codoped samples with different rare earth concentrations were analyzed by XRD. As presented in Figure 1a, almost all diffraction peaks from the ceramic phosphor plate could be indexed to the standard tetragonal LiYF4 phase (PDF#77-0816) with low impurity phases, indicating that the samples were successfully crystallized, and TeO2 concentrations did not cause significant changes in the host structure. According to Figure 1a, it was confirmed that LiYF4 samples possessed a tetragonal crystal structure with space group I41/a. The obtained results were comparable to those reported by Kim et. al., where the tetragonal phase appeared as the main crystalline phase when the Eu3+ concentration was below 40 mol% in the LiYF4 matrix [28]. For this study, the total lanthanide concentrations in LiYF4 samples were 28, 28.15, 28.3, 29.5, 36.5, and 42 mol%. Additional impurity peaks appeared from 28 to 29.5 mol%, which were associated with a Y2Te6O15 (PDF#37-1393) phase, whereas from 36.5 to 42 mol% the peak centered at 2θ = 28.2° disappeared, and the purity of the tetragonal phase increased. The lanthanide concentration also modified the diffraction peaks, shifting to a high angle side, when the lanthanide increased from 28 mol% to 42 mol%. Such a fact was associated with the substitution of larger ionic radii of Tb3+ and Eu3+ by a smaller Y3+ ionic radius. An SEM image of 28 mol% of Tb3+, 28 mol% of Tb/0.15 mol% of Eu, and 14 mol% of Eu3+ are shown in Figure 1b. The increase in Eu3+ concentration did not promote a significant change on the surface of the phosphor plate.
Figure 2a shows the photoluminescence excitation (PLE) from 325 to 500 nm in Tb3+- and Eu3+-doped ceramic plates monitored at 544 and 614 nm, respectively. The spectrum shows four main excitation bands centered at 353, 373, 378, and 485 nm, which were assigned to 7F65D2, 7F65L10, 7F65D3, and 7F65D4 transitions of Tb3+ respectively [29]. The 7F65D3 transition showed two excitation peaks at 373 and 378 nm, with the 378 nm shoulder being slightly weaker. From 331 to 338 nm there was a continuous excitation, while from 388 to 475 nm no excitation peaks were observed in Tb3+-doped samples. Figure 2b shows the photoluminescence emission (PL) of Tb3+-doped and Eu3+-doped ceramic phosphor plates under 373 and 396 nm excitation wavelengths. The bands centered at 490, 545, 589, and 622 nm were assigned to 5D47F6, 5D47F5, 5D47F4, and 5D47F3 transitions of Tb3+ respectively. These visible bands were the result of the well-known down-conversion process, and their emitted colors depended on the concentration of Tb3+ ions. In this work, the Tb3+ concentration was varied systematically (not show here) from 2 to 28 mol%, and results showed that there was no quenching concentration indicium. These results were similar to those found by other research groups, where 40 mol% of dopant ions were incorporated, and the green emission (545 nm) was the dominant color for the LiYF4 matrix. The inset in Figure 2a shows the picture of the opaque ceramic phosphor plate.
