Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Origin of saturated green emission from europium in zinc thiogallate

Open Access Open Access

Abstract

Europium doped zinc thiogallate, ZnGa2S4:Eu2+, has been reported as a saturated green emitting phosphor, suitable as conversion phosphor in white LEDs for lighting or displays. Up to now, no direct proof for the incorporation of Eu2+ in ZnGa2S4 has been given. We combined X-ray diffraction (XRD), cathodoluminescence in electron microscopy (SEM-CL) and X-ray absorption spectroscopy (XAS) to study the incorporation of the europium ions in the host material. The previously reported green luminescence was found to originate from small amounts of unintentionally formed EuGa2S4, and not from europium ions incorporated into ZnGa2S4. EuGa2S4 has a low quantum efficiency (< 20%) and shows strong thermal quenching, already below room temperature. The XAS data analysis suggests that a certain amount of europium might occupy octahedral voids inside the zinc thiogallate lattice in a divalent state. The zinc ion next to these interstitial dopants is then removed for charge compensation. Notwithstanding the possible, but limited, incorporation of Eu2+ in ZnGa2S4, these ions do not activate any luminescence as was shown with SEM-CL.

©2013 Optical Society of America

1. Introduction

Conversion phosphors emitting a saturated green color are very interesting for white light-emitting diode (LED) applications but are at the same time surprisingly scarce. Both in lighting and imaging, a white phosphor converted LED (pc-LED), composed of a blue pumping (In,Ga)N LED and one (a yellow, for example Y3Al5O12:Ce3+ [1]) or two (a green and red, for example SrSi2O2N2:Eu2+ and Sr2Si5N8:Eu2+ [2] or CaAlSiN3:Eu2+ [3]) conversion phosphors are currently used devices. When a violet or a near-ultraviolet pumping LED is used, a blue emitting conversion phosphor is needed in addition.

For lighting, broad emission bands are advantageous for reaching high color rendering, while narrow emission bands are needed in displays. Indeed, displays typically use RGB primary colors which should span a wide color gamut. The RGB primaries should therefore be saturated colors, requiring relatively narrow emission bands. When broadband phosphors are used, obtaining saturated primaries necessitates filtering a significant amount of the emitted light. This considerably lowers the overall device efficiency. Using saturated phosphors for which no color filtering is needed to obtain the correct primary does not only increases the device efficiency, but also increases the attainable color gamut [4].

Saturated green phosphors are also of interest for the construction of green LEDs. The external quantum efficiency of (In,Ga)N diodes decreases considerably with increasing indium content, as required to shift the emission wavelength to the green spectral region [5]. By combining a saturated green phosphor with a violet or blue pumping LED, a more energy efficient green LED is obtained, provided a high conversion efficiency phosphor is available [6]. In a recent publication, Oh et al. described a monochromatic green pc-LED, using a dichroic band-pass filter (BPF) and a green orthosilicate conversion phosphor [7]. Since the color purity of this phosphor was rather poor, a short-pass dichroic filter (SPDF) was added to absorb a significant fraction of the generated light. This issue can be resolved with an alternative green phosphor with a superior color purity.

Recently, six requirements were outlined which a conversion phosphor should fulfill in order to be a suitable candidate for the above-mentioned applications [8]. Besides the appropriate emission color, the phosphor should be optimally excitable with the emission from the pumping LED. Furthermore, the conversion process must occur with high quantum efficiency (QE) in order to maximize the overall efficiency of the final device. Moreover, the thermal quenching should be sufficiently low to prevent a color drift of the LED when the phosphor heats during operation.

In Table 1, an overview of green phosphors in recent literature is presented. Phosphors which cannot be efficiently excited with blue or violet light, such as Tb3+ doped phosphors, were omitted. From this table, it is clear that the number of saturated green phosphors (with 50 nm or less full width at half maximum (FWHM) of the emission band) is very limited. Next to SrGa2S4:Eu2+ and β-SiAlON:Eu2+, no phosphor is known that exhibits a sufficiently narrow emission band, combined with a high QE.

Tables Icon

Table 1. Overview of luminescent properties of green phosphors from recent literature. T0.5 describes the thermal quenching. It is the temperature at which the emission intensity is 50% of the value at low temperature. QE (int/ext) is the internal/external quantum efficiency.

Europium doped zinc thiogallate, ZnGa2S4:Eu2+ is, like every Eu2+ activated phosphor, expected to be a broadband emitter due to the 4f65d → 4f7 transition that is responsible for the luminescence. The emission band of this material has been reported to be rather narrow, yielding a saturated green color [2125]. In this paper, we investigate whether ZnGa2S4:Eu2+ has the same advantageous properties as the chemically similar SrGa2S4:Eu2+ (Table 1). The particular motivation for this work is the contradiction between a number of reports which have been recently published on ZnGa2S4:Eu2+ phosphors.

The photoluminescence of ZnGa2S4:Eu2+ was first described by Yuta and White [21], reporting broadband emission at 535 nm and a Stokes shift of 0.78 eV. Kim and Kim elaborated on the photoluminescence (PL) and cathodoluminescence (CL) of ZnGa2S4:Eu2+ and (Ca,Zn)Ga2S4:Eu2+ mixtures [22, 26]. The synthesis parameters of the solid-state reaction yielding the powders were optimized and a saturated green emission at 540 nm was reported. The lack of efficient CL was ascribed to the small particle size of the phosphor powders. By adding only a small amount of calcium to the mixture, the emission band at 540 nm disappeared abruptly and a new emission band was formed at longer wavelengths (> 550 nm). Recently, Yu et al. reported the thermal properties of ZnGa2S4:Eu2+. Herein, it was found that this material has a relatively low thermal quenching temperature, but it remains a potential candidate for LED applications because of the favorable excitation spectrum [24]. Reported key parameters about ZnGa2S4:Eu2+ (at room temperature) are an emission band at 540 nm (FWHM of 50 nm), a luminescence lifetime ranging from 126 to 79 ns, depending on the europium concentration and a quenching temperature T0.5 of 407 K. T0.5 was defined as the temperature where the emission intensity dropped to half the intensity value at 300 K, which is somewhat in contrast to the common approach to comparing the intensity to the value at low temperature, where no thermal quenching is noticeable. In [25], luminescence in ZnS-ZnGa2S4:Eu2+ mixed compounds was explained by energy transfer from ZnS:Eu2+ to ZnGa2S4:Eu2+. In contrast to the above reports, an anomalously broad emission at 565nm has been reported in [27]. Up to now, no value of the quantum efficiency of ZnGa2S4:Eu2+ has been reported.

