Next Article in Journal
Multi-Detection Size Exclusion Chromatography as an Advanced Tool for Monitoring Enzyme–Antibody Conjugation Reaction and Quality Control of a Final Product
Next Article in Special Issue
A Novel Dye-Modified Metal–Organic Framework as a Bifunctional Fluorescent Probe for Visual Sensing for Styrene and Temperature
Previous Article in Journal
Wheat Roll Enhanced by Buckwheat Hull, a New Functional Food: Focus on the Retention of Bioactive Compounds
Previous Article in Special Issue
In Situ Filling of the Oxygen Vacancies with Dual Heteroatoms in Co3O4 for Efficient Overall Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Mn4+-Activated K2Nb1−xMoxF7 (0 ≤ x ≤ 0.15) Solid Solution Red Phosphors with Superior Moisture Resistance and Good Thermal Stability

College of Chemistry and Materials Science, Sichuan Normal University, Chengdu 610066, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(11), 4566; https://doi.org/10.3390/molecules28114566
Submission received: 13 April 2023 / Revised: 27 May 2023 / Accepted: 31 May 2023 / Published: 5 June 2023
(This article belongs to the Collection Green Energy and Environmental Materials)

Abstract

:
Nowadays, Mn4+-activated fluoride red phosphors with excellent luminescence properties have triggered tremendous attentions for enhancing the performance of white light-emitting diodes (WLEDs). Nonetheless, the poor moisture resistance of these phosphors impedes their commercialization. Herein, we proposed the dual strategies of “solid solution design” and “charge compensation” to design K2Nb1−xMoxF7 novel fluoride solid solution system, and synthesized the Mn4+-activated K2Nb1−xMoxF7 (0 ≤ x ≤ 0.15, x represents the mol % of Mo6+ in the initial solution) red phosphors via co-precipitation method. The doping of Mo6+ not only significantly improve the moisture resistance of the K2NbF7: Mn4+ phosphor without any passivation and surface coating, but also effectively enhance the luminescence properties and thermal stability. In particular, the obtained K2Nb1−xMoxF7: Mn4+ (x = 0.05) phosphor possesses the quantum yield of 47.22% and retains 69.95% of its initial emission intensity at 353 K. Notably, the normalized intensity of the red emission peak (627 nm) for the K2Nb1−xMoxF7: Mn4+ (x = 0.05) phosphor is 86.37% of its initial intensity after immersion for 1440 min, prominently higher than that of the K2NbF7: Mn4+ phosphor. Moreover, a high-performance WLED with high CRI of 88 and low CCT of 3979 K is fabricated by combining blue chip (InGaN), yellow phosphor (Y3Al5O12: Ce3+) and the K2Nb1−xMoxF7: Mn4+ (x = 0.05) red phosphor. Our findings convincingly demonstrate that the K2Nb1−xMoxF7: Mn4+ phosphors have a good practical application in WLEDs.

1. Introduction

White light-emitting diodes (WLEDs) is regarded as the green lighting or the fourth generation solid-state lighting on account of brilliant efficiency, long operation lifetime, high reliability, small size, environmentally friendly, and low power consumption [1,2,3,4,5,6,7,8,9]. Recently, WLEDs have been zealously explored as the main light source for display and illumination in numerous areas including auto mobile, liquid crystal displays (LCDs), medical treatment, agriculture and so on [10,11,12,13,14,15]. Currently, the most commercially popular approach for fabricating WLEDs is an integration of blue diode chips (InGaN) and yellow Y3Al5O12: Ce3+ (YAG: Ce3+) phosphor [16,17,18]. Nevertheless, YAG: Ce3+-based WLEDs faces two severe shortcomings of high correlated color temperature (CCT > 4500 K) and low color rendering index (CRI < 80), which is the result of notoriously insufficient red contribution, thus restricting their wider commercial application [1,10,19,20]. Interestingly, adding suitable red phosphor to supplement the scarcity of red emission could overcome the bottleneck mentioned above [1]. As a typical non-rare earth activator of the red emission, Mn4+ ion has received vast attention due to its characteristics of wide absorption, strong and sharp emission, low reabsorption, easily available and low cost. At the same time, Mn4+-activated oxides and fluorides are developing into the two categories of red phosphors with a great prospect [21]. Since the covalent bond of Mn4+‒O2− in oxide is stronger distinctly than that of Mn4+‒F in fluoride, so the emission wavelength of Mn4+ in oxide is too long to be sensitive to human eyes. Furthermore, the strongest excitation wavelength of Mn4+-activated fluoride is largely matching the emission peak of blue light LED chip and it is easily perceived by the human eye because the emission peak falls in the red light region, together with good chemical stability, high efficiency and simple synthesis method, making it be an ideal red phosphor for WLEDs [20].
To our knowledge, broad absorption and narrow-band emission at about 630 nm is the remarkable characteristics of Mn4+-activated fluoride red phosphors owing to particular 3d3 outer electron configuration of Mn4+ [22,23,24,25]. According to the recent years reports, the Mn4+-activated fluoride can be divided into three types:(1) A3X(Ⅲ)F6: Mn4+ (A = Li, Na, K, Cs; X = Al, Ga, Sc), A2BX(Ⅲ)F6: Mn4+ (A/B = Li, Na, K, Rb, Cs; X = Al, Ga, Sc) [26,27,28]; (2) A2M(Ⅳ)F6: Mn4+ (A = NH4, Li, Na, K, Rb, Cs; M = Si, Ti, Ge, Sn, Zr, Hf), BM(Ⅳ)F6: Mn4+ (B = Ba, Zn; M = Si, Ge, Ti), AA′M(Ⅳ)F6: Mn4+ (A/A′ = Li, Na, K; M = Si, Ti, Ge) [29,30,31,32,33,34,35,36,37,38,39]; (3) A2X(Ⅴ)F7: Mn4+ (A = Li, Na, K, Rb, Cs; X = Nb, Ta) [40,41,42]. Unfortunately, the first two compounds possess symmetrical [MnF6]2− and thereby lack the special spectral phenomenon of the zero phonon line (ZPL), which is crucial to advance the color purity quality of WLEDs [40]. Remarkably, in C1 (or C2v) group symmetry, K2NbF7: Mn4+ undergoes highly distorted octahedral environment, making the distinctive vibronic transitions and intense zero phonon line appear [40]. Substantially, it may dominate the emission of Mn4+, in favor of the 2Eg4A2g spin-forbidden transition of Mn4+ and the enhancement of the color purity [43]. Hence, Mn4+-activated K2NbF7 phosphors have evoked plentiful attention. However, K2NbF7: Mn4+ phosphors can be easily hydrolyzed into manganese oxides and hydroxides in damp condition due to the inherent poor moisture resistance when [MnF6]2− complex ions locate on the phosphor surface, leading to a reduction in the intensity of the red emission and the performance deterioration of the WLEDs during long-term operation. Therefore, it is very important and valuable to improve the moisture resistance of Mn4+-activated K2NbF7 red phosphors [44,45,46,47].
To solve these problems, we propose the dual strategies of “solid solution design” and “charge compensation” to design novel Mn4+-activated fluoride luminescent material, and the as-designed K2Nb1−xMoxF7: Mn4+ red phosphors were synthesized via co-precipitation method. It should be noted that the replacement of Nb5+ by Mn4+ will probably produce an anion vacancy (fluorine vacancy) in the matrix, while the substitution of Nb5+ with Mo6+ will generate a cation vacancy and plays a role of charge compensation [42]. After doping of Mo6+, the structural diversity and structural rigidity enhancement of the K2NbF7: Mn4+ can be achieved simultaneously, thus effectively enhancing the comprehensive luminescence properties of the Mn4+-activated fluoride phosphors. Our results clearly demonstrated that not only the moisture resistance but also the thermal stability and luminescence performance are immensely enhanced by doping Mo6+. This work can provide an effective and facile strategy to design and synthesize the novel Mn4+-activated fluoride phosphors with superior luminescence properties, moisture resistance and good thermal stability, it also exploits new opportunities for these fluoride red phosphors used in WLEDs in outdoor high humidity working conditions.

