Next Article in Journal
Effect of Alumina Incorporation on the Surface Mineralization and Degradation of a Bioactive Glass (CaO-MgO-SiO2-Na2O-P2O5-CaF2)-Glycerol Paste
Next Article in Special Issue
Tuning the Photoluminescence of Graphene Quantum Dots by Photochemical Doping with Nitrogen
Previous Article in Journal
The Biological Properties of OGI Surfaces Positively Act on Osteogenic and Angiogenic Commitment of Mesenchymal Stem Cells
Previous Article in Special Issue
Optical Characterization of Nano- and Microcrystals of EuPO4 Created by One-Step Synthesis of Antimony-Germanate-Silicate Glass Modified by P2O5
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Co-Precipitation Synthesis and Optical Properties of Mn4+-Doped Hexafluoroaluminate w-LED Phosphors

Condensed Matter and Interfaces, Debye Institute for Nanomaterials Science, Utrecht University, P.O. Box 80000, 3508 TA Utrecht, The Netherlands
*
Author to whom correspondence should be addressed.
Materials 2017, 10(11), 1322; https://doi.org/10.3390/ma10111322
Submission received: 29 September 2017 / Revised: 2 November 2017 / Accepted: 7 November 2017 / Published: 17 November 2017
(This article belongs to the Special Issue Luminescent Materials 2017)

Abstract

:
Mn4+-activated hexafluoroaluminates are promising red-emitting phosphors for white light emitting diodes (w-LEDs). Here, we report the synthesis of Na3AlF6:Mn4+, K3AlF6:Mn4+ and K2NaAlF6:Mn4+ phosphors through a simple two-step co-precipitation method. Highly monodisperse large (~20 μm) smoothed-octahedron shaped crystallites are obtained for K2NaAlF6:Mn4+. The large size, regular shape and small size distribution are favorable for application in w-LEDs. All Mn4+-doped hexafluoroaluminates show bright red Mn4+ luminescence under blue light excitation. We compare the optical properties of Na3AlF6:Mn4+, K3AlF6:Mn4+ and K2NaAlF6:Mn4+ at room temperature and 4 K. The luminescence measurements reveal that multiple Mn4+ sites exist in M3AlF6:Mn4+ (M = Na, K), which is explained by the charge compensation that is required for Mn4+ on Al3+ sites. Thermal cycling experiments show that the site distribution changes after annealing. Finally, we investigate thermal quenching and show that the luminescence quenching temperature is high, around 460–490 K, which makes these Mn4+-doped hexafluoroaluminates interesting red phosphors for w-LEDs. The new insights reported on the synthesis and optical properties of Mn4+ in the chemically and thermally stable hexafluoroaluminates can contribute to the optimization of red-emitting Mn4+ phosphors for w-LEDs.

1. Introduction

White light emitting diodes (w-LEDs) are nowadays widely applied in general lighting and consumer electronics because of their high energy efficiency and long operation lifetime [1,2,3,4,5]. Commercial w-LEDs generate white light by combining blue-emitting (In,Ga)N LED chips with inorganic phosphors that convert part of the blue LED emission to green, yellow and/or red light [5,6]. Currently, the typical red phosphors in w-LEDs are Eu2+-doped nitrides. These phosphors can have quantum yields (QYs) exceeding 90%, but a drawback is that the Eu2+ emission band extends into the deep red and near-IR regions where the sensitivity of the human eye is low [6,7,8]. As a result, there are significant efficacy losses (reduced lumen/W output) at high color rendering indices (CRIs) when using Eu2+-doped nitrides as red phosphors in w-LEDs [9]. To overcome this issue, efficient narrow band red-emitting phosphors with λmax ~ 620 nm have to be developed.
Mn4+-doped fluorides are a promising class of materials to improve the color rendering of w-LEDs while maintaining a high luminous efficacy [8,9,10,11]. Upon blue photoexcitation Mn4+-doped fluorides show narrow red line emission in the 600–640 nm spectral region. Furthermore, they can have a QY > 90% and are prepared through simple wet-chemical synthesis at room temperature [8,12]. These characteristics make Mn4+-doped fluorides interesting narrow band red phosphors for w-LEDs. In recent years, a large number of Mn4+-activated fluorides with the general chemical formulas A2MF6:Mn4+ (A = Na, K, Rb, Cs and NH4; M = Si, Ge, Ti, Zr and Sn) and BaMF6:Mn4+ (M = Si, Ge, Ti and Sn) have been reported [11,12,13]. In these compounds Mn4+ substitutes for the M4+ cation that is octahedrally coordinated by six F ions. Most work has been done on the phosphor K2SiF6:Mn4+ (KSF) [14,15]. The Mn4+-doped fluorides have excellent luminescence properties. The deep red color of the 600–640 nm Mn4+ emission is particularly favorable for extending the color gamut of displays, and KSF is already widely applied in displays. In lighting, large-scale application is still limited, partly hampered by issues related to thermal and chemical stability and saturation at high pump powers [14,16].
Recently, the synthesis and luminescence properties of Mn4+-doped hexafluoraluminates with the compositions M3AlF6:Mn4+ (M = Li, Na, K) were reported [17,18,19,20]. The ionic radius of Mn4+ is similar to the ionic radius of Al3+ (0.53 versus 0.54 Å), and as a result, Mn4+ can easily substitute for Al3+ in fluoride hosts [17,21]. The M3AlF6:Mn4+ phosphors have potential advantages over K2SiF6:Mn4+ and other Mn4+-doped fluorides. First, Na3AlF6 and K3AlF6 have a melting point of ~1000 °C and therefore have a much better thermal stability than K2SiF6, which already decomposes at 350–400 °C [10,22,23,24]. Secondly, the M3AlF6 compounds have a lower water solubility than K2SiF6, making them more chemically stable under high humidity conditions [14,25]. Thirdly, hexafluoroaluminates are already produced worldwide on a large scale, as they are used as solvents for bauxite in the industrial extraction of aluminum [26]. This may facilitate cheap large-scale production of Mn4+-activated hexafluoroaluminates.
Until now, different methods have been used to synthesize M3AlF6:Mn4+ phosphors. Song et al. prepared Na3AlF6:Mn4+ phosphors via a co-precipitation method [17], while others synthesized K3AlF6:Mn4+ and K2NaAlF6:Mn4+ by cation exchange [18,19]. Furthermore, K2LiAlF6:Mn4+ was synthesized via a hydrothermal method [20]. A single convenient synthesis method for preparing M3AlF6:Mn4+ phosphors is thus so far lacking. The previous works on M3AlF6:Mn4+ have reported luminescence spectra and decay curves for different Mn4+ doping concentrations in the temperature range of 300 to 500 K [17,18,19,20]. However, they provided no insight into the influence of (charge compensating) defects on the luminescence spectra and quantum yield of M3AlF6:Mn4+. Furthermore, no explanations for the thermal quenching of the Mn4+ luminescence were given.
In this work we report a new synthesis route for Na3AlF6:Mn4+, K3AlF6:Mn4+ and K2NaAlF6:Mn4+ based on a simple two-step co-precipitation method. We synthesize Mn4+-doped hexafluoroaluminates by initially preparing the Mn4+-precursor K2MnF6 and then in a second step precipitating M3AlF6:Mn4+ (M = Na, K) from an aqueous HF solution containing Al3+, Mn4+ and Na+/K+ ions. The presented method gives good control over the composition and homogeneity of the M3AlF6:Mn4+ phosphors. All synthesized M3AlF6:Mn4+ phosphors exhibit bright red Mn4+ luminescence around 620 nm. For K2NaAlF6:Mn4+ we obtain highly monodisperse and large (~20 µm) phosphor particles that exhibit narrow size and shape distributions that are superior to the size and shape distributions of other reported Mn4+-doped phosphors. This makes K2NaAlF6:Mn4+ interesting for use in w-LEDs, as monodisperse, uniform and large particles are favorable for reproducible and high packing density of phosphors in w-LEDs.
We perform both room-temperature and low-temperature (T = 4 K) spectral measurements of M3AlF6:Mn4+. The measurements at 4 K reveal that multiple Mn4+ sites exist in M3AlF6:Mn4+, which was not observed in previous works on Mn4+-doped hexafluoroaluminates. The formation of multiple Mn4+ sites can be understood from the need for charge compensation for Mn4+ ions on an Al3+ site. Further evidence for the presence of multiple sites is obtained from thermal cycling experiments, which show a change in site distribution after high temperature annealing. The charge compensation and associated defects have a large influence on the luminescence properties (e.g., quantum yield) of M3AlF6:Mn4+.
Finally, we study the thermal quenching behavior for M3AlF6:Mn4+ by measuring the luminescence intensity as a function of temperature between 300 and 600 K. The luminescence quenching temperature we find for M3AlF6:Mn4+ is 460–490 K, which is above the operating temperature of phosphors in high-power w-LEDs. The thermal quenching is explained by thermally activated crossover from the 4T2 excited state to the 4A2 ground state. Furthermore, there is luminescence quenching due to non-radiative energy transfer from Mn4+ ions to quenching sites (defects and impurities).