To distinguish the PLE spectra of Tb3+ and Eu3+, a 14 mol% of Eu3+-doped sample was synthesized. Figure 2a shows the excitation spectra of the Eu3+-doped ceramic phosphor plate recorded at 614 nm emission. The sample showed five dominant bands centered at 362, 380, 393, 414, and 464 nm, which were assigned to 7F05D4, 7F05L7, 7F05L6, 7F05D3, and 7F05D2 transitions of Eu3+ respectively. Upon excitation at 393 nm, a magenta color appeared, with a spectral range from 585 to 710 nm, which are shown in Figure 2b. The 592, 614, 653, and 701 nm bands were attributed to 5D07F1, 5D07F2, 5D07F3, and 5D07F4 transitions of Eu3+, respectively, where the 592 and 614 nm emission bands were the feature emissions for LiYF4:Eu3+. Among these transitions, the electric dipole 5D07F2 transition was the most intense, followed by less intense magnetic dipole 5D07F1 transitions. This indicated that Eu3+ ions were located at noninversion symmetric sites [32]. An attractive detail of the emission spectrum was observed at the 701 nm band, which typically was 40% less intense than the 614 nm band in the LiYF4 matrix [28]. However, in this work the 701 nm band was debilitated, as it was 78% weaker than 614 nm band. For solid-state lighting applications, the 701 nm band is a waste of energy because the eye sensitivity is zero, whereas 614 nm bands are considered the optimal red emission to obtain high luminous efficacy and high CRI in warm white LEDs [33].
Figure 3 shows the excitation spectra of the Tb3+/Eu3+-codoped ceramic phosphor plate. By adding only 0.15 mol% of Eu3+ to the Tb3+-doped sample, the intensity ratios between splitting peaks at 373 and 378 nm changed slightly, where the shoulder at 378 nm was more intense. This fact was due to the spectral overlapping of the excitation bands at 373 and 380 nm of Tb3+ and Eu3+ ions, respectively. Such bands became more pronounced when the Eu3+ concentration increased from 0.05 to 5 mol%. From the point of view of solid-state lighting, the red shift of the excitation bands became significant because the optical power in the commercial LED chip at 380 nm was higher than 365 nm, and there was a lower probability of dispersion at larger wavelengths. In addition, upon 378 nm excitation, both Tb3+ and Eu3+ were excited efficiently (vertical gray line), producing simultaneously green and red emission bands. As a result, a higher luminous flux was obtained from a phosphor converter device. As presented in Figure 2b, the excitation bands at 378, 393, 464, and 485 nm for Tb3+/Eu3+-codoped samples increased with the Eu3+ concentration; however, the intensity of the band at 393 nm depended strongly on Eu3+ content rather than the other bands.
Figure 4 shows the emission spectra of Tb3+/Eu3+-codoped samples as a function of Eu3+ concentration. Simultaneous emissions from Tb3+ and Eu3+ were observed under 378 nm, which indicated the existence of energy transfer between Tb3+ and Eu3+. Intensity of emission bands of Tb3+ at 490, 545, and 589 nm decreased monotonically as the Eu3+ concentration increased from 0.15 mol% to 14 mol%. Among these bands, the 545 nm emission was the most influenced band by Eu3+ content, which decreased 36% of the initial intensity with only 0.15 mol% of Eu3+. Concurrently, an increment of 33% was observed for the 614 nm emission band of Eu3+ ions. Figure 4 shows that almost all energy was transferred from Tb3+ to Eu3+ when the Eu3+ reached 14 mol%, where red was the main color.