2. Results and Discussion

2.1. Phase Structure, Morphology and Composition

Figure 1a exhibits the XRD patterns of the prepared K2Nb1−xMoxF7: Mn4+ samples. Apparently, the diffraction peaks of all the samples are very consistent with the standard card of K2NbF7 (PDF#22-0839) and no impurity phase appearances, indicating that the obtained K2Nb1−xMoxF7: Mn4+ phosphors are pure solid solution. Since Nb5+ (69 pm, CN = 7), Mo6+ (73 pm, CN = 7) and Mn4+ (53 pm, CN = 6) have similar ionic radii, they are able to occupy the same lattice site in the K2MF7 fluoride. It also can be seen that the strongest peak of the samples located at 2θ of ~17.5° gradually shift to the lower 2θ side with the increasing x (see in the enlarged patterns), further confirming that the replacement of Nb5+ by Mo6+ in K2NbF7 host and the formation of K2Nb1−xMoxF7: Mn4+ solid solution. The crystal structure of K2Nb1−xMoxF7: Mn4+ solid solution is presented in Figure 1b. It can be intuitively seen that there are seven F around Nb5+ and Mo6+, forming [NbF7]2− decahedron and [MoF7] decahedron, respectively. Therefore, the defect equations of Mn4+ and Mo6+ doped into K2NbF7 host can be expressed as follows [42,48]:
K 2 M n F 6 K 2 N b F 7 K K × + M n N b / + 6 F F × + V F
K M o F 7 K 2 N b F 7 K K × + M o N b + 7 F F × + V K /
As shown in Equations (1) and (2), the replacement of Nb5+ by Mn4+ will probably produce anion vacancy (i.e., F ion vacancy), while the substitution of Nb5+ with Mo6+ will probably produce cation vacancy (i.e., K+ vacancy) in K2NbF7 host material. Cation vacancy can also compensate the charge of F vacancy, which is beneficial to stabilize the structure of the obtained solid solution.
The percentage difference of radius (Dr) between doped ion Mo6+ and replacement ion Nb5+ calculated by Formula (1) is 5%, which is less than 15%, indicating that Nb5+ and Mo6+ can form K2Nb1−xMoxF7: Mn4+ continuous solid solution.
Dr = R 1 ( CN ) R 2 ( CN ) R 1 ( CN ) × 100 %
where R1 and R2 are radii of the substituted ion and matrix ion, respectively, and CN represents the coordination number.
To further explore the effect of Mo6+-doping concentration on the phase purity and cell parameters of the K2Nb1−xMoxF7: Mn4+ samples, the Rietveld refinement was carried out based on the recorded XRD data of all samples by using GSAS software (GSAS_EXPGUI). The detailed refinement results of the K2Nb1−xMoxF7: Mn4+ samples are provided in Figure S1 and Table 1, respectively. As shown in Figure S1, the original XRD data (recorded data) of all the samples could well match with the corresponding calculated results based on the GSAS software, suggesting that these samples possess high phase purity. All of the K2Nb1−xMoxF7: Mn4+ samples are monoclinic structure with the same space group of P121/c1. In particular, the unit cell volume of the K2Nb1−xMoxF7: Mn4+ samples gradually increased from 631.624 Å3 to 632.067 Å3 as Mo6+-doping concentration (x) increasing from 0 to 0.15, this should be attributed the cell expansion induced by the replacement of Nb5+ by Mo6+ with the larger ion radius. The obtained relatively lower reliability factors including Rp, Rwp and χ2 for all the K2Nb1−xMoxF7: Mn4+ samples indicate that the obtained refinement results are credible. The Rietveld refinement results once again confirm that the doped Mo6+ ions successfully replaced Nb5+ ions in the K2NbF7 host and formed K2Nb1−xMoxF7: Mn4+ solid solution.
To investigate the influence of Mo6+-doping on the surface microstructure of K2Nb1−xMoxF7: Mn4+ samples, as the examples, Figure 2 gives the SEM images of K2NbF7: Mn4+ and K2Nb1−xMoxF7: Mn4+ with x = 0.05 and 0.15, respectively. It can be observed that all the samples are mainly rod-like microcrystals with the length between 10–70 μm and the width between 2–5 μm, representing excellent crystallinity. Additionally, all the samples have similar uniformity and regularity, suggesting that the doping concentration of Mo6+ has little influence on the surface microstructure of the K2Nb1−xMoxF7: Mn4+ solid solution crystals. Reasonably, the influence of the surface microstructure (morphology and size) on the luminescence properties of the K2Nb1−xMoxF7: Mn4+ solid solution phosphors can be ignored in the subsequent study.
The actual mole percentages of the doped Mo6+ and Mn4+ ions in K2Nb1−xMoxF7: Mn4+ solid solution crystals were measured by ICP-MS, and the results are depicted in Table 2. As can be seen, the Mo6+-doping concentration in these phosphors gradually increases with the increasing concentration of Mo6+ in the starting solution, although it is much lower than that of the starting solution. It should be noted that the actual Mn4+-doping concentration in the phosphors is between 1.19–1.48%, basically keeping in the narrow range of 1.35 ± 0.15 % . Thus, the influence of Mn4+-doping concentration on the luminescence properties of the K2Nb1−xMoxF7: Mn4+ phosphors also can be neglected in the following studies.