2. Materials and Methods

The M3AlF6:Mn4+ (M = Na, K) phosphors were synthesized through a two-step chemical co-precipitation method. In the first step the Mn4+-precursor K2MnF6 was synthesized and in the second step the M3AlF6:Mn4+ phosphor was prepared. Since K2MnF6 and M3AlF6:Mn4+ are synthesized in corrosive HF solutions, all reactions described were carried out in plastic or Teflon beakers.

2.1. Chemicals

The following chemicals were purchased from Sigma-Aldrich: KMnO4 (≥99.0%), KF (≥99.0%), H2O2 (30 wt % solution, ACS reagent), Al(OH)3 (reagent grade), Na2CO3 (≥99.95%) and K2CO3 (≥99.0%). Hydrofluoric acid (HF, 40% aqueous solution) was purchased from Riedel de Haën. All chemicals were used without any further purification.

2.2. Synthesis of K2MnF6

K2MnF6 was prepared according to the method of Bode [27,28]. KMnO4 (4 g) and KF (59.5 g) were dissolved in 250 mL of a 40% HF solution. The black-purple solution obtained was stirred for 30 min and then cooled with an ice bath. Under constant cooling and stirring, a 30 wt % H2O2 solution was added dropwise which resulted in the gradual precipitation of yellow K2MnF6 powder. The dropwise addition of H2O2 was stopped when the reaction solution turned from purple to red-brown, indicating the formation of Mn4+. The K2MnF6 product was isolated by decanting the red-brown solution, washing the precipitate twice with 100 mL of acetone and finally drying the precipitate at 60 °C for 5 h.

2.3. Synthesis of M3AlF6:Mn4+

Al(OH)3 (10 mmol, 0.78 g) was dissolved in 15 mL 40% HF by stirring and heating at 60 °C. After cooling down to room temperature, 0.1 mmol K2MnF6 (1 mol % doping concentration) was added. Simultaneously, a solution of M+ ions (M = Na, K) in 40% HF was prepared by gradually adding M2CO3 or MF to 40% HF (aq) while stirring. Table 1 lists the amounts of Na2CO3, K2CO3, KF and 40% HF used for preparing the M+/HF solutions. The M+/HF solution was added to the Al3+/Mn4+/HF solution, which resulted in the precipitation of Na3AlF6:Mn4+ and K2NaAlF6:Mn4+ but not K3AlF6:Mn4+ phosphor. K3AlF6:Mn4+ was precipitated by adding 50 mL of ethanol to the mixed M+/Al3+/Mn4+/HF solution (ethanol acts as anti-solvent). The synthesized phosphors were isolated by decanting the solution, washing the precipitate twice with ethanol and subsequently drying at 75 °C for 3 h. Different Mn4+ doping concentrations could be obtained by changing the amount of K2MnF6 that was used in the synthesis.

2.4. Characterization

Powder X-ray diffraction (XRD) patterns were measured on a Philips PW1729 X-ray diffractometer using Cu Kα radiation (λ = 1.5418 Å). Scanning electron microscopy (SEM) images of the phosphors were obtained using a Philips XL30S FEG microscope operating at 20 keV. The manganese concentrations in the phosphors were determined with inductively coupled plasma optical emission spectroscopy (ICP-OES) performed on a Perkin-Elmer Optima 8300 spectrometer. For the ICP-OES measurements the M3AlF6:Mn4+ phosphors were dissolved in aqua regia.
Photoluminescence (PL) measurements were performed on an Edinburgh Instruments FLS900 fluorescence spectrometer equipped with a double 0.22 m excitation monochromator. For recording emission and excitation spectra, we used a 450 W Xe lamp as excitation source and a Hamamatsu R928 photomultiplier tube (PMT) to detect the emission. For PL measurements down to 4 K, the phosphors were cooled in an Oxford Instruments liquid helium flow cryostat. PL measurements between 300 K and 600 K were performed by heating the phosphors with a Linkam THMS600 temperature controlled stage. PL quantum yields were determined with a calibrated home-built setup which consisted of a 65 W Xe lamp, excitation monochromator, integrating sphere (Labsphere) and CCD camera (Avantes AvaSpec-2048).