3.2. Rate Equation Model and Energy Transfer

To clarify the emissions corresponding to Tb3+ and Eu3+ ions, the following simplified model was proposed. (1) First, some Tb3+ ions were promoted from the ground state 7F6 to the excited 5D3 level as a result of pumping at 373 nm. The absorption rate was denoted by A02 (s−1). (2) Once some Terbium ions were in the 5D3 level, they relaxed nonradiatively to the 5D4 level; this multiphonon relaxation was denoted by A21 (s−1). Subsequently, two processes could occur, phonon relaxation A10 (s−1) or energy transfer from Tb3+ to Eu3+ ions (W). (3) Tb3+ emission wavelength peaks were 490, 545, 589, and 622 nm. (4) The energy transfer promotes some Eu3+ ions from the 7F0 ground state to 5D0 level where the emission rates B10 (s−1) of Eu3+ occurred in the 592, 614, 653, and 701 nm peaks, as is shown in Figure 5.
Therefore, the following ratio equations are proposed:
d N T b 2 d t = A 02 N T b 0 A 21
d N T b 1 d t = A 21 N T b 2 A 10 N T b 1 W N E u 0 N T b 1
d N E u 1 d t = B 10 N E u 1 + W N E u 0 N T b 1
where N T b 2 , N T b 1 , and N T b 0 (ions/cm3) are the Tb3+ ion populations in the 5D3, 5D4, and 7F6 energy levels, respectively. N E u 1 and N E u 0 (ions/cm3) are the Eu3+ ion populations in the 5D0 and 1F0 levels, respectively. For low-excitation pumping, the ground populations are proportional to the nominal concentration, i.e., N T b 0 N T b and N E u 0 N E u . In stationary conditions the solutions are:
N T b 1 = A 02 N T b ( W N E u + A 01 )
N E u 1 = W A 02 N T b N E u B 10 ( W N E u + A 01 )
N T b 2 = N T b A 02 A 21
Then, ion populations are related to the emission spectrum by:
N T b 1 = k I T b d λ
N E u 1 = k I E u d λ
where ITb and IEu are the emission spectra related to the Tb3+ and Eu3+ emission transitions, and k is a proportional constant. Thus, to obtain the ITb and IEu emission spectra, deconvolution of the spectra was performed for both ions, as is shown in Figure 6. Then, we define:
P T b = N T b 1 N T b 1 + N E u 1 = B 10 W N E u + B 10
P E u = B 21 N E u 1 N T b 1 + N E u 1 = W N E u W N E u + B 10
P T b = I T b d λ I T b d λ + I E u d λ
P E u = I E u d λ I T b d λ + I E u d λ
On the other hand, the dynamic solution for Tb3+ in the level 5D4 is:
N T b 1 ( t ) = A 21 A 02 N T b 0 A 21 A 10 N E u 0 W ( e ( A 01 + W N E u 0 ) t e A 21 t )
This equation has two terms related to the rise time and lifetime. In this way, the lifetime is expressed by:
τ = 1 A 01 + W N E u 0
Figure 7 shows the experimental lifetime curves, and the inset graph is the experimental fitting of Equation (14), with A01 = 226.8/s and W = 53.4/s mol%. In our samples, the lifetimes of the 5D4 level of Tb3+ showed values of 4.3, 4.28, 4.28, 3.44, 1.42, and 1.01 ms when Eu3+ concentration increased to 0, 0.15, 0.3, 1.5, 8.5, and 14 mol%, respectively (see Table 1).
Figure 8 shows the experimental normalized emission spectra according to Equations (11) and 12. Additionally, adjustment of the experimental data was carried out using Equations (9) and (10), with B10 = 131.2/s. The model simultaneously adjusted the emission spectrum and lifetime curves. Luminescent efficiency can be defined by the ratio of population loss by emission and the population gain of the level:
η = A 10 N T b 1 A 21 N T b 2
Substituting Equations (4) and (5) we find:
η = A 01 A 01 + W N E u
Energy transfer efficiency can be defined by the ratio of population loss by energy transfer and the population gain of the level:
E T = W N T b 1 N E u 0 A 21 N T b 2 = W N E u A 01 + W N E u
Equations (16) and (17) fulfill the following relationship:
E T + η = 1
Then, energy transfer can be rewritten as:
E T = 1 τ τ 0
where τ = ( A 01 + W N E u ) 1 is the lifetime, and τ 0 = ( A 01 ) 1 is the radiative lifetime. In this work, τ = τ T b E u and τ 0 = τ T b were the fluorescence lifetimes of the 5D4 level of Tb3+ for Tb3+-doped and Tb3+/Eu3+-codoped ceramic phosphor plates. The calculated ET efficiency increased from 5 to 76% when the Eu3+ concentration increased from 0.15 to 14 mol%. ET increased rapidly to 30% with the addition of 1.5 mol% of Eu3+; after this concentration, no big changes were observed, and the ET was kept almost constant at 14 mol% of Eu3+.