2.2. Luminescence Properties

The photoluminescence excitation spectra (PLE) of the K2Nb1−xMoxF7: Mn4+ phosphors were recorded under the monitoring wavelength of 627 nm, as illustrated in Figure 3a. Impressively, all samples have two wide absorption peaks in the wavelength range of 300–550 nm. The peak located in the near ultraviolet (~362 nm) district is attributed to the spin permissive transition of 4A2g4T1g of the 3d electrons of Mn4+ in the octahedron, whereas the peak located in blue (~470 nm) region is derived from the spin permissive transition of 4A2g4T2g of the 3d electrons of Mn4+ in the octahedron [49,50]. The K2Nb1−xMoxF7: Mn4+ phosphors have a strong absorption band in blue light region also means that these phosphors could be effectively excited by the InGaN LED blue light chip. Therefore, K2Nb1−xMoxF7: Mn4+ phosphors are promising red phosphors for WLEDs, which can be expected to improve CCT and Ra of the packaged WLEDs [19]. Figure 3b depicts the emission spectra (PL) of the K2Nb1−xMoxF7: Mn4+ phosphors in the range of 500–700 nm under excitation of 470 nm blue light. All the phosphors display the similar emission spectra, which are composed of the seven typical sharp emission peaks at 597 nm, 606 nm, 611 nm, 619 nm, 627 nm, 631 nm and 644 nm, respectively. These peaks are attributed to the anti-Stokes shift, ZPL and Stokes shift originating from 2Eg4A2g spin-forbidden transition of Mn4+, respectively [43,44]. Interestingly, the intensities of these emission peaks for the K2Nb1−xMoxF7: Mn4+ phosphors initially increase and then decease with the increasing Mo6+-doing concentration (x), giving the largest at x = 5%. It can be reasonably concluded that the best enhancement effect of luminescent properties induced by Mo6+-doping in K2NbF7: Mn4+ system should appear at x = 5% in our case (Figure 3b).
In general, the main factors that determine the luminescent properties of the phosphors include the doping concentration of activator, surface microstructures (morphology and size) and the local environment around the activator ions. In present study, the changes in the doping concentration of Mn4+ and the surface microstructures of the K2Nb1−xMoxF7: Mn4+ phosphors are almost negligible (see the relevant discussion above). Thus, the variation of luminescent properties in the K2Nb1−xMoxF7: Mn4+ phosphors should be attributed to the change of the local environment around Mn4+ in the lattices after being doped by Mo6+ ions. As mentioned above, the unit cell of the K2Nb1−xMoxF7: Mn4+ solid solution will expand owing to the larger ion radius of Mo6+ than that of the Nb5+. Hence, the substitution of Nb5+ by Mo6+ will inevitably lead to the variation of the sublattice of the K2NbF7: Mn4+, further changing the local environment around Mn4+, thus effectively breaking through the prohibition rule to improve d-d transition probability and enhancing red emission to a certain extent.
The UV-visible diffuse reflection spectra and band gaps of K2NbF7: Mn4+ (KNF) and K2Nb1−xMoxF7: Mn4+ with x = 0.05 (KNMF) are shown in Figure S2, and the corresponding basis for the band gap calculation is also provided in the Supplementary Materials. It can be seen that the two samples have strong absorption bands in both the near ultraviolet area and the blue light area, corresponding to the transitions of 4A2g4T1g and 4A2g4T2g of Mn4+, respectively. More importantly, the absorption band in the blue light area is stronger than the absorption band in the ultraviolet region, proving that the obtained phosphors can be well combined with the blue light InGaN chip. The obtained Eg value for KNF and KNMF are 4.62 eV and 4.38 eV, respectively. Apparently, the band gap of KNMF obviously decreased after Mo6+ being doped in KNF, indicating that the Mo6+-doping can effectively improve excitation efficiency of the K2NbF7: Mn4+ phosphors.
Since the charge number of Mo6+ is greater than that of Nb5+, after a certain amount of Nb5+ ions are replaced by Mo6+ ions, the crystal field strength of the [MoF7] complexation ion in the formed K2Nb1−xMoxF7: Mn4+ matrix is stronger than that of the [NbF7]2− complexation ion, thus effectively narrowing the band gap of the K2NbF7: Mn4+ phosphor. Therefore, the obtained K2Nb1−xMoxF7: Mn4+ solid solution phosphors have the lower photon energy required in the excitation luminescence process, leading to the higher excitation efficiency and the superior luminescence properties in comparison to K2NbF7: Mn4+.
To comprehensively understand the luminescent properties of the as-obtained phosphors, the quantum yields (η) of the KNF and KNMF were measured. On the basis of the PL spectra of the two samples, their quantum efficiencies (η) are calculated through the following formula:
η = L s E R E S
where LS, ES and ER are the emission spectrum of the phosphor sample, the spectra of excitation light with and without samples in the integral sphere, respectively [51], the quantum yield (η) of KNF and KNMF are calculated to be 39.65% and 47.22%, respectively (Figure S3). After doping a suitable concentration of Mo6+ ions, the quantum yield of K2Nb1−xMoxF7: Mn4+ solid solution phosphors also can be effectively enhanced compared with the K2NbF7: Mn4+ phosphor.
The decay curves of K2Nb1−xMoxF7: Mn4+ (x = 0, 0.05 and 0.15) under excitation of 470 nm blue light (monitored at 627 nm) are depicted in Figure 3d. All the decay curves can be well fitted in a double-exponential function below.
I ( t ) = I 0 + A 1 e ( t / τ 1 ) + A 2 ( t / τ 2 )
τ = ( A 1 τ 1 2 + A 2 τ 2 2 ) ( A 1 τ 1 + A 2 τ 2 )