3. Results and Discussion

3.1. Structural Properties

To investigate the composition, size and shape of the M3AlF6:Mn4+ (M = Na, K) phosphor particles, we used different characterization techniques such as XRD, SEM and ICP-OES. Figure 1 shows XRD patterns of the M3AlF6:Mn4+ phosphors. The XRD patterns are in perfect agreement with the reference patterns for Na3AlF6 (PDF 00-025-0772, red), K2NaAlF6 (PDF 01-072-2434, blue) and K3AlF6 (PDF 00-057-0227, green). No other crystal phases can be observed, which confirms that the phosphor samples are single phase.
The XRD measurements show that incorporation of Mn4+ on the Al3+ sites does not significantly change the crystal structure of M3AlF6, which is expected as the ionic radii of Mn4+ and Al3+ are similar (0.53 versus 0.54 Å) [21]. The space groups of Na3AlF6, K2NaAlF6 and K3AlF6 are listed in Table 2. K2NaAlF6 has a highly symmetric cubic crystal structure (space group is Fm 3 ¯ m), while Na3AlF6 and K3AlF6 have a crystal structure with lower symmetry (space groups are P21/n and I41/a, respectively) [29,30,31]. In the M3AlF6 crystal structure, the Al3+ ions are octahedrally coordinated by six F ions. Depending on the M3AlF6 lattice, the AlF6 octahedron can be highly symmetric (Oh in K2NaAlF6) or distorted (Ci in Na3AlF6 and C1 in K3AlF6). The average Al–F distances in the AlF6 octahedra are around 1.8 Å.
By using ICP-OES, we measured the manganese concentrations in the synthesized M3AlF6:Mn4+phosphors (see Materials and Methods section). The XRD patterns displayed in Figure 1 were measured for M3AlF6:Mn4+ phosphors containing 0.4 mol % (Na3AlF6:Mn4+), 1.2 mol % (K3AlF6:Mn4+) and 0.9 mol % Mn (K2NaAlF6:Mn4+). The results presented in the rest of this work were obtained for M3AlF6:Mn4+ phosphors having these doping concentrations. For K3AlF6:Mn4+ and K2NaAlF6:Mn4+ the measured Mn concentration is very close to the 1 mol % Mn added during the synthesis, which demonstrates that our synthesis method provides good control over the Mn4+ doping concentration. Also a substantially higher fraction of Mn4+ is incorporated into K3AlF6 compared to previously reported cation exchange methods for preparing K3AlF6:Mn4+ [18].
Figure 2 displays SEM images of Na3AlF6:Mn4+ (0.4%), K3AlF6:Mn4+ (1.2%) and K2NaAlF6:Mn4+ (0.9%) phosphor particles. The Na3AlF6:Mn4+ (Figure 2a) and K3AlF6:Mn4+ (Figure 2b) phosphors consist of irregularly shaped particles and clusters of particles with sizes ranging from ~100 nm to >10 µm. For K3AlF6:Mn4+, we attribute the large variety in shape and size to the rapid and forced crystallization following addition of the anti-solvent ethanol. In contrast, the synthesis of K2NaAlF6:Mn4+ (Figure 2c,d) yields highly monodisperse phosphor particles with a large average diameter of 22.5 ± 6.1 µm. The K2NaAlF6:Mn4+ particles have a smoothed octahedral shape, as expected from the cubic elpasolite structure of K2NaAlF6 [10,19,29]. The K2NaAlF6:Mn4+ particles prepared with our co-precipitation method have a more uniform shape and size than the K2NaAlF6:Mn4+ particles prepared by Zhu et al. via cation exchange [19]. Moreover, the K2NaAlF6:Mn4+ (0.9%) phosphor shown in Figure 2c,d exhibits particle size and shape distributions that are superior to the size and shape distributions of other reported Mn4+-doped fluoride phosphors [10,24,32,33,34,35,36,37].
The high monodispersity of the K2NaAlF6:Mn4+ crystallites reported here may originate from the synthesis method used. In order to achieve a narrow size distribution, it is necessary to temporally separate the particle nucleation and growth stages [38]. As described in the Materials and Methods section, we dissolve all the starting materials in HF solutions prior to the formation of the phosphor particles. Mixing of the precursor solutions results in oversaturation and the rapid formation of crystal nuclei. Once the nuclei have been formed, the precursor concentrations drop and no new nuclei are formed. The particles can grow to monodisperse and large crystallites during the growth stage. This differs from syntheses where Mn4+-doped particles are prepared via cation exchange or chemical etching [14]. With these methods, new precursor ions are constantly supplied to the reaction solution preventing temporal separation of nucleation and growth.
The superior size distribution gives K2NaAlF6:Mn4+ potential advantages in LED applications. Monodisperse and large crystallites are favorable for uniform and reproducible packing of phosphors, which is very important in the production of w-LEDs. In addition, phosphors with large and highly crystalline particles often display higher quantum yields because of reduced concentrations of defects which act as quenching sites. Finally, a large particle size aids the long-term stability of a phosphor.