3.3. White Light Device Fabrication

Figure 9 shows multicolor light devices that were constructed using the Tb3+/Eu3+-codoped ceramic phosphor plate and the 380 nm UV LED chip.
The electroluminescence of fabricated devices as a function of Eu3+ concentration with a bias current of 20 mA is shown in Figure 10. Representative samples with 0.15, 0.3, and 1.5 mol% of Eu3+ clearly showed 380 (UV LED), 544 (Tb3+), 592 (Eu3+), and 614 nm (Eu3+) bands, where warm white was the feature color emissions of these devices. The maximum luminous efficacy was measured to be 13.08 lm/W for 28 mol% of Tb3+, whereas a decrement from 9.22 to 6.04 lm/W was observed as the Eu3+ concentration increased (see Table 1). The low values of luminous efficacy were associated with the opacity of the samples, the low efficiency of UV LED (InGaN), as well as the poor contribution of color from LED.
Figure 11 shows the color coordinates of the samples under study. The values were located on the edge of the chromaticity diagram, predominantly in yellow and red regions. The emitted color of the device changed from green to red through warm white by keeping the Tb3+ concentration and changing the Eu3+ content. The obtained results showed that the optical properties of LED were strongly influenced by Eu3+ content. Interestingly, by doping with only 0.15 mol% of Eu3+, it was possible to modify both CRI and CCT. CRI showed an increase from 34 to 74.81, and CCT diminished from 5497 to 3658 K; these features were very adequate for indoor lighting. The samples with 0.3 and 1.5 mol% of Eu3+ gave the highest CRIs of approximately 82.5 and 82.6 with CCTs of 3136 and 2225 K, respectively. Although the luminous efficacy of the devices was low, we expected to further increase such values by introducing Tb3+ and Eu3+ in an adequate matrix or by codoping with Ce3+ to increase the absorption strength. The obtained results showed that by choosing properly the Eu3+ concentration, it was possible to produce comfortable white light devices for vivid applications in daily life.

4. Conclusions

In summary, the fabrication of LiYF4 ceramic phosphor plates doped with Tb3+/Eu3+ ions is reported. The maximum excitation peak of a single Tb3+ was at 373 nm; however, this changed to 378 nm when LiYF4:10Tb3+ was codoped with Eu3+ ions. Based on the experimental results, it was concluded that intensity ratios between emission bands could be tuned by choosing properly the ion concentrations of both Tb3+ and Eu3+ ions. By placing the LiYF4:28Tb3+/yEu3+ (mol%) ceramic phosphor plates on the top of 380 nm LED chip, green, warm, and red color emissions were obtained. It was found that warm white was achieved by adding only 0.15 mol% of Eu3+ without serious detriment to the luminous efficacy. However, when the concentration of Eu3+ increased to 0.3 mol% in LiYF4:28Tb3+/yEu3+ (mol%), a CRI of 82.3 and a CCT of 3136 K were measured. With an increase in Eu3+ concentration, the yellow and red bands were improved, but the blue and green bands were reduced. Then, it was necessary to compromise CRI, CCT, and luminous efficacy to define the ion concentration. The obtained CRI was one of the highest reported in the literature for the Tb3+/Eu3+ system, which, in combination with low CCT, made LiYF4 ceramic phosphor plates a good candidate for solid-state lighting applications.

Author Contributions

Data curation, J.B.-P and P.C.; Investigation, H.D., P.C. and J.M-G.; Methodology, J.M.-G. and O.M.; Writing—original draft, H.D.; Writing—review & editing, H.D and J.M-G.

Funding

This work was supported by CONACyT, México through grant 288572.