Additionally, there is still discussion about the position of the Eu2+ ions in the zinc thiogallate lattice. Due to the size mismatch between the Eu2+ and Zn2+ ions, one can expect a difficult incorporation of the europium ions in the zinc-based host material. The ions could (1) substitute for the zinc ions, (2) occupy the tetrahedral vacancy sites in the lattice, or (3) substitute at octahedral voids of the host lattice. Yuta and White proposed an energy scheme for Eu2+ in ZnGa2S4, based on tetrahedral coordination [21]. The difficult incorporation of Eu2+ on small tetrahedral coordinated sites was already drawn to attention by Wickleder et al. [28]. They argued that the octahedral voids are favored, based on the 4f65d → 4f7 transition energy and the observation that no europium doped compounds exist where the doping ion is fourfold coordinated as would be the case on the zinc or vacancy sites. However, no direct proof for any of the europium positions has been given yet [28, 29].

The hypothesis which is explored in the present paper is that the observed luminescence originates from small amounts of unintentionally formed EuGa2S4. This stoichiometric phosphor is known to exhibit efficient green luminescence, centered around 545 nm [12]. This is surprisingly similar to the reported values for the ZnGa2S4:Eu2+ phosphors. In EuGa2S4, the europium ions are eightfold coordinated by sulfur ions.

In this paper, the internal and external quantum efficiencies and the other vital properties of the luminescence of ZnGa2S4:Eu2+ are characterized and discussed, to complement the previously published data on this phosphor. Next, the suitability of ZnGa2S4:Eu2+ as a conversion phosphor for lighting and display applications is evaluated. Emphasis is put on the microscopic structure of the phosphor, both at the level of individual phosphor particles and at the atomic level, to get a thorough understanding of the doping of Eu2+ into ZnGa2S4.

2. Experimental

Zn1-xEuxGa2S4 powders with europium concentrations of x = 0.01, 0.02, 0.03, 0.04, 0.06, 0.08 and 0.10 were synthesized by a solid state reaction between ZnS (Alfa Products, 99.9%), Ga2S3 (Alfa Aesar, 99.99%) and EuF3 (Alfa Aesar, 99.5%). Stoichiometric quantities of these starting materials were mixed and heat treated at 1000 °C for 2 hours in alumina crucibles under a H2S flux. After the heat treatment, the powders cooled naturally and were lightly ground. Pure EuGa2S4 was synthesized in a similar fashion, however using EuS instead of EuF3 and without ZnS.

Photoluminescence (PL) emission and excitation spectra were measured with an Edinburgh FS920 fluorescence spectrometer. Decay profiles were collected with an intensified CCD (Andor DH720), combined with a pulsed LED (peak wavelength of 400 nm, pulse frequency of 10 kHz) as excitation source. Measurements as a function of temperature were performed using an Oxford Optistat CF cryostat.

Diffuse reflectance spectra were measured with a Varian Cary 500 spectrophotometer, equipped with an internal integrating sphere. External quantum efficiencies were obtained inside the Edinburgh fluorescence spectrometer, using a phosphor with known QE as a standard (SrSi2O2N2:Eu2+ with an internal QE of 90% [15]). Subsequently, the absorption of the phosphors was measured inside an integrating sphere. From this data, the internal quantum efficiency of the samples was calculated.

SEM-EDX-CL measurements were performed on a Hitachi S-3400N scanning electron microscope (SEM), equipped with an energy dispersive X-ray (EDX) detector (Thermo Scientific Noran 7). The cathodoluminescence (CL) light was captured in an optical fiber and analyzed by a CCD (Princeton Instruments ProEM 16002), attached to a spectrograph (Princeton Instruments Acton SP2358).

The crystal structure of the powders was obtained with X-ray diffraction (XRD), using Cu radiation on a Siemens D5000 diffractometer (40 kV, 40 mA).

EXAFS (extended x-ray absorption fine structure) measurements were performed at the Eu LIII edge and at the Zn K edge at the Dubble beamline (BM26A) at the ESRF synchrotron facility in Grenoble (France). A silicon (111) double crystal monochromator was used and all EXAFS data were recorded in fluorescence mode [30]. The EXAFS data reduction and analysis were performed by fitting the data to theoretical standards generated from the FEFF6 code 16 using the Athena and Artemis software by Ravel and Newville [31].

3. Results and discussion

3.1 Photoluminescence and diffuse reflectance spectra

The photoluminescence emission and excitation spectra of the powders with dopant concentrations up to x = 0.10 were measured at room temperature. Both the position of the maximum (at about 542 nm) and the full width at half maximum (FWHM, about 50 nm) of the emission peak show no significant dependence on the europium concentration. The spectra of Zn0.99Eu0.01Ga2S4, which are also representative for all other doping concentrations, are shown in Fig. 1. We refer to §3.2 for a comparison of the absolute intensity of the luminescence. The emission band of the phosphor corresponds with a saturated green color (CIE x = 0.31, y = 0.66) which exceeds the green color point of the EBU gamut.

 figure: Fig. 1

Fig. 1 Top: Diffuse reflectance spectra for ZnGa2S4 (x = 0.00), Zn0.99Eu0.01Ga2S4 (x = 0.01) and EuGa2S4 (x = 1.00) at room temperature. Bottom: PL emission (at 450 nm excitation, solid lines) and excitation (at 540 nm emission, dashed lines) spectra for Zn0.99Eu0.01Ga2S4 (x = 0.01) and EuGa2S4 (x = 1.00) at room temperature, normalized for better readability. The dip in the excitation spectra around 470 nm is an artifact.