3.2. Mn4+Luminescence

Figure 3 shows the room-temperature emission and excitation spectra of M3AlF6:Mn4+ (M = Na, K). Upon blue photoexcitation the Mn4+-doped hexafluoroaluminates show narrow red emission lines around 620 nm (see Figure 3b). The emission lines are in a spectral range where the eye sensitivity is still relatively high, which is beneficial for applications. The red emission lines are assigned to spin- and parity-forbidden Mn4+ 2E → 4A2 transitions. Figure 3a shows that the red luminescence of M3AlF6:Mn4+ has two broad excitation bands in the ultraviolet (UV) to blue spectral region. These bands correspond to the 4A24T1 and 4A24T2 transitions of Mn4+. In all three lattices the 4A24T2 excitation band is positioned at 460 nm, which indicates that the crystal field splitting is approximately equal for the investigated M3AlF6 hosts.
The emission spectra of M3AlF6:Mn4+ in Figure 3b resemble the emission spectra observed for other Mn4+-doped fluoride phosphors [33,39,40]. By analogy, the 2E → 4A2 emission spectra of M3AlF6:Mn4+ consist of a zero-phonon line (ZPL) at ~620 nm and anti-Stokes and Stokes vibronic (ν3, ν4 and ν6) lines on the high and low energy side of the ZPL, respectively [39]. The Mn4+ emission spectra are dominated by vibronic lines because the parity selection rule is relaxed by coupling of the 2E → 4A2 electronic transition with odd-parity vibrations (ν3, ν4 and ν6 vibrational modes) [39,41]. It is noted that the 2E → 4A2 ZPL intensity of M3AlF6:Mn4+ is relatively strong when compared to other Mn4+-doped fluorides. For example, in K2SiF6:Mn4+ the ZPL intensity is less than 5% of the Stokes ν6 intensity, while in K3AlF6:Mn4+ the ZPL intensity is almost half of the Stokes ν6 intensity [40]. The intense ZPL is an interesting property for (w-LED) applications, as it improves the color quality of the red Mn4+ phosphor. Furthermore, the observation of relatively strong zero-phonon lines is a sign that the MnF6 centers lack inversion symmetry. The presence of odd-parity crystal field components relaxes the parity selection rule by inducing mixing with opposite parity states. As a result, the radiative lifetime of the 2E → 4A2 emission is shorter (which is beneficial to reduce saturation effects) and the 4A24T2 absorption is stronger (and thus less material or a lower Mn4+ concentration is needed to absorb the desired fraction of blue LED light).
The strong ZPL intensity in K3AlF6:Mn4+ can be attributed to the low symmetry for Mn4+ on the Al3+ sites, i.e., the AlF6 octahedron lacks an inversion center (C1 symmetry, see Table 2). As discussed, without inversion symmetry, the 2E → 4A2 ZPL intensity increases due to relaxation of the parity selection rule by odd-parity crystal field components that mix odd-parity states into the 3d wavefunctions [42]. In Na3AlF6 and K2NaAlF6 the AlF6 octahedra have inversion centers (Ci and Oh symmetry, respectively) and the 2E → 4A2 ZPL is expected to be weak, since there are no odd-parity crystal field components to relax the parity selection rule. The emission spectra in Figure 3b however show that the ZPLs of Na3AlF6:Mn4+ and K2NaAlF6:Mn4+ are significant, which indicates that at least for a part of the Mn4+ ions the site symmetry is lower than Ci (no inversion symmetry). This we explain by the charge compensation required for the Mn Al sites (the Kröger-Vink notation is used to identify defects). The proximity of charge compensating defects (probably V K , Na vacancies or F i interstitials) gives rise to local deformation of the MnF6 octahedra and lifts the inversion symmetry, causing the 2E → 4A2 ZPL intensity of Na3AlF6:Mn4+ and K2NaAlF6:Mn4+to increase.
Besides influencing the 2E → 4A2 ZPL intensity, the charge compensating defects have a large influence on the integrated luminescence intensity and emission lifetime of M3AlF6:Mn4+. Typical luminescence quantum yield (QY) values measured for the M3AlF6:Mn4+ phosphors are 39% for Na3AlF6:Mn4+, 53% for K2NaAlF6:Mn4+ and 55% for K3AlF6:Mn4+. These QY values are below unity, which is attributed to non-radiative transfer of excitation energy from Mn4+ to crystal defects. At the defects the excitation energy is lost non-radiatively as heat, i.e., the defects act as quenching sites. The defect concentration will increase with the Mn4+ concentration, and it is therefore expected that the luminescence quenching becomes stronger at higher Mn4+ concentrations. Previously, it has been measured that the luminescence intensity and emission lifetime of M3AlF6:Mn4+ significantly decrease with increasing Mn4+ concentration already at doping levels of 4% Mn4+ [17,18,19]. In these works, the decrease in intensity and lifetime with the Mn4+ concentration was explained by concentration quenching, i.e., energy migration between Mn4+ ions to quenchers (defects, impurities). Energy migration is however not expected at doping concentrations of 4%, as most Mn4+ ions will not have Mn4+ neighbors in this concentration range (see also [43]). The results presented here indicate that quenching becomes stronger due to an increase in the defect concentration connected to the need for charge compensation and not because of enhanced energy migration among the Mn4+ ions.
We performed low-temperature (T = 4 K) spectral measurements to get more insight into the Mn4+ sites in M3AlF6:Mn4+. In addition, the measurements at 4 K allow us to accurately compare the energy of the emitting 2E level in the different M3AlF6 hosts. Figure 4 displays emission spectra (λexc = 460 nm) at T = 4 K of M3AlF6:Mn4+. In line with the luminescence spectra at room temperature, the 4 K emission spectra of M3AlF6:Mn4+ consist of zero-phonon and vibronic 2E → 4A2 emission lines (labeled ZPL, ν3, ν4 and ν6). Multiple lines are observed in the ZPL region. The peaks marked with a star can be due to 2E → 4A2 electronic transitions that couple with low energy rotatory or translatory lattice vibrational modes [40,44]. Vibronic lines due to these modes are usually found at 50–150 cm−1 relative to the ZPL in emission spectra of Mn4+. Alternatively, the peaks marked with a star can be ZPLs of Mn4+ ions located on different sites than the Mn4+ ions yielding the most intense zero-phonon emission line (labeled ZPL in Figure 4). Mn4+ emission lines caused by lattice modes are typically very weak, and therefore it is more probable that the peaks marked with stars are ZPLs of Mn4+ ions with other geometric environments [33,44]. In addition, multiple Mn4+ sites can be expected, based on the need for charge compensation. Below, further evidence is given which indicates that the various sharp emission lines around 620 nm arise from MnF6 groups with different local geometries related to charge compensation.
In Figure 5 we investigate the presence of multiple Mn4+ sites by measuring 4 K luminescence spectra of K2NaAlF6:Mn4+ for various excitation and emission wavelengths. The excitation spectra in Figure 5a show that the structure in the 4A24T2 excitation band of K2NaAlF6:Mn4+ changes significantly with emission wavelength. If only one Mn4+ site would be present in K2NaAlF6:Mn4+, the excitation spectrum will have the same shape and structure for all emission wavelengths. However, here, the excitation spectrum changes significantly with emission wavelength, which indicates that more than one Mn4+ site is present in K2NaAlF6:Mn4+. Furthermore, the spectra in Figure 5b show that the shape of the 2E → 4A2 spectrum is different for various excitation wavelengths within the 4A24T2 band. This confirms that multiple Mn4+ sites exist in K2NaAlF6:Mn4+, and likely also in K3AlF6:Mn4+ and Na3AlF6:Mn4+. The presence of more than one Mn4+ site in M3AlF6:Mn4+ was not observed in previous reports on Mn4+-doped hexafluoraluminates [17,18]. Instead, from time-resolved measurements it was concluded that only one Mn4+ emission site was present in M3AlF6:Mn4+. The formation of geometrically different Mn4+ sites in M3AlF6:Mn4+ is expected as charge compensation is required for the Mn Al center. The charge compensating defect can be local or distant, i.e., in the first shell of cations around the Mn4+ ion or further away in the lattice. A distant defect will not influence the local geometry around the Mn4+ ion, whereas a local defect can cause a deformation of the fluorine octahedron around the Mn4+ ion. This will give rise to multiple differently charge compensated Mn4+ sites within the lattice, depending on the type and local geometry of charge compensation.
To study the influence of the M3AlF6 (M = Na, K) host on the energy of the Mn4+2E level, we compare the positions of the highest-energy 2E → 4A2 ZPLs in M3AlF6:Mn4+. The energies of these ZPLs (labeled ZPL in Figure 4) are 16,200 cm−1 (K3AlF6:Mn4+), 16,167 cm−1 (Na3AlF6:Mn4+) and 16,082 cm−1 (K2NaAlF6:Mn4+). The 2E level energies are also listed in Table 2. The energy of the 2E level for Mn4+ in M3AlF6 is in good agreement with the 2E level energy observed for Mn4+ in other fluoride hosts [45,46]. Furthermore, it is observed that the energy of the Mn4+ 2E level increases from K2NaAlF6 to Na3AlF6 to K3AlF6 (see dashed line in Figure 4). This indicates that the local structure and type of M+ cation in the second coordination sphere around the Mn4+ ion has an influence on the 2E level energy. Previous studies on M2SiF6:Mn4+ (M = Na, K, Rb or Cs) also show an influence of the M+ cation on the 2E level energy [33,40,44]. In these compounds the energy E of the 2E level follows the trend E(Na+) > E(K+) > E(Rb+) > E(Cs+), which suggests that the 2E level energy decreases with the radius or electron affinity of the M+ ion [21]. This is however not confirmed by our results for the M3AlF6:Mn4+ phosphors, where the highest 2E energy is found for K3AlF6:Mn4+. Instead, the results in Table 2 indicate that the energy of the 2E level increases when the Mn–F (Al–F) distance becomes longer or when the symmetry of the Mn4+ site (Al3+ site) is reduced. It is however not possible to draw definite conclusions from the data in Table 2 as the symmetry and distances in (part of) the MnF6 octahedra will change due to deformations caused by nearby charge compensating defects.