Acknowledgments

Jorge Molina-González gratefully acknowledges to CONACyT for providing scholarship for PhD studies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fujita, S.; Yoshihara, S.; Sakamoto, A.; Yamamoto, S.; Tanabe, S. YAG Glass-Ceramic Phosphor for White LED (I): Background and Development; Ferguson, I.T., Carrano, J.C., Taguchi, T., Ashdown, I.E., Eds.; SPIE: San Diego, CA, USA, 2005; p. 594111. [Google Scholar]
  2. Zhang, X.; Wang, J.; Huang, L.; Pan, F.; Chen, Y.; Lei, B.; Peng, M.; Wu, M. Tunable Luminescent Properties and Concentration-Dependent, Site-Preferable Distribution of Eu2+ Ions in Silicate Glass for White LEDs Applications. ACS Appl. Mater. Interfaces 2015, 7, 10044–10054. [Google Scholar] [CrossRef] [PubMed]
  3. Fujita, S.; Sakamoto, A.; Tanabe, S. Luminescence Characteristics of YAG Glass–Ceramic Phosphor for White LED. IEEE J. Sel. Top. Quantum Electron. 2008, 14, 1387–1391. [Google Scholar] [CrossRef]
  4. Lee, Y.K.; Lee, J.S.; Heo, J.; Im, W.B.; Chung, W.J. Phosphor in glasses with Pb-free silicate glass powders as robust color-converting materials for white LED applications. Opt. Lett. 2012, 37, 3276. [Google Scholar] [CrossRef] [PubMed]
  5. Li, S.; Zhu, Q.; Wang, L.; Tang, D.; Cho, Y.; Liu, X.; Hirosaki, N.; Nishimura, T.; Sekiguchi, T.; Huang, Z.; et al. CaAlSiN3:Eu2+ translucent ceramic: A promising robust and efficient red color converter for solid state laser displays and lighting. J. Mater. Chem. C 2016, 4, 8197–8205. [Google Scholar] [CrossRef]
  6. Ji, E.-K.; Song, Y.-H.; Bak, S.H.; Jung, M.K.; Jeong, B.W.; Lee, D.B.; Yoon, D.-H. The design of a ceramic phosphor plate with functional materials for application in high power LEDs. J. Mater. Chem. C 2015, 3, 12390–12393. [Google Scholar] [CrossRef]
  7. Ghoneim, N.A.; Halawa, M.M. Effect of boron oxide on the thermal conductivity of some sodium silicate glasses. Thermochim. Acta 1985, 83, 341–345. [Google Scholar] [CrossRef]
  8. Hanna, B.; Bohn, R.G. Thermal Conductivity of Li2O·Al2O3·nSiO2 Glass—Ceramics between 5 and 100 K. J. Am. Ceram. Soc. 1991, 74, 3035–3038. [Google Scholar] [CrossRef]
  9. Zhang, R.; Lin, H.; Yu, Y.; Chen, D.; Xu, J.; Wang, Y. A new-generation color converter for high-power white LED: Transparent Ce3+:YAG phosphor-in-glass: Ce:YAG phosphor-in-glass for high-power white LED. Laser Photonics Rev. 2014, 8, 158–164. [Google Scholar] [CrossRef]
  10. Yoo, H.; Kouhara, Y.; Yoon, H.C.; Park, S.J.; Oh, J.H.; Do, Y.R. Sn–P–F containing glass matrix for the fabrication of phosphor-in-glass for use in high power LEDs. RSC Adv. 2016, 6, 111640–111647. [Google Scholar] [CrossRef]
  11. Kim, E.; Unithrattil, S.; Sohn, I.S.; Kim, S.J.; Chung, W.J.; Im, W.B. Facile one-step fabrication of 2-layered and 4-quadrant type phosphor-in-glass plates for white LEDs: An insight into angle dependent luminescence. Opt. Mater. Express 2016, 6, 804. [Google Scholar] [CrossRef]
  12. Feldmann, C.; Jüstel, T.; Ronda, C.R.; Schmidt, P.J. Inorganic Luminescent Materials: 100 Years of Research and Application. Adv. Funct. Mater. 2003, 13, 511–516. [Google Scholar] [CrossRef]