Download Full Size | PDF

Both the emission and excitation spectra of the powders are very similar to the spectra of the stoichiometric phosphor EuGa2S4 which are also included in the figure. This is surprising given the very different structure of the ZnGa2S4 and EuGa2S4 host crystals. The difference in the excitation spectra below 420 nm can be explained in terms of interband absorption in ZnGa2S4. The absorbed energy is not transferred to Eu2+ ions and is dissipated non-radiatively.

Diffuse reflectance spectra are also included in Fig. 1, for Zn0.99Eu0.01Ga2S4, EuGa2S4 and undoped ZnGa2S4. The steep increase in the ZnGa2S4 absorption around 400 nm is due to interband absorption of the host material. The bandgap of ZnGa2S4, 3.4 eV, was obtained from a fit to the Kubelka-Munk spectrum, calculated from the reflection spectrum. This is in correspondence with the reported value of 3.22 eV [28]. EuGa2S4 starts to absorb light from ± 520 nm due to electronic excitation towards the 5d levels of Eu2+.

3.2 Quantum efficiency

The internal quantum efficiency considers only the conversion process after a photon is absorbed, the external quantum efficiency also takes the absorption into account:

ηext=NemNincηint=NemNabsηext=Aηint
Herein, Nem, Nabs and Ninc are the number of photons emitted and absorbed by the phosphor and incident on the phosphor respectively. A is the fraction of the incident light being absorbed. The internal and external quantum efficiencies of the Zn1-xEuxGa2S4 powders are depicted in Fig. 2. The internal quantum efficiency remains essentially constant as a function of Eu-concentration at a value of about 18%. This is striking since concentration quenching is expected to occur because of the increasing probability for energy transfer between two adjacent Eu2+ ions when they are closer together. Still, some small changes as a function of doping concentration can be observed. If one assumes that small grains of luminescent EuGa2S4 are mixed in ZnGa2S4 powder (this assumption will be substantiated in the remainder of this paper), these changes can be attributed to scattering effects, the nonzero absorption of ZnGa2S4 (at low x-values, see Fig. 1) and reabsorption of converted light by other EuGa2S4 grains (at higher x-values). In any case, the low conversion efficiency of this phosphor hampers the possibilities for technological applications.

 figure: Fig. 2

Fig. 2 Absorption (fraction of the incident photons that are absorbed), internal and external quantum efficiency for Zn1-xEuxGa2S4 at room temperature, measured at 460 nm excitation.

Download Full Size | PDF

3.3 Decay and thermal behavior

The decay profiles of the samples were measured at different temperatures with the pulsed excitation from an LED emitting at 400 nm. A luminescent lifetime was determined by fitting the initial decay with a single exponential function. The lifetime of 127 ± 24 ns did not change as a function of the europium concentration. This is a remarkably short luminescent lifetime for 4f65d 4f7 transitions in Eu2+. In comparison, the luminescent lifetime of other europium doped thiogallate and thioaluminate phosphors are given in Table 2. Emission maxima are also displayed because of the intrinsic dependence of the lifetime on the emission color, shorter wavelengths corresponding with shorter lifetimes (if the refractive indices are assumed to be similar) [32]. At low temperature (77 K), the lifetime increased to ± 350 ns. The other, more efficient thioaluminate and thiogallate phosphors do not show a change in lifetime as a function of temperature unless thermal quenching sets in [33].

Tables Icon

Table 2. Emission maxima (λmax) and full width at half maximum (FWHM), luminescent lifetimes (τ) and quenching temperatures (T0.5) of multiple Eu2+ doped thiogallate and thioaluminate phosphors. The displayed parameters were recorded at room temperature.

The integrated emission intensity of the Zn1-xEuxGa2S4 powders was monitored as a function of temperature (Fig. 3). Common LED phosphors show a more or less flat profile of the emission intensity as a function of temperature until a sudden steep drop at a well-defined activation temperature [38]. This activation temperature is associated with an energy barrier that, when crossed, leads to the non-radiative depopulation of the excited state. It is notable that the emission intensity does not show a stable region as a function of temperature, but decreases over the entire temperature range. This is a common behavior for stoichiometric phosphors [13, 39]. The quenching temperature, defined as the temperature where the emission intensity is half of the intensity at low temperature (here, 100 K), is about 240 K for x = 0.06. The obtained result is in accordance to the quenching temperature of 407 K as reported in [24], taking into account that the reference level was then taken at 300 K. No significant influence of the doping concentration on the decay or thermal quenching behavior was observed.

 figure: Fig. 3

Fig. 3 Normalized PL intensity and luminescent lifetimes as function of temperature for Zn0.94Eu0.06Ga2S4 upon excitation at 400 nm.

Download Full Size | PDF

From these measurements, we can state that the luminescence of these phosphor powders is essentially quenched over the whole studied temperature range, and is characterized by an unusual short luminescent lifetime.

The temperature dependence of the corresponding time constant is very similar to that of the non-radiative decay path, as identified in SrGa2S4:Eu2+ [11]. For the latter phosphor, this decay path is only relevant at high doping concentrations or temperatures above 400 K. It was found that in this phosphor, which in contrast to ZnGa2S4 is isostructural with EuGa2S4, the environment of Eu locally resembles that of EuGa2S4 due to clustering of europium [11].

The luminescent lifetime for EuGa2S4 was evaluated to be 116 ns (at 300 K). Within the error range, this is the same as the decay time of the ZnGa2S4:Eu2+ powders. The thermal quenching profile of EuGa2S4 was measured by Iida et al. [12]. Their result is very similar to the profiles we obtained for ZnGa2S4:Eu2+.