3.3. Thermal Quenching in M3AlF6:Mn4+

For high-power w-LED applications, the thermal quenching behavior of a phosphor is very important, as the temperature of the on-chip phosphor layer in a high-power w-LED reaches temperatures as high as 450 K. The thermal quenching behavior of K3AlF6:Mn4+ and Na3AlF6:Mn4+ has been investigated by Song et al. [17,18]. They reported that thermal quenching sets in around 423 K (150 °C) for K3AlF6:Mn4+ and Na3AlF6:Mn4+. More interestingly, they found that the integrated photoluminescence (PL) intensity of these phosphors doubles between room temperature and 423 K. This is a very useful property for high temperature applications, but is also very unexpected. For most Mn4+-doped fluorides, the PL intensity is relatively constant between room temperature and the temperature at which thermal quenching sets in [8,9,10].
To investigate the thermal quenching of the Mn4+ emission in M3AlF6:Mn4+ (M = Na, K), we measure the integrated PL intensity of M3AlF6:Mn4+ as a function of temperature between 298 K and 600 K. Figure 6a shows emission spectra of K3AlF6:Mn4+ recorded in this temperature range. Emission spectra of Na3AlF6:Mn4+ and K2NaAlF6:Mn4+ between 298 K and 600 K can be found in Figure S1. The results show that the PL intensity of M3AlF6:Mn4+ slowly decreases up to 423 K (150 °C). Above this temperature, there is rapid quenching, with no emission intensity remaining at 573 K. After heating to 573 K, around half of the initial room-temperature PL intensity is retained. Part of the PL intensity is lost upon heating due to e.g., degradation of the phosphor, reduction/oxidation of the Mn4+ ions and formation of Mn-oxide impurities.
From the emission spectra recorded between 298 K and 600 K we obtain the temperature dependence of the integrated PL intensity (IPL), which is displayed in Figure 6b. The intensity is given relative to the integrated PL intensity at room temperature (IRT). The PL intensity gradually decreases between 300 K and 450 K, but then at higher temperatures rapidly drops due to an increased probability for non-radiative transitions from the 2E excited state (luminescence quenching). The luminescence quenching temperature T½, the temperature at which the PL intensity is half of its initial value, is around 460 K for K3AlF6:Mn4+ and Na3AlF6:Mn4+ and 485 K for K2NaAlF6:Mn4+. The T½ values fall in the range of quenching temperatures reported for Mn4+-doped fluoride phosphors [39,47]. The quenching temperatures of M3AlF6:Mn4+ are above the phosphor operating temperatures of high-power w-LEDs.
The temperature dependences we obtain for K3AlF6:Mn4+ and Na3AlF6:Mn4+ (Figure 6b) are significantly different from the work by Song et al. [17,18]. We observe a small decrease in the PL intensity between 298 K and 423 K, while they reported a doubling of the PL intensity between these temperatures. The decrease in PL intensity between 298 K and 423 K for M3AlF6:Mn4+ we ascribe to an increase of the energy transfer from Mn4+ ions to quenching sites (defects and impurities) with temperature [48]. The rapid quenching of the Mn4+ luminescence above 430 K is attributed to thermally activated crossing of the Mn4+ 4T2 excited state and 4A2 ground state, as is explained in [49].
Finally, we observe some interesting changes in the emission spectrum of K2NaAlF6:Mn4+ upon heating to 573 K. This is illustrated in Figure 7, which displays emission spectra recorded at T = 4 K and 298 K of K2NaAlF6:Mn4+ (0.9%) for as-synthesized K2NaAlF6:Mn4+ (blue spectra) and K2NaAlF6:Mn4+ phosphor that was heated to 573 K (red spectra). The room-temperature spectra in Figure 7a show that the structure of the 2E → 4A2 ZPL emission in K2NaAlF6:Mn4+ changes upon heating to 573 K. This effect is even more pronounced in the spectra measured at 4 K (Figure 7b). Four ZPLs of similar intensity are observed for the as-synthesized phosphor, while two ZPLs dominate the spectrum after heating at 573 K. In addition, the results in Figure 7b show that heating to 573 K changes the structure and intensity of the vibronic 2E → 4A2 emission lines. The changes in the emission spectra can be caused by a phase transition in the K2NaAlF6 crystal structure. However, the XRD patterns of as-synthesized K2NaAlF6:Mn4+ and K2NaAlF6:Mn4+ phosphor heated to 573 K both match the reference pattern of elpasolite K2NaAlF6, which indicates that no phase transition occurs (see Figure S2). We therefore explain the changes in the emission spectra upon high temperature annealing by a rearrangement of the Mn4+ sites in K2NaAlF6:Mn4+ upon heating to 573 K. Furthermore, the fact that two ZPLs dominate the emission spectrum of K2NaAlF6:Mn4+ after heating indicates that there is a redistribution in the abundance of different charge compensated Mn4+ sites. The results presented in Figure 7 show that post-synthesis heating can have a large effect on the luminescence properties of M3AlF6:Mn4+ and other Mn4+-doped fluoride phosphors.

4. Conclusions

Mn4+-doped fluorides have recently attracted considerable attention due to their potential for application in w-LEDs. For application in w-LEDs, it is important to understand and control the synthesis and luminescence properties of Mn4+-doped fluoride phosphors. Here, we report the synthesis of different M3AlF6:Mn4+ (M = Na, K) phosphors via a simple two-step co-precipitation method. Our synthesis method provides good control over Mn4+ doping and yields highly monodisperse ~20 μm smoothed octahedron shaped crystallites for K2NaAlF6:Mn4+, while irregularly shaped particles with a broad size distribution are obtained for K3AlF6:Mn4+ and Na3AlF6:Mn4+. All synthesized M3AlF6:Mn4+ phosphors show narrow red Mn4+ 2E → 4A2 luminescence that can be excited in the UV and blue spectral region. Luminescence spectra recorded at T = 4 K reveal that multiple Mn4+ sites are present in M3AlF6:Mn4+, which was not observed in previous reports. The presence of multiple Mn4+ sites is attributed to charge compensation that is required for Mn4+ on Al3+ sites. The results show that charge compensating defects have a large influence on the luminescence properties (e.g., spectra, QY, luminescence lifetime) of Mn4+-doped fluorides. Lowering of the QY by defect quenching is a problem for application of this class of Mn4+ phosphors. Finally, we investigated the thermal quenching behavior for M3AlF6:Mn4+. The luminescence quenching temperature of M3AlF6:Mn4+ is between 460 K and 490 K, which is above the phosphor operating temperature in high-power w-LEDs. If the QY can be improved by suppressing defect quenching, Mn4+-doped hexafluoroaluminates are a promising class of materials as their chemical and thermal stability is superior to the presently used commercial Mn4+ phosphors.

Supplementary Materials

The following are available online at www.mdpi.com/1996-1944/10/11/1322/s1. Figure S1: Emission spectra of (a) Na3AlF6:Mn4+ (0.4%) and (b) K2NaAlF6:Mn4+ (0.9%) at various temperatures between 298 and 573 K (λexc = 450 nm); Figure S2: XRD patterns of K2NaAlF6:Mn4+ (2.9%) for as-synthesized K2NaAlF6:Mn4+ phosphor and K2NaAlF6:Mn4+ phosphor that has been heated to 573 K. The XRD patterns are in agreement with the reference diffraction pattern for K2NaAlF6 (PDF 01-072-2434, red); Video S1: Movie of synthesis of K2NaAlF6:Mn4+ phosphor. The phosphor shows bright red Mn4+ luminescence under 405 nm illumination.

Acknowledgments

The authors thank Stephan Zevenhuizen for performing the SEM measurements. Mart Peeters is acknowledged for measuring the photoluminescence quantum yields. This work is financially supported by Technologiestichting STW, which is part of the Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO).