  13. 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]
  14. 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]
  15. Zhu, C.; Chaussedent, S.; Liu, S.; Zhang, Y.; Monteil, A.; Gaumer, N.; Yue, Y. Composition dependence of luminescence of Eu3+ and Eu3+/Tb3+ doped silicate glasses for LED applications. J. Alloys Compd. 2013, 555, 232–236. [Google Scholar] [CrossRef]
  16. Zhu, C.; Wang, J.; Zhang, M.; Ren, X.; Shen, J.; Yue, Y. Eu-,Tb-, and Dy-Doped Oxyfluoride Silicate Glasses for LED Applications. J. Am. Ceram. Soc. 2014, 97, 854–861. [Google Scholar] [CrossRef]
  17. Zhang, X.; Zhou, L.; Shi, J.; Gong, M. Luminescence and energy transfer of a color tunable phosphor Tb3+, Eu3+ co-doped KCaY(PO4)2. Mater. Lett. 2014, 137, 32–35. [Google Scholar] [CrossRef]
  18. Wang, B.; Ren, Q.; Hai, O.; Wu, X. Luminescence properties and energy transfer in Tb3+ and Eu3+ co-doped Ba2P2O7 phosphors. RSC Adv. 2017, 7, 15222–15227. [Google Scholar] [CrossRef]
  19. Xu, M.; Wang, L.; Jia, D.; Zhao, H. Tuning the Color Emission of Sr2P2O7: Tb3+, Eu3+ Phosphors Based on Energy Transfer. J. Am. Ceram. Soc. 2015, 98, 1536–1541. [Google Scholar] [CrossRef]
  20. Jiang, T.; Yu, X.; Xu, X.; Yu, H.; Zhou, D.; Qiu, J. Realization of tunable emission via efficient Tb3+–Eu3+ energy transfer in K3Gd(PO4)2 for UV-excited white light-emitting-diodes. Opt. Mater. 2014, 36, 611–615. [Google Scholar] [CrossRef]
  21. Zhang, X.; He, P.; Zhou, L.; Shi, J.; Gong, M. Energy transfer and luminescent properties of a green-to-red color tunable Tb3+, Eu3+ co-doped K2Y(WO4)(PO4) phosphor. Mater. Res. Bull. 2014, 60, 300–307. [Google Scholar] [CrossRef]
  22. Xia, Y.; Huang, Y.; Long, Q.; Liao, S.; Gao, Y.; Liang, J.; Cai, J. Near-UV light excited Eu3+, Tb3+, Bi3+ co-doped LaPO4 phosphors: Synthesis and enhancement of red emission for WLEDs. Ceram. Int. 2015, 41, 5525–5530. [Google Scholar] [CrossRef]
  23. Zhang, B.; Zou, H.; Guan, H.; Dai, Y.; Song, Y.; Zhou, X.; Sheng, Y. Lu2O2S:Tb3+, Eu3+ nanorods: Luminescence, energy transfer, and multicolour tuneable emission. CrystEngComm 2016, 18, 7620–7628. [Google Scholar] [CrossRef]
  24. Huo, J.; Dong, L.; Lü, W.; Shao, B.; You, H. Novel tunable green-red-emitting oxynitride phosphors co-activated with Ce3+, Tb3+, and Eu3+: Photoluminescence and energy transfer. Phys. Chem. Chem. Phys. 2017, 19, 17314–17323. [Google Scholar] [CrossRef] [PubMed]
  25. Xia, Z.; Zhuang, J.; Liu, H.; Liao, L. Photoluminescence properties and energy transfer of Ba2Lu(BO3)2Cl:Eu2+/Eu3+, Tb3+ phosphors. J. Phys. Appl. Phys. 2012, 45, 015302. [Google Scholar] [CrossRef]
  26. Jiang, Y.; Xia, H.; Zhang, J.; Yang, S.; Jiang, H.; Chen, B. Growth and Optical Spectra of Tb3+/Eu3+ Co-doped Cubic NaYF4 Single Crystal for White Light Emitting Diode. J. Mater. Sci. Technol. 2015, 31, 1232–1236. [Google Scholar] [CrossRef]
  27. Mahalingam, V.; Vetrone, F.; Naccache, R.; Speghini, A.; Capobianco, J.A. Colloidal Tm3+/Yb3+-Doped LiYF4 Nanocrystals: Multiple Luminescence Spanning the UV to NIR Regions via Low-Energy Excitation. Adv. Mater. 2009, 21, 4025–4028. [Google Scholar] [CrossRef]