where A1 and A2 are constant (represent the initial intensities of the τ1 and τ2, i.e., their relative weightings), It is the luminescence intensity at time t, I0 is the luminescence intensity at initial time, τ1 and τ2 represent the decay lifetime, and τ is the average decay time. Among them, τ1 is short lifetime, while τ2 is long lifetime. For the decay model by double-exponential fitting, the obtained fast and slow lifetimes can be understood as follows: Assuming that the long lifetime describes the intrinsic emission decay of all Mn4+ ions while the short lifetime represents the part of Mn4+ ions that have some defect centers nearby, and a corresponding efficient channel of energy transfer to quenching centers. The decay curves deviate from the single exponential model, indicating that an extra energy decay channel is active. This extra energy decay channel may have stemmed from the cation (such as K+) vacancy defects in the vicinity of Mn4+ ions for charge compensation. The fitting parameters of decay lifetimes for the K2Nb1−xMoxF7: Mn4+ phosphors are provided in Table S1. The calculated lifetime for the K2Nb1−xMoxF7: Mn4+ phosphors with x = 0, 0.05 and 0.15 is 2.1459 ms, 3.4917 ms and 3.1883 ms, respectively. Obviously, the decay lifetime of K2Nb1−xMoxF7: Mn4+ phosphors are significantly larger than that of the K2NbF7: Mn4+ phosphor, further suggesting that Mo6+-doping effectively boosted the PL emission intensity of the K2Nb1−xMoxF7: Mn4+ phosphors. Accordingly, the K2Nb1−xMoxF7: Mn4+ phosphor (x = 0.05) exhibits the longest decay lifetime could be attributed the optimum Mo6+-doping concentration in the K2Nb1−xMoxF7: Mn4+ solid solution in our case.
The corresponding Commission Internationale de L’Eclairage (CIE) coordinates of the K2Nb1−xMoxF7: Mn4+ phosphors are shown in Figure 3c and Table 3, the color purity and the correlated color temperature (CCT) of these phosphors were calculated (the relevant calculation basis and details are provided in Supplementary Materials), which are also listed in Table 3. The coordinates of these phosphors all fall in the redlight region and almost overlap, which are close to the coordinates specified by the National Television Standards Committee (NTSC) (x = 0.67, y = 0.33), manifesting again that K2Nb1−xMoxF7: Mn4+ samples are superior red phosphors. It is worth mentioning that the obtained K2Nb1−xMoxF7: Mn4+ samples have the higher color purity ~94% (above 90%) and the lower CCT of ~3500 K (below 4500 K), exhibiting the great application potential in warm WLEDs for indoor lighting.
Normally, WLEDs need to work at temperature much higher than room temperature, so the thermal stability of the phosphors is an important parameter to evaluate their actual application. As an example, the thermal stability of the K2Nb1−xMoxF7: Mn4+ with x = 0.05 (KNMF) was assessed, and the results are shown in Figure 4. Figure 4a depicts the temperature-dependent emission spectra of KNMF phosphor in the test temperature range of 303–423 K. It is clearly observed that the PL emission intensity of the phosphor gradually decreases with the increasing test temperature, this should be ascribed the thermal quenching effect stemming from the nonradiative transition [42,52]. Whereas the PL patterns do not vary with the increasing temperature, suggests that the luminescence mechanism of the K2Nb1−xMoxF7: Mn4+ phosphors has not changed in the test temperature range. Generally, the ratio of luminescence intensity of a phosphor at the test temperature of 353K (80 °C) to that at the initial temperature is used for evaluating its thermal stability. Figure 4b shows the normalized emission intensity of the red emission peak at 627 nm as the function of temperature for KNMF. Excitingly, the luminescence intensity of the KNMF retains 69.95% of the initial intensity at 353 K, showing the good thermal stability. For a better understanding of the thermal quenching characteristics for the K2Nb1−xMoxF7: Mn4+ (x = 0.05) (KNMF) phosphor, the activation energy Ea is estimated (the details are provided in Supplementary Materials) and given in Figure 4c. The calculated Ea is 0.74 eV, further verifying that the KNMF phosphor owns good thermal stability.
Table 4 also gives comparison of the thermal stability between K2Nb1−xMoxF7: Mn4+ (x = 0.05) (KNMF) phosphor with the reported some typical Mn4+-activated fluoride phosphors. Comparatively speaking, the as-obtained phosphor shows good thermal stability. The above results may be due to the formation of K2Nb1−xMoxF7: Mn4+ solid solution caused by Mo6+-doping and the simultaneous change of the crystal field local environment of Mn4+, which not only increases the structural rigidity of a phosphor, but also inhibits the non-radiation transition, thus effectively improving the thermal stability of the phosphors. Figure 4d displays the energy conversion process and mechanism of the thermal quenching of Mn4+ in K2Nb1−xMoxF7, which can better explain the thermal quenching effect. Under excitation of 470 nm blue light, the electrons of Mn4+ transition from the ground state 4A2g to the excited state 4T1g and 4T2g. These unstable electrons will then make a non-radiative transition to a lower intermediate state 2Eg. Afterwards, the electrons at 2Eg will return to their ground state by a radiative transition, emitting red light. Nevertheless, some electrons in 2Eg will assimilate the activation energy under the high temperature, return to the ground state through the intersection of 4T2g and 4A2g. This process is a non-radiative transition, reducing the possibility of normal radiative transition and resulting in the weakening of the intensity of the red emission of Mn4+, thus generating thermal quenching behavior.