Author Contributions

Tim Senden and Andries Meijerink conceived and designed the experiments; Tim Senden and Robin G. Geitenbeek synthesized the phosphor particles; Tim Senden performed the characterization measurements; Tim Senden, Robin G. Geitenbeek and Andries Meijerink analyzed the data; Tim Senden and Andries Meijerink wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Charge of the LED brigade: A global switch to LEDs will change the lighting business. Economist. 2011. http://www.economist.com/node/21526373.
  2. Krames, M.R.; Shchekin, O.B.; Mueller-Mach, R.; Mueller, G.O.; Zhou, L.; Harbers, G.; Craford, M.G. Status and future of high-power light-emitting diodes for solid-state lighting. IEEE/OSA J. Disp. Technol. 2007, 3, 160–175. [Google Scholar] [CrossRef]
  3. Harbers, G.; Bierhuizen, S.J.; Krames, M.R. Performance of high power light emitting diodes in display illumination applications. IEEE/OSA J. Disp. Technol. 2007, 3, 98–109. [Google Scholar] [CrossRef]
  4. Nakamura, S. Current status of GaN-based solid-state lighting. MRS Bull. 2009, 34, 101–107. [Google Scholar] [CrossRef]
  5. 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]
  6. Setlur, A.A. Phosphors for LED-based solid-state lighting. Electrochem. Soc. Interface 2009, 18, 32–36. [Google Scholar]
  7. Xie, R.J.; Hirosaki, N. Silicon-based oxynitride and nitride phosphors for white LEDs—A review. Sci. Technol. Adv. Mater. 2007, 8, 588–600. [Google Scholar] [CrossRef]
  8. Zhu, H.; Lin, C.C.; Luo, W.; Shu, S.; Liu, Z.; Liu, Y.; Kong, J.; Ma, E.; Cao, Y.; Liu, R.-S.; et al. Highly efficient non-rare-earth red emitting phosphor for warm white light-emitting diodes. Nat. Commun. 2014, 5, 4312. [Google Scholar] [CrossRef] [PubMed]
  9. Setlur, A.A.; Radkov, E.V.; Henderson, C.S.; Her, J.-H.; Srivastava, A.M.; Karkada, N.; Kishore, M.S.; Kumar, N.P.; Aesram, D.; Deshpande, A.; et al. Energy-efficient, high-color-rendering LED lamps using oxyfluoride and fluoride phosphors. Chem. Mater. 2010, 22, 4076–4082. [Google Scholar] [CrossRef]
  10. Sijbom, H.F.; Joos, J.J.; Martin, L.I.D.J.; Van den Eeckhout, K.; Poelman, D.; Smet, P.F. Luminescent behavior of the K2SiF6:Mn4+ red phosphor at high fluxes and at the microscopic level. ECS J. Solid State Sci. Technol. 2016, 5, R3040–R3048. [Google Scholar] [CrossRef] [Green Version]
  11. Lin, C.C.; Meijerink, A.; Liu, R.-S. Critical red components for next-generation white LEDs. J. Phys. Chem. Lett. 2016, 7, 495–503. [Google Scholar] [CrossRef] [PubMed]
  12. Nguyen, H.-D.; Liu, R.-S. Narrow-band red-emitting Mn4+-doped hexafluoride phosphors: Synthesis, optoelectronic properties, and applications in white light-emitting diodes. J. Mater. Chem. C 2016, 4, 10759–10775. [Google Scholar] [CrossRef]
  13. Hoshino, R.; Adachi, S. Photo-induced degradation and thermal decomposition in ZnSnF6·6H2O:Mn4+ red-emitting phosphor. Opt. Mater. 2015, 48, 36–43. [Google Scholar] [CrossRef]
  14. Sijbom, H.F.; Verstraete, R.; Joos, J.J.; Poelman, D.; Smet, P.F. K2SiF6:Mn4+ as a red phosphor for displays and warm-white LEDs: A review of properties and perspectives. Opt. Mater. Express 2017, 7, 3332–3365. [Google Scholar] [CrossRef]
  15. Murphy, J.E.; Garcia-Santamaria, F.; Setlur, A.A.; Sista, S. PFS, K2SiF6:Mn4+: The red-line emitting LED phosphor behind GE’s TriGain TechnologyTM platform. SID Symp. Dig. Tech. Pap. 2015, 46, 927–930. [Google Scholar] [CrossRef]
  16. Nguyen, H.-D.; Lin, C.C.; Liu, R.-S. Waterproof alkyl phosphate coated fluoride phosphors for optoelectronic materials. Angew. Chem. Int. Ed. 2015, 54, 10862–10866. [Google Scholar] [CrossRef] [PubMed]
  17. Song, E.H.; Wang, J.Q.; Ye, S.; Jiang, X.F.; Peng, M.Y.; Zhang, Q.Y. Room-temperature synthesis and warm-white LED applications of Mn4+ ion doped fluoroaluminate red phosphor Na3AlF6:Mn4+. J. Mater. Chem. C 2016, 4, 2480–2487. [Google Scholar] [CrossRef]
  18. Song, E.; Wang, J.; Shi, J.; Deng, T.; Ye, S.; Peng, M.; Wang, J.; Wondraczek, L.; Zhang, Q. Highly efficient and thermally stable K3AlF6:Mn4+ as a red phosphor for ultra-high-performance warm white light-emitting diodes. ACS Appl. Mater. Interfaces 2017, 9, 8805–8812. [Google Scholar] [CrossRef] [PubMed]
  19. Zhu, Y.; Cao, L.; Brik, M.G.; Zhang, X.; Huang, L.; Xuan, T.; Wang, J. Facile synthesis, morphology and photoluminescence of a novel red fluoride nanophosphor K2NaAlF6:Mn4+. J. Mater. Chem. C 2017, 5, 6420–6426. [Google Scholar] [CrossRef]
  20. Zhu, Y.; Huang, L.; Zou, R.; Zhang, J.; Yu, J.; Wu, M.; Wang, J.; Su, Q. Hydrothermal synthesis, morphology and photoluminescent properties of an Mn4+-doped novel red fluoride phosphor elpasolite K2LiAlF6. J. Mater. Chem. C 2016, 4, 5690–5695. [Google Scholar] [CrossRef]
  21. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  22. Thompson, W.T.; Goad, D.G.W. Some thermodynamic properties of K3AlF6–KAlF4 melts. Can. J. Chem. 1976, 54, 3342–3349. [Google Scholar] [CrossRef]
  23. Rumble, J.R. Handbook of Chemistry and Physics, 98th ed.; CRC Press: Boca Raton, FL, USA; Taylor & Francis: Boca Raton, FL, USA, 2017. [Google Scholar]
  24. Lv, L.; Jiang, X.; Huang, S.; Chen, X.; Pan, Y. The formation mechanism, improved photoluminescence and LED applications of red phosphor K2SiF6:Mn4+. J. Mater. Chem. C 2014, 2, 3879–3884. [Google Scholar] [CrossRef]
  25. Setlur, A.A.; Lyons, R.J.; Nammalwar, P.K.; Hanumanlha, R.; Murphy, J.E.; Garcia, F. Moisture-Resistant Phosphor Compositions and Associate Methods. World Patent WO 2,015,102,876, 9 July 2015. [Google Scholar]
  26. Shriver, D.F.; Atkins, P.W. Inorganic Chemistry, 4th ed.; Oxford University Press: Oxford, UK, 2006. [Google Scholar]