This luminescence characterization of Zn1-xEuxGa2S4 powders indicates that the light emission most likely originates from EuGa2S4 impurities. The PL emission and excitation spectra, quantum efficiency, thermal behavior and decay dynamics can be explained with this conjecture. In the next part, structural analysis will be applied to confirm this finding.

3.4 Luminescence on a microscopic scale

By the combined mapping of cathodoluminescence and characteristic X-rays in a scanning electron microscope setup, local variations in the chemical composition of a powder sample can be related to differences in luminescent properties [40]. In Fig. 4, a SEM-CL-EDX mapping of the Zn0.99Eu0.01Ga2S4 powder is displayed. In the EDX map, the areas which are colored red (green) are the result from the mapping on Eu (Zn). One can clearly discern EuGa2S4 grains, where no Zn is detected, among the majority of ZnGa2S4 grains, where no Eu is detected. The CL map shows that the characteristic green luminescence is indeed originating from the EuGa2S4 grains, due to the perfect correlation between the CL and the EDX maps. No light output is detected from the ZnGa2S4 phase. This shows that the presence of EuGa2S4 in the synthesized phosphors cannot be ignored from a luminescence point of view.

 figure: Fig. 4

Fig. 4 SEM-CL-EDX mapping of Zn0.99Eu0.01Ga2S4 powder, obtained at 250 K. Top left: phosphor morphology obtained by backscattered electron imaging. The integrated CL emission intensity is shown as an overlay in false colors. Top right: elemental distribution by EDX, where the colors are determined by color coding with green for Zn and red for Eu. Simultaneous detection would lead to a yellow color. Bottom: CL emission spectra for the points indicated in the top right figure, obtained in clockwise direction.

Download Full Size | PDF

There are two possible reasons for the absence of luminescence in the ZnGa2S4 grains. First, it is possible that no europium is incorporated in the ZnGa2S4 lattice at all. This is plausible due to the size mismatch between the Zn2+ and Eu2+ ions (88 pm versus 131 pm for sixfold coordination [41]).

Secondly, incorporation of europium ions in the host lattice does not necessarily cause light emission [42]. It would not be surprising if Eu in ZnGa2S4:Eu2+ is not luminescent. When ZnGa2S4 doped with europium would–coincidentally–have a very similar emission spectrum as EuGa2S4, then the absorption energy is estimated to be 2.4 eV, for the transition between the 4f 7 ground state and the lowest 4f 65d excited state of europium. This absorption energy is the value where the PL excitation spectrum amounts to 20% of its maximum value [43]. Given the small bandgap of this host material (3.2-3.4 eV), the excited states will probably overlap with – or be in close distance of – the conduction band of the ZnGa2S4 host, totally quenching any europium activated luminescence [42, 43].

One could also argue that the absence of CL does not automatically imply the absence of PL. Although there can be a noticeable difference between the efficiency of CL and PL, a complete absence of cathodoluminescence for a lanthanide doped inorganic crystal is not expected [44].

3.5 X-ray diffraction

The XRD patterns show that crystalline powders are obtained, corresponding with the defect-stannite structure of ZnGa2S4 (Fig. 5) [45, 46]. However, only a limited amount of europium can apparently be dissolved in the ZnGa2S4 lattice without clustering, as diffraction peaks originating from a EuGa2S4, having a different crystal structure, are observed. Careful inspection of earlier reports on XRD patterns of ZnGa2S4:Eu2+, also suggests the presence of EuGa2S4 as an impurity phase [22, 25, 47].

 figure: Fig. 5

Fig. 5 X-ray diffraction patterns for Zn1-xEuxGa2S4 powders, compared with reference patterns for ZnGa2S4 (ICSD 69542) and EuGa2S4 (ICSD 8053) [45, 47].

Download Full Size | PDF

The emerging of the EuGa2S4 (4 4 2) peak at 24.1° is studied, together with the ZnGa2S4 (1 1 0) peak at 23.8° to estimate the dependence of the ZnGa2S4 – EuGa2S4 fractional content on the doping concentration. The separate peak intensities, I, were calculated by a fitting procedure in the 2θ range from 23° to 25° with three pseudo-voigt peak shapes. The third peak is the less intense EuGa2S4 (4 4 0) peak at 24.4°. The pseudo-voigt profile is calculated as a linear combination of a Gaussian and a Lorentzian peak shape [48].

In Fig. 6, the values of

y(x)=IEuGa2S4IZnGa2S41xx
are displayed. Herein, I are the integrated areas of the XRD peaks for the EuGa2S4 (4 4 2) and ZnGa2S4 (1 1 0) peaks. One can expect a constant y as function of x when no europium is incorporated in ZnGa2S4 or zinc in EuGa2S4. However, when a reasonable amount of Eu is incorporated in ZnGa2S4, one expects a disproportionately low amount of EuGa2S4 in the powders with small x-values, resulting in a smaller y-value. Since there does not seem to be a variation of y for different europium concentrations, no incorporation of europium in ZnGa2S4 is discerned within the accuracy of the XRD measurements and fitting procedure. Therefore, if any europium is incorporated, it is restricted to a small fraction, unlike the case of conventional phosphor materials. This is underpinned by the constant XRD peak locations upon changing the doping concentration. To measure the incorporation with a better accuracy, different experimental techniques have to be addressed.

 figure: Fig. 6

Fig. 6 y(x), as defined in Eq. (2).

Download Full Size | PDF

3.6 Incorporation of Eu2+ in ZnGa2S4

X-ray absorption spectroscopy is ideally suited to study both the valence state and dopant incorporation in polycrystalline inorganic materials. The valence state can be probed by means of XANES (x-ray absorption near edge spectroscopy) [49], while EXAFS gives details on the first coordination spheres of the element under investigation [50].