2.3. Moisture Resistance Properties

As is well known, the practical application of Mn4+ activated fluoride red phosphors are greatly limited by their poor moisture resistance, thus the moisture resistance is an important performance parameter that should be considered for this kind of phosphors. The moisture resistance properties of K2NbF7: Mn4+ (KNF) and K2Nb1−xMoxF7: Mn4+ with x = 0.05 (KNMF) phosphors were measured and shown in Figure 5. The two phosphors were measured based on an extreme moisture resistance method–water immersing [20]. After immersing for various times (0–1440 min), the photographs of the KNF and KNMF red phosphors under the sunlight and blue light irradiation are revealed in Figure 5a,b, respectively, and normalized intensity of the strongest peak at 627 nm of the two phosphors are shown in Figure 5c. Under irradiation of the sunlight, the color of the KNF solution turned into dark purple after being immersed in deionized water for 1440 min, suggests that [MnF6]2− ions in KNF phosphor have been obviously hydrolyzed. In contrast, the KNMF solution is almost colorless under the same immersing time (1440 min), implying the superior water stability of [MnF6]2− ions in KNMF phosphor (Figure 5a). Moreover, the KNMF phosphor displays a stronger red light than that of the KNF phosphor under the excitation of blue light at 470 nm, intuitively confirming that the better moisture resistance of KNMF phosphors (Figure 5b). It can be seen that from Figure 5c, the red emission intensity (627 nm) of the two phosphors both gradually decrease with the increase of the immersing time, while the decline trend of the red emission intensity for the KNMF phosphor is obviously slower than that for KNF phosphor. More importantly, the red emission intensity of the KNMF phosphor can still maintain 86.37% of its initial level even after immersing for 1440 min, obviously higher than that of KNF phosphor at the same immersing time. These results confirmed that the moisture resistance of K2NbF7: Mn4+ has been obviously enhanced after doping a proper amount of Mo6+ ions. Compared with the previously reported Mn4+-activated fluorides, the K2Nb1−xMoxF7: Mn4+ (x = 0.05) (KNMF) phosphor show excellent water-proof properties (Table 5). The reason is that the formation of solid solutions inhibits the hydrolysis of [MnF6]2−, thus enhances the moisture resistance of the Mn4+-activated fluoride phosphors [43].

2.4. Performances of WLEDs

Given that K2Nb1−xMoxF7: Mn4+ phosphors exhibit excellent emission efficiency, good thermal stability, and superior moisture resistance, it is essential to assess their practical application in WLEDs. Three representative WLEDs were packaged: a blue chip (InGaN) with a yellow YAG: Ce3+ phosphor (A); a blue chip with a mixtures of YAG: Ce3+ and KNF (B); and a blue chip with a mixtures of YAG: Ce3+ and KNMF (C). The electroluminescence (EL) spectra (the driving current of 0.7A) and photographs of the WLEDs are exhibited in Figure 6a. It can be seen that the obtained two red phosphors (KNF and KNMF) are efficiently excited by the blue LED chip, resulting in a sharp emission peak in the range of 600–650 nm. In addition, the two WLEDs (B and C) packaged with a blue chip and mixtures of YAG: Ce3+ and the K2Nb1−xMoxF7: Mn4+ (x = 0 and 0.05) red phosphors display warm white light, while the WLED packaged with a blue chip (InGaN) and yellow YAG: Ce3+ phosphor shows cold white light (insets in Figure 6a). The corresponding chromaticity coordinates of these WLEDs are displayed in Figure 6b. Among them, the WLED packaged with a blue chip (InGaN) and yellow YAG: Ce3+ phosphor has the CIE color coordinates of (0.3174, 0.3463), a CCT of 6170 K, and the color rendering index (CRI) of 72, indicating that it is not suitable for indoor lighting due to the absence of a red-light component. After adding KNF phosphor as the red component, the packaged WLED possesses the CIE color coordinates of (0.3510, 0.3767), a high CRI of 83.7, and a low CCT of 4883 K. It is worth mentioning that the packaged WLED incorporating KNMF red phosphor has the CIE color coordinates of (0.3886, 0.4035), a higher CRI of 88, and a lower CCT of 3979 K. Compared with currently commercially used WLEDs, the obvious enhancement in CRI and the significant decrease in CCT for the packaged WLED using the synthesized K2Nb1−xMoxF7: Mn4+ phosphors as the red component can be simultaneously achieved. These results further confirm that the as-obtained K2Nb1−xMoxF7: Mn4+ phosphors are promising luminescent materials with potential application in warm WLEDs.

3. Materials and Methods

3.1. Materials

The raw materials KMnO4 (99.5%), HF (40 wt%), H2O2, and anhydrous ethanol were obtained from the Chengdu Kelon Chemical Reagent Factory. KHF2 (99%), Nb2O5 (99.5%), and H24Mo7N6O24·4H2O (99%) were purchased from Aladdin Biotechnology Co., Ltd. (Xi’an, China). H2O2 and anhydrous ethanol were analytical grade and required no further purification.