  27. Bode, H.; Jenssen, H.; Bandte, F. Über eine neue darstellung des kalium-hexafluoromanganats (IV). Angew. Chem. 1953, 65, 304. [Google Scholar] [CrossRef]
  28. Roesky, H.W. Efficient Preparations of Fluorine Compounds; Roesky, H.W., Ed.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2012; ISBN 9781118409466. [Google Scholar]
  29. Morss, L.R. Crystal structure of dipotassium sodium fluoroaluminate (elpasolite). J. Inorg. Nucl. Chem. 1974, 36, 3876–3878. [Google Scholar] [CrossRef]
  30. Hawthorne, F.C.; Ferguson, R.B. Refinement of the crystal structure of cryolite. Can. Mineral. 1975, 13, 377–382. [Google Scholar]
  31. Abakumov, A.M.; King, G.; Laurinavichute, V.K.; Rozova, M.G.; Woodward, P.M.; Antipov, E.V. The crystal structure of α-K3AlF6: Elpasolites and double perovskites with broken corner-sharing connectivity of the octahedral framework. Inorg. Chem. 2009, 48, 9336–9344. [Google Scholar] [CrossRef] [PubMed]
  32. Adachi, S.; Takahashi, T. Photoluminescent properties of K2GeF6:Mn4+ red phosphor synthesized from aqueous HF/KMnO4 solution. J. Appl. Phys. 2009, 106, 013516. [Google Scholar] [CrossRef]
  33. Xu, Y.K.; Adachi, S. Properties of Na2SiF6:Mn4+ and Na2GeF6:Mn4+ red phosphors synthesized by wet chemical etching. J. Appl. Phys. 2009, 105, 013525. [Google Scholar] [CrossRef]
  34. Jiang, X.; Pan, Y.; Huang, S.; Chen, X.; Wang, J.; Liu, G. Hydrothermal synthesis and photoluminescence properties of red phosphor BaSiF6:Mn4+ for LED applications. J. Mater. Chem. C 2014, 2, 2301–2306. [Google Scholar] [CrossRef]
  35. Fang, M.-H.; Nguyen, H.-D.; Lin, C.C.; Liu, R. Preparation of a novel red Rb2SiF6:Mn4+ phosphor with high thermal stability through a simple one-step approach. J. Mater. Chem. C 2015, 3, 7277–7280. [Google Scholar] [CrossRef]
  36. Zhou, Q.; Tan, H.; Zhou, Y.; Zhang, Q.; Wang, Z.; Yan, J.; Wu, M. Optical performance of Mn4+ in a new hexa-coordinative fluorozirconate complex of Cs2ZrF6. J. Mater. Chem. C 2016, 4, 7443–7448. [Google Scholar] [CrossRef]
  37. Wang, Z.; Zhou, Y.; Yang, Z.; Liu, Y.; Yang, H.; Tan, H.; Zhang, Q.; Zhou, Q. Synthesis of K2XF6:Mn4+ (X = Ti, Si and Ge) red phosphors for white LED applications with low-concentration of HF. Opt. Mater. 2015, 49, 235–240. [Google Scholar] [CrossRef]
  38. De Mello Donegá, C. Synthesis and properties of colloidal heteronanocrystals. Chem. Soc. Rev. 2011, 40, 1512–1546. [Google Scholar] [CrossRef] [PubMed]
  39. Paulusz, A.G. Efficient Mn(IV) emission in fluorine coordination. J. Electrochem. Soc. 1973, 120, 942–947. [Google Scholar] [CrossRef]
  40. Takahashi, T.; Adachi, S. Mn4+-activated red photoluminescence in K2SiF6 phosphor. J. Electrochem. Soc. 2008, 155, E183–E188. [Google Scholar] [CrossRef]
  41. Senden, T.; Broers, F.T.H.; Meijerink, A. Comparative study of the Mn4+ 2E → 4A2 luminescence in isostructural RE2Sn2O7:Mn4+ pyrochlores (RE3+ = Y3+, Lu3+ or Gd3+). Opt. Mater. 2016, 60, 431–437. [Google Scholar] [CrossRef]
  42. Blasse, G.; Grabmaier, B.C. Luminescent Materials; Springer: Berlin, Germany, 1994. [Google Scholar]
  43. Garcia-Santamaria, F.; Murphy, J.E.; Setlur, A.A.; Sista, S.P. Concentration quenching in K2SiF6:Mn4+ phosphors. ECS J. Solid State Sci. Technol. 2018, 7, R3030–R3033. [Google Scholar] [CrossRef]
  44. Chodos, S.L.; Black, A.M.; Flint, C.D. Vibronic spectra and lattice dynamics of Cs2MnF6 and A12MIVF6:MnF62−. J. Chem. Phys. 1976, 65, 4816–4824. [Google Scholar] [CrossRef]
  45. Brik, M.G.; Srivastava, A.M. On the optical properties of the Mn4+ ion in solids. J. Lumin. 2013, 133, 69–72. [Google Scholar] [CrossRef]
  46. Brik, M.G.; Camardello, S.J.; Srivastava, A.M. Influence of covalency on the Mn4+ 2Eg4A2g emission energy in crystals. ECS J. Solid State Sci. Technol. 2014, 4, R39–R43. [Google Scholar] [CrossRef]
  47. Sakurai, S.; Nakamura, T.; Adachi, S. Rb2SiF6:Mn4+ and Rb2TiF6:Mn4+ red-emitting phosphors. ECS J. Solid State Sci. Technol. 2016, 5, R206–R210. [Google Scholar] [CrossRef]
  48. Bachmann, V.; Ronda, C.; Meijerink, A. Temperature quenching of yellow Ce3+ luminescence in YAG:Ce. Chem. Mater. 2009, 21, 2077–2084. [Google Scholar] [CrossRef]
  49. Senden, T.; van Dijk-Moes, R.J.A.; Meijerink, A. Quenching of the red Mn4+ luminescence in Mn4+-doped fluoride LED phosphors. ECS J. Solid State Sci. Technol. 2017. Unpublished work. [Google Scholar]
Figure 1. Powder X-ray diffraction (XRD) patterns of M3AlF6:Mn4+ (M = Na, K). The XRD patterns of the synthesized phosphors are in excellent agreement with the reference patterns for Na3AlF6 (PDF 00-025-0772, red), K2NaAlF6 (PDF 01-072-2434, blue) and K3AlF6 (PDF 00-057-0227, green).
Figure 1. Powder X-ray diffraction (XRD) patterns of M3AlF6:Mn4+ (M = Na, K). The XRD patterns of the synthesized phosphors are in excellent agreement with the reference patterns for Na3AlF6 (PDF 00-025-0772, red), K2NaAlF6 (PDF 01-072-2434, blue) and K3AlF6 (PDF 00-057-0227, green).
Materials 10 01322 g001
Figure 2. Representative scanning electron microscopy (SEM) images of (a) Na3AlF6:Mn4+ (0.4%) phosphor particles; (b) K3AlF6:Mn4+ (1.2%) phosphor particles and (c,d) K2NaAlF6:Mn4+ (0.9%) phosphor particles.
Figure 2. Representative scanning electron microscopy (SEM) images of (a) Na3AlF6:Mn4+ (0.4%) phosphor particles; (b) K3AlF6:Mn4+ (1.2%) phosphor particles and (c,d) K2NaAlF6:Mn4+ (0.9%) phosphor particles.
Materials 10 01322 g002