Figure 7 shows the Eu LIII edge XANES spectrum of Zn0.96Eu0.04Ga2S4 compared with those of Eu2O3 and EuS, measured as reference compounds for Eu3+ and Eu2+, respectively. The shape of the spectrum of Zn0.96Eu0.04Ga2S4 is almost the same as that for EuS having a sharp single peak at 6972 eV and different from that for Eu2O3 having a single peak 8 eV higher in energy. This result shows that most of Eu atoms in Zn0.96Eu0.04Ga2S4 are divalent. This implies that the europium, which is doped in a trivalent state through EuF3 is effectively reduced during the heat treatment in the H2S atmosphere.

 figure: Fig. 7

Fig. 7 Eu LIII XANES spectrum of (top) Zn0.96Eu0.04Ga2S4, compared with the spectra of reference compounds for Eu2+ and Eu3+ (bottom), EuS (solid line) and Eu2O3 (dashed line).

Download Full Size | PDF

The extracted EXAFS spectra, k2χ(k) of the Eu LIII edge and the Zn K edge in ZnGa2S4:Eu2+ (4%) are shown in Fig. 8. This figure already suggests that the Eu ions do not occupy the Zn2+ sites in the lattice, but are more likely to be present in the powders as EuGa2S4.

 figure: Fig. 8

Fig. 8 Top: Zn K edge extracted EXAFS spectrum, Middle: Eu LIII edge extracted EXAFS spectrum, both in Zn0.96Eu0.04Ga2S, Bottom: Eu LIII edge extracted EXAFS spectrum in EuGa2S4.

Download Full Size | PDF

To investigate the structure around dopant ions in detail, FEFF simulations were carried out. In the first step, the EuGa2S4 powder was fitted using the known crystallographic structure [47]. In this way the Debye-Waller parameters were determined.

First, a simulation with only EuGa2S4 was performed. The correspondence with the experimental spectrum was already very good. This indicates that the true composition of the powders does not deviate far from this model and the majority of the europium occurs as EuGa2S4, as already established with SEM-CL-EDX and XRD.

Secondly, the simulation where a small amount of the Eu2+ ions substitute on the tetrahedral coordinated zinc sites showed a bigger deviation from the experimental spectrum. This is not surprising considering the limited volume of the Zn first coordination sphere and the size mismatch between Zn2+ and Eu2+. This clearly confirms the hypothesis by Wickleder et al. that Eu2+ is not substituting for Zn2+ [28].

Thirdly, a simulation with a small amount of the Eu on the octahedral voids in ZnGa2S4 and the majority in the form of EuGa2S4, was performed. The simulated Fourier transforms of Eu LIII edge EXAFS are shown in Fig. 9, compared with experimental data. In this case, the Fourier transforms were performed in the k range of [2.1 Å−1 – 7.8 Å−1]. As can be seen in the figure, there is a very good correspondence between the two. The volume of the octahedral voids in the ZnGa2S4 structure is ~24 Å3, being much larger than the volume of the coordination tetrahedron of the Zn site, but still significantly smaller than the volume of the coordination octahedron of other sulfide hosts for Eu2+ (~31 Å3 for CaS, ~36 Å3 for SrS [51]). The simulation also showed that the closest zinc neighbor of the incorporated europium ions disappears. This observation can be explained by charge compensation: the europium ions are divalent as well as the zinc ions. Furthermore it is observed that some of the sulfuric atoms move away from the dopants. This effect can be caused by the vacancy that occurs at the cationic position in the lattice. A crystal model of this simulation is displayed in Fig. 10.

 figure: Fig. 9

Fig. 9 Fourier transform ofk2χ(k)to radial distance (r) space for a Zn0.92Eu0.08Ga2S4 powder (solid line) and the result of the simulation, based on the structure of EuGa2S4 and the europium ions occupying the octahedral voids (circles).

Download Full Size | PDF

 figure: Fig. 10

Fig. 10 Left: Crystal structure of ZnGa2S4 where an Eu2+ dopant occupies an octahedral void (sixfold coordinated). The neighboring zinc ion is removed to compensate for the excess positive charge [45]. Right: Crystal structure of EuGa2S4, with (slightly distorted) square antiprismatic coordination polyhedra for Eu2+ (eightfold coordinated) [47].

Download Full Size | PDF

4. Conclusions

In this paper, green zinc thiogallate phosphors with different doping concentrations were investigated. The luminescent properties of these materials were described in detail. Furthermore, the structure of the phosphor was investigated and the distribution and incorporation of europium in the powders was resolved with XRD, SEM-CL-EDX and XAS measurements. With these complementary techniques, we were able to explain the anomalous luminescent properties of these phosphors.

The emission color of this phosphor is saturated green. The material can be excited with a broad range of blue and violet wavelengths. However, the quantum efficiency of this material is far too low for any application. Furthermore, the thermal quenching is excessive. Since desirable photoluminescence properties should always occur simultaneously with low thermal quenching and high quantum efficiency, this material should be discarded for application as conversion phosphor, despite earlier reports.

Analysis of the XRD and EXAFS experiments proved that only a limited amount of europium – if any - can be incorporated into the ZnGa2S4 lattice. If these ions are incorporated, they occupy octahedral voids instead of the smaller zinc or vacancy sites with tetrahedral coordination. The neighboring zinc ion is then removed to compensate for the excess positive charge. The majority of the europium ions form the EuGa2S4 phase.

Scanning electron microscope measurements, combined with CL and EDX allowed to link light output of the powder particles to their composition. It was found that the green luminescence which is observed by different authors is coming from the small fraction of the stoichiometric EuGa2S4 phosphor that is formed. The majority of the powder particles, consisting of the ZnGa2S4 phase, do not show any light emission.

Acknowledgments

This work is financially supported by the agency for Innovation by Science and Technology (IWT) and the Research Foundation Flanders (FWO). The authors would like to acknowledge Olivier Janssens for the X-ray diffraction measurements. We gratefully thank Koen Van den Eeckhout and Katrien Meert for assisting in the synchrotron data collection.