  28. Kim, S.Y.; Won, Y.-H.; Jang, H.S. A Strategy to enhance Eu3+ emission from LiYF4:Eu nanophosphors and green-to-orange multicolor tunable, transparent nanophosphor-polymer composites. Sci. Rep. 2015, 5, 7866. [Google Scholar] [CrossRef]
  29. Szpikowska-Sroka, B.; Pawlik, N.; Goryczka, T.; Pisarski, W.A. Technological aspects for Tb3+-doped luminescent sol-gel nanomaterials. Ceram. Int. 2015, 41, 11670–11679. [Google Scholar] [CrossRef]
  30. Jiang, Y.; Xia, H.; Yang, S.; Zhang, J.; Jiang, D.; Wang, C.; Feng, Z.; Zhang, J.; Gu, X.; Zhang, J.; et al. Luminescence of Tb3+/Eu3+ codoped LiYF4 single crystals under UV excitation for white-light LEDs. Chin. Opt. Lett. 2015, 13, 071601. [Google Scholar] [CrossRef]
  31. Zhang, S.; Li, Y.; Lv, Y.; Fan, L.; Hu, Y.; He, M. A full-color emitting phosphor Ca9Ce(PO4) 7:Mn2+, Tb3+: Efficient energy transfer, stable thermal stability and high quantum efficiency. Chem. Eng. J. 2017, 322, 314–327. [Google Scholar] [CrossRef]
  32. Ptacek, P.; Schäfer, H.; Kömpe, K.; Haase, M. Crystal Phase Control of Luminescing α-NaGdF4:Eu3+and β-NaGdF4:Eu3+ Nanocrystals. Adv. Funct. Mater. 2007, 17, 3843–3848. [Google Scholar] [CrossRef]
  33. Žukauskas, A.; Vaicekauskas, R.; Ivanauskas, F.; Vaitkevičius, 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]
Figure 1. (a) X-ray diffraction pattern of the phosphor ceramic plate. The peaks marked with an asterisk correspond to the phase Y2Te6O15. (b) SEM image of the ceramic phosphor plate synthetized trough melt quenching method.
Figure 1. (a) X-ray diffraction pattern of the phosphor ceramic plate. The peaks marked with an asterisk correspond to the phase Y2Te6O15. (b) SEM image of the ceramic phosphor plate synthetized trough melt quenching method.
Materials 12 02240 g001
Figure 2. (a) Excitation and (b) emission spectra for Eu3+- and Tb3+-doped samples. The inset shows the picture of the opaque ceramic phosphor plate.
Figure 2. (a) Excitation and (b) emission spectra for Eu3+- and Tb3+-doped samples. The inset shows the picture of the opaque ceramic phosphor plate.
Materials 12 02240 g002
Figure 3. Excitation spectra of the Tb3+/Eu3+-codoped ceramic phosphor plate monitored at 614 nm.
Figure 3. Excitation spectra of the Tb3+/Eu3+-codoped ceramic phosphor plate monitored at 614 nm.
Materials 12 02240 g003
Figure 4. Emission spectra of the Tb3+/Eu3+-codoped ceramic phosphor plate excited at 378 nm.
Figure 4. Emission spectra of the Tb3+/Eu3+-codoped ceramic phosphor plate excited at 378 nm.
Materials 12 02240 g004
Figure 5. Schematic level diagram of Tb3+ and Eu3+ and the energy transfer mechanism.
Figure 5. Schematic level diagram of Tb3+ and Eu3+ and the energy transfer mechanism.
Materials 12 02240 g005
Figure 6. Deconvolution to the experimental emission spectrum (blue). Red and green areas represent the emissions related to the Eu3+ and Tb3+ ions, respectively.