Figure 3. Room-temperature luminescence spectra of M3AlF6:Mn4+ (M = Na, K) phosphors. (a) Excitation spectra of K3AlF6:Mn4+ (1.2%) (red, λem = 628 nm), Na3AlF6:Mn4+ (0.4%) (green, λem = 628 nm) and K2NaAlF6:Mn4+ (0.9%) (blue, λem = 631 nm). The broad excitation bands correspond to the Mn4+ 4A24T1 and 4A24T2 transitions; (b) Emission spectra (λexc = 460 nm) of K3AlF6:Mn4+ (1.2%) (red), Na3AlF6:Mn4+ (0.4%) (green) and K2NaAlF6:Mn4+ (0.9%) (blue). The Mn4+emission spectra consist of zero-phonon (ZPL) and (anti-)Stokes vibronic (νi) 2E → 4A2 emission lines.
Figure 3. Room-temperature luminescence spectra of M3AlF6:Mn4+ (M = Na, K) phosphors. (a) Excitation spectra of K3AlF6:Mn4+ (1.2%) (red, λem = 628 nm), Na3AlF6:Mn4+ (0.4%) (green, λem = 628 nm) and K2NaAlF6:Mn4+ (0.9%) (blue, λem = 631 nm). The broad excitation bands correspond to the Mn4+ 4A24T1 and 4A24T2 transitions; (b) Emission spectra (λexc = 460 nm) of K3AlF6:Mn4+ (1.2%) (red), Na3AlF6:Mn4+ (0.4%) (green) and K2NaAlF6:Mn4+ (0.9%) (blue). The Mn4+emission spectra consist of zero-phonon (ZPL) and (anti-)Stokes vibronic (νi) 2E → 4A2 emission lines.
Materials 10 01322 g003
Figure 4. Emission spectra of K3AlF6:Mn4+ (1.2%) (red), Na3AlF6:Mn4+ (0.4%) (green) and K2NaAlF6:Mn4+ (0.9%) (blue) at T = 4 K. The excitation wavelength is 460 nm. Labels are assigned to the highest-energy zero-phonon line (ZPL) and vibronic 2E → 4A2 emissions (ν3, ν4 and ν6). The stars mark lines assigned to ZPLs of other Mn4+ sites (see text).
Figure 4. Emission spectra of K3AlF6:Mn4+ (1.2%) (red), Na3AlF6:Mn4+ (0.4%) (green) and K2NaAlF6:Mn4+ (0.9%) (blue) at T = 4 K. The excitation wavelength is 460 nm. Labels are assigned to the highest-energy zero-phonon line (ZPL) and vibronic 2E → 4A2 emissions (ν3, ν4 and ν6). The stars mark lines assigned to ZPLs of other Mn4+ sites (see text).
Materials 10 01322 g004
Figure 5. Emission and excitation spectra of K2NaAlF6:Mn4+ (0.9%) at T = 4 K. (a) Excitation spectra for λem = 622 nm (red), 625 nm (green), 630 nm (blue) and 637 nm (orange); (b) Emission spectra for λexc = 460 nm (red), 462 nm (green), 467 nm (blue) and 485 nm (orange).
Figure 5. Emission and excitation spectra of K2NaAlF6:Mn4+ (0.9%) at T = 4 K. (a) Excitation spectra for λem = 622 nm (red), 625 nm (green), 630 nm (blue) and 637 nm (orange); (b) Emission spectra for λexc = 460 nm (red), 462 nm (green), 467 nm (blue) and 485 nm (orange).
Materials 10 01322 g005
Figure 6. (a) Emission spectra of K3AlF6:Mn4+ (1.2%) at various temperatures between 298 K and 573 K (λexc = 450 nm); (b) Integrated PL intensity of K3AlF6:Mn4+ (1.2%) (blue), Na3AlF6:Mn4+ (0.4%) (green) and K2NaAlF6:Mn4+ (0.9%) (red) as a function of temperature between 300 and 600 K. The integrated PL intensity IPL is scaled to the integrated PL intensity at room temperature IRT.
Figure 6. (a) Emission spectra of K3AlF6:Mn4+ (1.2%) at various temperatures between 298 K and 573 K (λexc = 450 nm); (b) Integrated PL intensity of K3AlF6:Mn4+ (1.2%) (blue), Na3AlF6:Mn4+ (0.4%) (green) and K2NaAlF6:Mn4+ (0.9%) (red) as a function of temperature between 300 and 600 K. The integrated PL intensity IPL is scaled to the integrated PL intensity at room temperature IRT.
Materials 10 01322 g006
Figure 7. (a) Room-temperature emission spectra (λexc = 450 nm) of K2NaAlF6:Mn4+ (0.9%) for as-synthesized K2NaAlF6:Mn4+ phosphor (blue) and K2NaAlF6:Mn4+ phosphor that has been heated to 573 K (red); (b) Emission spectra of K2NaAlF6:Mn4+ (0.9%) at T = 4 K (λexc = 460 nm) for as-synthesized K2NaAlF6:Mn4+ phosphor (blue) and K2NaAlF6:Mn4+ phosphor that has been heated to 573 K (red).
Figure 7. (a) Room-temperature emission spectra (λexc = 450 nm) of K2NaAlF6:Mn4+ (0.9%) for as-synthesized K2NaAlF6:Mn4+ phosphor (blue) and K2NaAlF6:Mn4+ phosphor that has been heated to 573 K (red); (b) Emission spectra of K2NaAlF6:Mn4+ (0.9%) at T = 4 K (λexc = 460 nm) for as-synthesized K2NaAlF6:Mn4+ phosphor (blue) and K2NaAlF6:Mn4+ phosphor that has been heated to 573 K (red).
Materials 10 01322 g007
Table 1. Amounts of Na2CO3, K2CO3, KF and 40% HF (aq) used in the synthesis of M3AlF6:Mn4+ (M = Na, K).
Table 1. Amounts of Na2CO3, K2CO3, KF and 40% HF (aq) used in the synthesis of M3AlF6:Mn4+ (M = Na, K).
PhosphorNa2CO3K2CO3KF40% HF
Na3AlF6:Mn4+15 mmol--15 mL
K2NaAlF6:Mn4+5 mmol10 mmol-15 mL
K3AlF6:Mn4+--40 mmol 13 mL
1 A 4:1 ratio of K:Al was used, as this gave K3AlF6 without impurity phases. With a 3:1 ratio, the obtained phosphor contained impurities of other crystal phases.
Table 2. Space group, Al3+ site symmetry and average Al–F distance for M3AlF6 (M = Na, K); zero-phonon line (ZPL) energy of the Mn4+ 2E → 4A2 emission in M3AlF6:Mn4+. Structural data obtained from Refs. [29,30,31].
Table 2. Space group, Al3+ site symmetry and average Al–F distance for M3AlF6 (M = Na, K); zero-phonon line (ZPL) energy of the Mn4+ 2E → 4A2 emission in M3AlF6:Mn4+. Structural data obtained from Refs. [29,30,31].
LatticeSpace GroupAl3+ Symmetry Al–F Distance (Å)ZPL Energy (cm−1)
Na3AlF6P21/nCi1.80816,167
K2NaAlF6Fm 3 ¯ mOh1.77816,082
K3AlF6I41/aC11.81016,200

Share and Cite

MDPI and ACS Style

Senden, T.; Geitenbeek, R.G.; Meijerink, A. Co-Precipitation Synthesis and Optical Properties of Mn4+-Doped Hexafluoroaluminate w-LED Phosphors. Materials 2017, 10, 1322. https://doi.org/10.3390/ma10111322

AMA Style

Senden T, Geitenbeek RG, Meijerink A. Co-Precipitation Synthesis and Optical Properties of Mn4+-Doped Hexafluoroaluminate w-LED Phosphors. Materials. 2017; 10(11):1322. https://doi.org/10.3390/ma10111322

Chicago/Turabian Style

Senden, Tim, Robin G. Geitenbeek, and Andries Meijerink. 2017. "Co-Precipitation Synthesis and Optical Properties of Mn4+-Doped Hexafluoroaluminate w-LED Phosphors" Materials 10, no. 11: 1322. https://doi.org/10.3390/ma10111322

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