3.2. Synthesis of K2MnF6 Precursor

Firstly, 10 g of KMnO4 and 5 g of KHF2 were dissolved in 100 mL of HF (40 wt%) solution. Afterwards, H2O2 was slowly added drop by drop after vigorously stirring and dissolving for 30 min. The color of the solution gradually changed from dark purple to brownish red, and the golden precipitation of K2MnF6 was produced. Ultimately, the K2MnF6 powders were obtained after centrifuging at 4000 r/min for 1min, washing with anhydrous ethanol three times, and drying at 70 °C for 8 h.

3.3. Preparation of K2Nb1−xMoxF7: Mn4+

In a typical synthesis procedure, Nb2O5 and H24Mo7N6O24·4H2O with a preset stoichiometric ratio (keeping the total amount of Nb and Mo of 1.9 mmol) and 6 mL HF were added into the Teflon-lined autoclave. Subsequently, the autoclave was maintained at 120 °C for 30 min. After the autoclave was naturally cooled to room temperature, 0.1 mmol of the synthesized K2MnF6 was added into the resultant solution and vigorously stirred for 10 min. Then 4 mmol of KHF2 was put into the as-obtained solution and continuously stirred for 10 min. After that, 5mL ethanol was added to the above solution and stirred thoroughly for 20 min until the reaction stopped. The sediments were collected, washed with anhydrous ethanol three times, and dried at 80 °C for 12 h. Lastly, the K2Nb1−xMoxF7: Mn4+ red phosphors were obtained (Figure S4).

3.4. Characterization

The crystal structures of the as-synthesized products were characterized with an Xray powder diffractometer (Cu Kα radiation, λ = 1.5406 Å). The structure was optimized by the GSAS program. The morphologies and the actual doping concentration of Mn4+ in a solid solution crystal were obtained with a scanning electron microscope (SEM, Quanta 250, FEI, St. Leonards, NSW, Australia) and Agilent 8900 ICP-MS (Santa Clara, CA, USA), respectively. Photoluminescence spectra (PL), photoluminescence excitation spectra (PLE), and temperature-dependent luminescence properties were tested employing a Hitachi F-4600 Fluorescence analyzer (Tokyo, Japan) equipped with a 150 W Xe lamp as an excitation light source. The diffuse reflection spectra of the samples were tested with a spectrophotometer (UV-3600, Shimadzu, Tokyo, Japan) and BaSO4 was used as the reflective material. The quantum efficiency (QE) of the samples was measured by a fluorescence spectrometer with an integrating sphere (Horiba FluoroMax-4, Tokyo, Japan). The WLEDs were tested by a high-precision array spectrometer (HSP 6000) under a steady current of 700 mA.

4. Conclusions

Novel Mn4+-activated K2Nb1−xMoxF7 phosphors were designed and successfully synthesized through a co-precipitation method. Benefiting from the synergistic effect of “solid solution formation” and “charge compensation” induced by Mo6+ doping, the obtained K2Nb1−xMoxF7 phosphors exhibit obviously enhanced comprehensive luminescence properties, including PL intensity, quantum yield, thermal stability, and moisture resistance. Notably, the obtained K2Nb1−xMoxF7: Mn4+ (x = 5%) phosphor exhibits the optimal comprehensive luminescence properties among the K2Nb1−xMoxF7: Mn4+ series. Using the K2Nb1−xMoxF7: Mn4+ (x = 5%) phosphor as a red component, the packaged WLED possesses a high CRI of 88 and a low CCT of 3979 K. The as-developed K2Nb1−xMoxF7: Mn4+ phosphors significantly boost the CRI (from 72 to 88) and decrease the CCT (from 6170K to 3979K) when compared with the commercial WLED packaged with a blue chip (InGaN) and yellow YAG: Ce3+ phosphor, with attractive application prospectives in WLEDs for indoor lighting.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28114566/s1, Figure S1. XRD Rietveld refinement patterns of K2Nb1−xMoxF7: Mn4+ samples. Figure S2. Diffuse reflection spectra of K2NbF7: Mn4+ (KNF) (a) and K2Nb1−xMoxF7: Mn4+ with x = 0.05 (KNMF) (c); band gaps of KNF (b) and KNMF (d). Figure S3. Quantum yields of K2NbF7: Mn4+ (KNF) (a) and K2Nb1−xMoxF7: Mn4+ with x = 0.05 (KNMF) (b). Figure S4. Schematic diagram of the preparation process for K2MnF6 and K2Nb1−xMoxF7: Mn4+ phosphors. Table S1. Fitting parameters of decay lifetimes for the K2Nb1−xMoxF7: Mn4+ phosphors.