Figure 6. Deconvolution to the experimental emission spectrum (blue). Red and green areas represent the emissions related to the Eu3+ and Tb3+ ions, respectively.
Materials 12 02240 g006
Figure 7. Experimental lifetime curves. The inset is a graph of experimental lifetime. Black lines represent the fitting model.
Figure 7. Experimental lifetime curves. The inset is a graph of experimental lifetime. Black lines represent the fitting model.
Materials 12 02240 g007
Figure 8. Normalized emission spectrum related to the Eu3+ (red) and Tb3+ (green) ions. Black lines represent the fitting model.
Figure 8. Normalized emission spectrum related to the Eu3+ (red) and Tb3+ (green) ions. Black lines represent the fitting model.
Materials 12 02240 g008
Figure 9. Luminescence photographs of fabricated devices under a 380 nm UV LED chip.
Figure 9. Luminescence photographs of fabricated devices under a 380 nm UV LED chip.
Materials 12 02240 g009
Figure 10. Electroluminescent spectra of LEDs fabricated using a 380 nm UV LED chip combined with Tb3+-doped and Tb3+/Eu3+-codoped ceramic phosphor plates under a forward bias of 20 mA.
Figure 10. Electroluminescent spectra of LEDs fabricated using a 380 nm UV LED chip combined with Tb3+-doped and Tb3+/Eu3+-codoped ceramic phosphor plates under a forward bias of 20 mA.
Materials 12 02240 g010
Figure 11. CIE diagram of chromaticity as function of Eu3+ content.
Figure 11. CIE diagram of chromaticity as function of Eu3+ content.
Materials 12 02240 g011
Table 1. Decay time, energy transfer, color coordinate (CC), color-correlated temperature (CCT), color rendering index (CRI), and luminous efficacy (LE) of the ceramic phosphor plate as a function of Eu3+ content.
Table 1. Decay time, energy transfer, color coordinate (CC), color-correlated temperature (CCT), color rendering index (CRI), and luminous efficacy (LE) of the ceramic phosphor plate as a function of Eu3+ content.
Tb3+/Eu3+ (mol%)τTb(5D4) (ms)ET (%)xyCCT (K)CRILE (lm/W)
284.300.3690.602549734.0613.08
28/0.154.280.40.4590.521365874.819.22
28/0.34.280.40.4820.500313682.548.13
28/1.53.44200.5340.453222582.597.09
28/8.51.42660.6140.381113642.817.80
28/140.93780.6290.368129435.546.04

Share and Cite

MDPI and ACS Style

Desirena, H.; Molina-González, J.; Meza, O.; Castillo, P.; Bujdud-Pérez, J. Multicolor and Warm White Emissions with a High Color Rendering Index in a Tb3+/Eu3+-Codoped Phosphor Ceramic Plate. Materials 2019, 12, 2240. https://doi.org/10.3390/ma12142240

AMA Style

Desirena H, Molina-González J, Meza O, Castillo P, Bujdud-Pérez J. Multicolor and Warm White Emissions with a High Color Rendering Index in a Tb3+/Eu3+-Codoped Phosphor Ceramic Plate. Materials. 2019; 12(14):2240. https://doi.org/10.3390/ma12142240

Chicago/Turabian Style

Desirena, Haggeo, Jorge Molina-González, Octavio Meza, Priscilla Castillo, and Juan Bujdud-Pérez. 2019. "Multicolor and Warm White Emissions with a High Color Rendering Index in a Tb3+/Eu3+-Codoped Phosphor Ceramic Plate" Materials 12, no. 14: 2240. https://doi.org/10.3390/ma12142240

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