References and links

1. V. Bachmann, C. Ronda, and A. Meijerink, “Temperature quenching of yellow Ce3+ luminescence in YAG:Ce,” Chem. Mater. 21(10), 2077–2084 (2009). [CrossRef]  

2. R. Mueller-Mach, G. Mueller, M. R. Krames, H. A. Hoppe, F. Stadler, W. Schnick, T. Juestel, and P. Schmidt, “Highly efficient all-nitride phosphor-converted white light emitting diode,” Physica Status Solidi a-Applications and Materials Science 202(9), 1727–1732 (2005). [CrossRef]  

3. I. H. Cho, G. Anoop, D. W. Suh, S. J. Lee, and J. S. Yoo, “On the stability and reliability of Sr1-xBaxSi2O2N2:Eu2+ phosphors for white LED applications,” Opt. Mater. Express 2(9), 1292–1305 (2012). [CrossRef]  

4. R. J. Xie, N. Hirosaki, and T. Takeda, “Wide color gamut backlight for liquid crystal displays using three-band phosphor-converted white light-emitting diodes,” Appl. Phys. Express 2, 022401 (2009). [CrossRef]  

5. M. H. Crawford, “LEDs for solid-state lighting: performance challenges and recent advances,” IEEE J. Sel. Top. Quant. 15(4), 1028–1040 (2009). [CrossRef]  

6. J. H. Oh, S. J. Yang, Y. G. Sung, and Y. R. Do, “Excellent color rendering indexes of multi-package white LEDs,” Opt. Express 20(18), 20276–20285 (2012). [CrossRef]   [PubMed]  

7. J. H. Oh, S. J. Yang, Y. G. Sung, and Y. R. Do, “Improved color coordinates of green monochromatic pc-LED capped with a band-pass filter,” Opt. Express 21(4), 4539–4550 (2013). [CrossRef]   [PubMed]  

8. P. F. Smet, A. B. Parmentier, and D. Poelman, “Selecting conversion phosphors for white light-emitting diodes,” J. Electrochem. Soc. 158(6), R37–R54 (2011). [CrossRef]  

9. T. L. Barry, “Fluorescence of Eu2+-activated phases in binary alkaline earth orthosilicate systems,” J. Electrochem. Soc. 115(11), 1181–1184 (1968). [CrossRef]  

10. Y. Shimomura, T. Honma, M. Shigeiwa, T. Akai, K. Okamoto, and N. Kijima, “Photoluminescence and crystal structure of green-emitting Ca3Sc2Si3O12:Ce3+ phosphor for white light emitting diodes,” J. Electrochem. Soc. 154(1), J35–J38 (2007). [CrossRef]  

11. J. J. Joos, K. W. Meert, A. B. Parmentier, D. Poelman, and P. F. Smet, “Thermal quenching and luminescence lifetime of saturated green Sr1-xEuxGa2S4 phosphors,” Opt. Mater. 34(11), 1902–1907 (2012). [CrossRef]  

12. S. Iida, A. Kato, M. Tanaka, H. Najafov, and H. Ikuno, “Photoluminescence characterization of rare-earth stoichiometric compounds of EuGa2S4,” J. Phys. Chem. Solids 64(9-10), 1815–1819 (2003). [CrossRef]  

13. R. J. Yu, H. K. Yang, B. K. Moon, B. C. Choi, and J. H. Jeong, “Luminescence properties of stoichiometric EuM2S4 (M = Ga, Al) conversion phosphors for white LED applications,” Physica Status Solidi a-Applications and Materials Science 209(12), 2620–2625 (2012). [CrossRef]  

14. C. Braun, M. Seibald, S. L. Börger, O. Oeckler, T. D. Boyko, A. Moewes, G. Miehe, A. Tücks, and W. Schnick, “Material properties and structural characterization of M3Si6O12N2:Eu2+ (M = Ba, Sr) - A comprehensive study on a promising green phosphor for pc-LEDs,” Chemistry 16(31), 9646–9657 (2010). [CrossRef]   [PubMed]  

15. V. Bachmann, T. Justel, A. Meijerink, C. Ronda, and P. J. Schmidt, “Luminescence properties of SrSi2O2N2 doped with divalent rare earth ions,” J. Lumin. 121(2), 441–449 (2006). [CrossRef]  

16. R. J. Xie, N. Hirosaki, Y. Q. Li, and T. Takeda, “Rare-earth activated nitride phosphors: synthesis, luminescence and applications,” Materials 3(6), 3777–3793 (2010). [CrossRef]  

17. M. Seibald, T. Rosenthal, O. Oeckler, C. Maak, A. Tücks, P. J. Schmidt, D. Wiechert, and W. Schnick, “New polymorph of the highly efficient LED-phosphor SrSi2O2N2:Eu2+ – polytypism of a layered oxonitridosilicate,” Chem. Mater. 25(9), 1852–1857 (2013). [CrossRef]  

18. R. J. Xie, N. Hirosaki, H. L. Li, Y. Q. Li, and M. Mitomo, “Synthesis and photoluminescence properties of beta-sialon: Eu2+ (Si6-zAlzOzN8-z:Eu2+),” J. Electrochem. Soc. 154, J314–J319 (2007). [CrossRef]  

19. Y. C. Chiu, C. H. Huang, T. J. Lee, W. R. Liu, Y. T. Yeh, S. M. Jang, and R. S. Liu, “Eu2+-activated silicon-oxynitride Ca3Si2O4N2: a green-emitting phosphor for white LEDs,” Opt. Express 19(S3Suppl 3), A331–A339 (2011). [CrossRef]   [PubMed]  

20. Y. Q. Li, C. M. Fang, G. de With, and H. T. Hintzen, “Preparation, structure and photoluminescence properties of Eu2+ and Ce3+-doped SrYSi4N7,” J. Solid State Chem. 177(12), 4687–4694 (2004). [CrossRef]  