Author Contributions

Methodology, data curation, writing—original draft, Y.G.; methodology, data curation, L.F.; investigation, data curation, L.W.; investigation, data curation, J.Z.; investigation, data curation, F.R.; investigation, data curation, S.L.; methodology, writing—review and editing, Z.N.; resources, investigation, T.Z.; resources, investigation, X.W.; resources, investigation, X.L.; conceptualization, writing—review and editing, supervision, D.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Training Program of Innovation and Entrepreneurship for Undergraduates (202210636001), National Natural Science Foundation of China (62104164), and Natural Science Foundation of Sichuan Province (2022NSFSC2001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Dong, Q.; Guo, C.; He, L.; Lu, X.; Yin, J. Improving the moisture resistance and luminescent properties of K2TiF6: Mn4+ by coating with CaF2. Mater. Res. Bull. 2019, 115, 98–104. [Google Scholar] [CrossRef]
  2. Senden, T.; van Dijk-Moes, R.J.A.; Meijerink, A. Quenching of the red Mn4+ luminescence in Mn4+-doped fluoride LED phosphors. Light Sci. Appl. 2018, 7, 8. [Google Scholar] [CrossRef] [PubMed]
  3. Kim, Y.H.; Ha, J.; Im, W.B. Towards green synthesis of Mn4+-doped fluoride phosphors: A review. J. Mater. Res. Technol. 2021, 11, 181–195. [Google Scholar] [CrossRef]
  4. Wu, J.; Li, Z.; Luo, L.; Xiong, Y.; Jiang, L.; Guo, R.; Meng, L. A facile two-step synthesis of an efficient narrow-band red-emitting K2NbF7: Mn4+ phosphor for warm white LEDs and its thermal quenching behavior. J. Alloys Compd. 2021, 863, 158058. [Google Scholar] [CrossRef]
  5. Ming, H.; Liu, S.; Liu, L.; Peng, J.; Fu, J.; Du, F.; Ye, X. Highly regular, uniform K3ScF6: Mn4+ phosphors: Facile synthesis, microstructures, photoluminescence properties, and application in light-emitting diode Devices. ACS Appl. Mater. Interfaces 2018, 10, 19783–19795. [Google Scholar] [CrossRef]
  6. Cao, R.; Wang, W.; Zhang, J.; Jiang, S.; Chen, Z.; Li, W.; Yu, X. Synthesis and luminescence properties of Li2SnO3: Mn4+ red-emitting phosphor for solid-state lighting. J. Alloys Compd. 2017, 704, 124–130. [Google Scholar] [CrossRef]
  7. Deng, T.T.; Song, E.H.; Su, J.; Zhou, Y.Y.; Wang, L.Y.; Ye, S.; Zhang, Q.Y. Stable narrowband red emission in fluorotellurate KTeF5: Mn4+via Mn4+ noncentral-site occupation. J. Mater. Chem. C 2018, 6, 4418–4426. [Google Scholar] [CrossRef]
  8. Ji, H.; Huang, Z.; Xia, Z.; Molokeev, M.S.; Atuchin, V.V.; Fang, M.; Huang, S. New yellow-emitting whitlockite-type structure Sr1.75Ca1.25(PO4)2: Eu2+ phosphor for near-UV pumped white light-emitting devices. Inorg. Chem. 2014, 53, 5129–5135. [Google Scholar] [CrossRef]
  9. Lin, C.C.; Liu, R.-S. Advances in phosphors for light-emitting diodes. J. Phys. Chem. Lett. 2011, 2, 1268–1277. [Google Scholar] [CrossRef]
  10. Hong, F.; Xu, H.; Pang, G.; Liu, G.; Dong, X.; Yu, W. Moisture resistance, luminescence enhancement, energy transfer and tunable color of novel core-shell structure BaGeF6: Mn4+ phosphor. Chem. Eng. J. 2020, 390, 124579. [Google Scholar] [CrossRef]
  11. Xiang, J.; Chen, J.; Zhang, N.; Yao, H.; Guo, C. Far red and near infrared double-wavelength emitting phosphor Gd2ZnTiO6: Mn4+, Yb3+ for plant cultivation LEDs. Dye. Pigment. 2018, 154, 257–262. [Google Scholar] [CrossRef]
  12. Huang, X.; Guo, H. Finding a novel highly efficient Mn4+-activated Ca3La2W2O12 far-red emitting phosphor with excellent responsiveness to phytochrome PFR: Towards indoor plant cultivation application. Dye. Pigment. 2018, 152, 36–42. [Google Scholar] [CrossRef]
  13. Jia, Y.; Pan, Y.; Li, Y.; Zhang, L.; Lian, H.; Lin, J. Improved moisture-resistant and luminescence properties of a red phosphor based on dodec-fluoride K3RbGe2F12: Mn4+ through surface modification. Inorg. Chem. 2021, 60, 231–238. [Google Scholar] [CrossRef] [PubMed]
  14. Zhou, Y.Y.; Song, E.H.; Brik, M.G.; Wang, Y.J.; Hu, T.; Xia, Z.G.; Zhang, Q.Y. Non-equivalent Mn4+ doping into A2NaScF6 (A = K, Rb, Cs) hosts toward short fluorescence lifetime for backlight display application. J. Mater. Chem. C 2019, 7, 9203–9210. [Google Scholar] [CrossRef]
  15. Pust, P.; Schmidt, P.J.; Schnick, W. A revolution in lighting. Nat. Mater. 2015, 14, 454–458. [Google Scholar] [CrossRef]
  16. Xu, W.; Chen, D.; Yuan, S.; Zhou, Y.; Li, S. Tuning excitation and emission of Mn4+ emitting center in Y3Al5O12 by cation substitution. Chem. Eng. J. 2017, 317, 854–861. [Google Scholar] [CrossRef]
  17. Gao, J.; Zhu, H.; Li, R.; Huang, D.; Luo, B.; You, W.; Ke, J.; Yi, X.; Shang, X.; Xu, J.; et al. Moisture-resistant and highly efficient narrow-band red-emitting fluoride phosphor K2NaGaF6: Mn4+ for warm white LED application. J. Mater. Chem. C 2019, 7, 7906–7914. [Google Scholar] [CrossRef]
  18. Shang, M.; Li, C.; Lin, J. How to produce white light in a single-phase host? Chem. Soc. Rev. 2014, 43, 1372–1386. [Google Scholar] [CrossRef]
  19. 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]