21. M. M. Yuta and W. B. White, “Photoluminescence of sulfide phosphors with the MB2S4 composition,” J. Electrochem. Soc. 139(8), 2347–2352 (1992). [CrossRef]  

22. J. W. Kim and Y. J. Kim, “Synthesis and luminescent characterization of zinc thiogallate,” J. Eur. Ceram. Soc. 27(13-15), 3667–3670 (2007). [CrossRef]  

23. Y. K. Kim, D. H. Cho, Y. K. Jeong, M. K. Nah, K. B. Kim, and J. G. Kang, “Luminescence characterization of (Ca1-xZnx)Ga2S4:Eu2+ phosphors for a white light-emitting diode,” Mater. Res. Bull. 45(8), 905–909 (2010). [CrossRef]  

24. R. Yu, R. X. Luan, C. F. Wang, J. T. Chen, Z. X. Wang, B. K. Moon, and J. H. Jeong, “Photoluminescence properties of green emitting ZnGa2S4:Eu2+ phosphor,” J. Electrochem. Soc. 159(5), J188–J192 (2012). [CrossRef]  

25. Y. K. Jeong, D. H. Cho, K. B. Kim, and J. G. Kang, “Highly luminescent (Zn0.6Sr0.3Mg0.1)2Ga2S5:Eu2+ green phosphors for a white light-emitting diode,” Bull. Korean Chem. Soc. 33(8), 2523–2528 (2012). [CrossRef]  

26. J. W. Kim and Y. J. Kim, “Optical properties of eu doped M-Ga2S4 (M: Zn, Ca, Sr) phosphors for white light emitting diodes,” J. Nanosci. Nanotechnol. 7(11), 4065–4068 (2007). [CrossRef]   [PubMed]  

27. B. G. Tagiev, G. G. Guseinov, R. B. Dzhabbarov, O. B. Tagiev, N. N. Musaeva, and A. N. Georgobiani, “Synthesis and luminescent properties of ZnGa2S4:Eu,F and ZnGa2O4:Eu,F,” Inorg. Mater. 36(12), 1189–1191 (2000). [CrossRef]  

28. C. Wickleder, S. Zhang, and H. Haeuseler, “Photoluminescence of ZnGa2S4:Eu2+,” Z. Kristallogr. 220(2-3-2005), 277–280 (2005). [CrossRef]  

29. H. A. Höppe, “Recent developments in the field of inorganic phosphors,” Angew. Chem. Int. Ed. Engl. 48(20), 3572–3582 (2009). [CrossRef]   [PubMed]  

30. G. Derbyshire, K. C. Cheung, P. Sangsingkeow, and S. S. Hasnain, “A low-profile monolithic multi-element Ge detector for X-ray fluorescence applications,” J. Synchrotron Radiat. 6(2), 62–63 (1999). [CrossRef]  

31. B. Ravel and M. Newville, “ATHENA, ARTEMIS, HEPHAESTUS: data analysis for X-ray absorption spectroscopy using IFEFFIT,” J. Synchrotron Radiat. 12(4), 537–541 (2005). [CrossRef]   [PubMed]  

32. S. H. M. Poort, A. Meyerink, and G. Blasse, “Lifetime measurements in Eu2+-doped host lattices,” J. Phys. Chem. Solids 58(9), 1451–1456 (1997). [CrossRef]  

33. R. B. Jabbarov, C. Chartier, B. G. Tagiev, O. B. Tagiev, N. N. Musayeva, C. Barthou, and P. Benalloul, “Radiative properties of Eu2+ in BaGa2S4,” J. Phys. Chem. Solids 66(6), 1049–1056 (2005). [CrossRef]  

34. P. F. Smet, J. E. Van Haecke, R. L. Van Meirhaeghe, and D. Poelman, “Crystallographic and luminescent properties of orthorhombic BaAl2S4:Eu powder and thin films,” J. Appl. Phys. 98(4), 043512 (2005). [CrossRef]  

35. K. T. Lethi, A. Garcia, F. Guillen, and C. Fouassier, “Investigation of the MS-Al2S3 systems (M = Ca, Sr, Ba) and luminescence properties of europium-doped thioaluminates,” Mat. Sci. Eng. B-Solid 14, 393–397 (1992). [CrossRef]  

36. J. E. Van Haecke, P. F. Smet, and D. Poelman, “Luminescent characterization of CaAl2S4:Eu powder,” J. Lumin. 126(2), 508–514 (2007). [CrossRef]  

37. P. Benalloul, C. Barthou, C. Fouassier, A. N. Georgobiani, L. S. Lepnev, Y. N. Emirov, A. N. Gruzintsev, B. G. Tagiev, O. B. Tagiev, and R. B. Jabbarov, “Luminescence of Eu2+ in calcium thiogallate,” J. Electrochem. Soc. 150(1), G62–G65 (2003). [CrossRef]  

38. V. Bachmann, C. Ronda, O. Oeckler, W. Schnick, and A. Meijerink, “Color point tuning for (Sr,Ca,Ba)Si2O2N2:Eu2+ for white light LEDs,” Chem. Mater. 21(2), 316–325 (2009). [CrossRef]  

39. B. Henderson and G. F. Imbusch, Optical Spectroscopy of Inorganic Solids (Oxford University Press, Incorporated, 2006).

40. A. B. Parmentier, P. F. Smet, F. Bertram, J. Christen, and D. Poelman, “Structure and luminescence of (Ca,Sr)2SiS4:Eu2+ phosphors,” J. Phys. D Appl. Phys. 43(8), 085401 (2010). [CrossRef]  

41. R. D. Shannon, “Revised effective ionic-radii and systematic studies of interatomic distances in halides and chalcogenides,” Acta Crystallogr. A 32(5), 751–767 (1976). [CrossRef]  

42. P. Dorenbos, “Thermal quenching of Eu2+ 5d-4f luminescence in inorganic compounds,” J. Phys.- Condens. Mat. 17(50), 8103–8111 (2005). [CrossRef]  

43. P. Dorenbos, “Energy of the first 4f7-> 4f65d transition of Eu2+ in inorganic compounds,” J. Lumin. 104(4), 239–260 (2003). [CrossRef]  

44. B. G. Yacobi and D. B. Holt, “Cathodoluminescence scanning electron-microscopy of semiconductors,” J. Appl. Phys. 59(4), R1–R24 (1986). [CrossRef]  

45. C. K. Lowe-Ma and T. A. Vanderah, “Structure of ZnGa2S4, a defect sphalerite derivative,” Acta Crystallogr. C 47(5), 919–924 (1991). [CrossRef]  

46. A. Eifler, G. Krauss, V. Riede, V. Kramer, and W. Grill, “Optical phonon modes and structure of ZnGa2Se4 and ZnGa2S4,” J. Phys. Chem. Solids 66(11), 2052–2057 (2005). [CrossRef]  

47. R. Roques, R. Rimet, J. P. Declercq, and G. Germain, “Determination of crystal-structure of EuGa2S4,” Acta Crystallogr. B 35(3), 555–557 (1979). [CrossRef]  

48. V. K. Pecharsky and P. Y. Zavalij, Fundamentals of Powder Diffraction and Structural Characterization of Materials (Springer, 2005).

49. N. Avci, K. Korthout, M. A. Newton, P. F. Smet, and D. Poelman, “Valence states of europium in CaAl2O4:Eu phosphors,” Opt. Mater. Express 2(3), 321–330 (2012). [CrossRef]  

50. K. Korthout, A. B. Parmentier, P. F. Smet, and D. Poelman, “A XAS study of the luminescent Eu centers in thiosilicate phosphors,” Phys. Chem. Chem. Phys. 15(22), 8678–8683 (2013). [CrossRef]   [PubMed]  

51. T. Kobayashi, K. Susa, and S. Taniguchi, “Preparation and semiconductive properties of rock salt type solid-solution systems, Cd1-xMxS (M=Sr, Ca, Mg, Pb, Sn),” J. Phys. Chem. Solids 40(10), 781–785 (1979). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1
Fig. 1 Top: Diffuse reflectance spectra for ZnGa2S4 (x = 0.00), Zn0.99Eu0.01Ga2S4 (x = 0.01) and EuGa2S4 (x = 1.00) at room temperature. Bottom: PL emission (at 450 nm excitation, solid lines) and excitation (at 540 nm emission, dashed lines) spectra for Zn0.99Eu0.01Ga2S4 (x = 0.01) and EuGa2S4 (x = 1.00) at room temperature, normalized for better readability. The dip in the excitation spectra around 470 nm is an artifact.
Fig. 2
Fig. 2 Absorption (fraction of the incident photons that are absorbed), internal and external quantum efficiency for Zn1-xEuxGa2S4 at room temperature, measured at 460 nm excitation.
Fig. 3
Fig. 3 Normalized PL intensity and luminescent lifetimes as function of temperature for Zn0.94Eu0.06Ga2S4 upon excitation at 400 nm.
Fig. 4
Fig. 4 SEM-CL-EDX mapping of Zn0.99Eu0.01Ga2S4 powder, obtained at 250 K. Top left: phosphor morphology obtained by backscattered electron imaging. The integrated CL emission intensity is shown as an overlay in false colors. Top right: elemental distribution by EDX, where the colors are determined by color coding with green for Zn and red for Eu. Simultaneous detection would lead to a yellow color. Bottom: CL emission spectra for the points indicated in the top right figure, obtained in clockwise direction.
Fig. 5
Fig. 5 X-ray diffraction patterns for Zn1-xEuxGa2S4 powders, compared with reference patterns for ZnGa2S4 (ICSD 69542) and EuGa2S4 (ICSD 8053) [45, 47].
Fig. 6
Fig. 6 y(x), as defined in Eq. (2).
Fig. 7
Fig. 7 Eu LIII XANES spectrum of (top) Zn0.96Eu0.04Ga2S4, compared with the spectra of reference compounds for Eu2+ and Eu3+ (bottom), EuS (solid line) and Eu2O3 (dashed line).
Fig. 8
Fig. 8 Top: Zn K edge extracted EXAFS spectrum, Middle: Eu LIII edge extracted EXAFS spectrum, both in Zn0.96Eu0.04Ga2S, Bottom: Eu LIII edge extracted EXAFS spectrum in EuGa2S4.
Fig. 9
Fig. 9 Fourier transform of k 2 χ(k) to radial distance (r) space for a Zn0.92Eu0.08Ga2S4 powder (solid line) and the result of the simulation, based on the structure of EuGa2S4 and the europium ions occupying the octahedral voids (circles).
Fig. 10
Fig. 10 Left: Crystal structure of ZnGa2S4 where an Eu2+ dopant occupies an octahedral void (sixfold coordinated). The neighboring zinc ion is removed to compensate for the excess positive charge [45]. Right: Crystal structure of EuGa2S4, with (slightly distorted) square antiprismatic coordination polyhedra for Eu2+ (eightfold coordinated) [47].

Tables (2)

Tables Icon

Table 1 Overview of luminescent properties of green phosphors from recent literature. T0.5 describes the thermal quenching. It is the temperature at which the emission intensity is 50% of the value at low temperature. QE (int/ext) is the internal/external quantum efficiency.

Tables Icon

Table 2 Emission maxima (λmax) and full width at half maximum (FWHM), luminescent lifetimes (τ) and quenching temperatures (T0.5) of multiple Eu2+ doped thiogallate and thioaluminate phosphors. The displayed parameters were recorded at room temperature.

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

η ext = N em N inc η int = N em N abs η ext =A η int
y(x)= I EuGa 2 S 4 I ZnGa 2 S 4 1x x
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.