  20. Fang, S.; Han, T.; Lang, T.; Zhong, Y.; Liu, B.; Cao, S.; Peng, L.; Yakovlev, A.N.; Korepanov, V.I. Synthesis of a novel red phosphor K2xBa1-xTiF6: Mn4+ and its enhanced luminescence performance, thermal stability and waterproofness. J. Alloys Compd. 2019, 808, 151697. [Google Scholar] [CrossRef]
  21. Yang, C.; Zhang, Z.; Hu, G.; Cao, R.; Liang, X.; Xiang, W. A novel deep red phosphor Ca14Zn6Ga10O35: Mn4+ as color converter for warm W-LEDs: Structure and luminescence properties. J. Alloys Compd. 2017, 694, 1201–1208. [Google Scholar] [CrossRef]
  22. Li, Q.; Yu, L.; Wu, W.; Liu, C.; Liu, M.; Huang, L.; Zhao, Y.; Lai, X.; Bi, J.; Gao, D. Novel BaGe1-xSixF6: Mn4+ (0 ≤ x ≤ 1) red phosphors for warm white LEDs: Hydrothermal synthesis and photoluminescence properties. J. Alloys Compd. 2021, 852, 156995. [Google Scholar] [CrossRef]
  23. Park, J.Y.; Joo, J.S.; Yang, H.K.; Kwak, M. Deep red-emitting Ca14Al10Zn6O35: Mn4+ phosphors for WLED applications. J. Alloys Compd. 2017, 714, 390–396. [Google Scholar] [CrossRef]
  24. Fang, M.-H.; Hsu, C.-S.; Su, C.; Liu, W.; Wang, Y.-H.; Liu, R.-S. Integrated surface modification to enhance the luminescence properties of K2TiF6: Mn4+ phosphor and Its application in white-light-emitting diodes. ACS Appl. Mater. Interfaces 2018, 10, 29233–29237. [Google Scholar] [CrossRef]
  25. Li, Q.; Luo, R.; Feng, L.; Dong, C.; Ning, Z.; Liu, M.; Zhao, Y.; Lai, X.; Bi, J.; Gao, D. A novel red phosphor of BaGe(1−x)TixF6: Mn4+ solid solution: Facile hydrothermal controlled synthesis, microstructures and luminescent properties. J. Mater. Chem. C 2019, 7, 11265–11275. [Google Scholar] [CrossRef]
  26. 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]
  27. Deng, T.T.; Song, E.H.; Zhou, Y.Y.; Wang, L.Y.; Zhang, Q.Y. Tailoring photoluminescence stability in double perovskite red phosphors A2BAlF6: Mn4+ (A = Rb, Cs; B = K, Rb) via neighboring-cation modulation. J. Mater. Chem. C 2017, 5, 12422–12429. [Google Scholar] [CrossRef]
  28. 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]
  29. Zhou, Y.; Ming, H.; Zhang, S.; Deng, T.; Song, E.; Zhang, Q. Unveiling Mn4+ substitution in oxyfluoride phosphor Rb2MoO2F4: Mn4+ applied to wide-gamut fast-response backlight displays. Chem. Eng. J. 2021, 415, 128974. [Google Scholar] [CrossRef]
  30. Xi, L.; Pan, Y.; Huang, S.; Liu, G. Mn4+ doped (NH4)2TiF6 and (NH4)2SiF6 micro-crystal phosphors: Synthesis through ion exchange at room temperature and their photoluminescence properties. RSC Adv. 2016, 6, 76251–76258. [Google Scholar] [CrossRef]
  31. Huang, L.; Zhu, Y.; Zhang, X.; Zou, R.; Pan, F.; Wang, J.; Wu, M. HF-free hydrothermal route for synthesis of highly efficient narrow-band red emitting phosphor K2Si1–xF6: xMn4+ for warm white light-emitting diodes. Chem. Mater. 2016, 28, 1495–1502. [Google Scholar] [CrossRef]
  32. Wu, W.-L.; Fang, M.-H.; Zhou, W.; Lesniewski, T.; Mahlik, S.; Grinberg, M.; Brik, M.G.; Sheu, H.-S.; Cheng, B.-M.; Wang, J.; et al. High color eendering index of Rb2GeF6: Mn4+ for light-emitting diodes. Chem. Mater. 2017, 29, 935–939. [Google Scholar] [CrossRef]
  33. Tang, F.; Su, Z.; Ye, H.; Wang, M.; Lan, X.; Phillips, D.L.; Cao, Y.; Xu, S. A set of manganese ion activated fluoride phosphors (A2BF6: Mn4+, A = K, Na, B = Si, Ge, Ti): Synthesis below 0 °C and efficient room-temperature photoluminescence. J. Mater. Chem. C 2016, 4, 9561–9568. [Google Scholar] [CrossRef]
  34. Zhong, J.S.; Chen, D.Q.; Wang, X.; Chen, L.F.; Yu, H.; Ji, Z.G.; Xiang, W.D. Synthesis and optical performance of a new red-emitting ZnTiF6·6H2O: Mn4+ phosphor for warm white-light-emitting diodes. J. Alloys Compd. 2016, 662, 232–239. [Google Scholar] [CrossRef]
  35. Zhou, Y.; Zhou, Q.; Liu, Y.; Wang, Z.; Yang, H.; Wang, Q. Hydrothermal synthesis and luminescent properties of BaTiF6: Mn4+ red phosphor for LED backlighting. Mater. Res. Bull. 2016, 73, 14–20. [Google Scholar] [CrossRef]
  36. 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]
  37. 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]
  38. 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]
  39. 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] [Green Version]
  40. Lin, H.; Hu, T.; Huang, Q.; Cheng, Y.; Wang, B.; Xu, J.; Wang, J.; Wang, Y. Non-Rare-Earth K2XF7: Mn4+ (X = Ta, Nb): A Highly-Efficient Narrow-Band Red Phosphor Enabling the Application in Wide-Color-Gamut LCD. Laser Photonics Rev. 2017, 11, 1700148. [Google Scholar] [CrossRef]
  41. Kumada, N.; Yanagida, S.; Takei, T.; Hong, B. Hydrothermal synthesis and crystal structure of new red phosphors, KNaMF7: Mn4+ (M: Nb, Ta). Mater. Res. Bull. 2019, 115, 170–175. [Google Scholar] [CrossRef]
  42. Jansen, T.; Baur, F.; Jüstel, T. Red emitting K2NbF7: Mn4+ and K2TaF7: Mn4+ for warm-white LED applications. J. Lumin. 2017, 192, 644–652. [Google Scholar] [CrossRef]
  43. Hong, F.; Pang, G.; Diao, L.; Fu, Z.; Liu, G.; Dong, X.; Yu, W.; Wang, J. Local structure modulation of Mn4+-doped Na2Si1-yGeyF6 red phosphors for enhancement of emission intensity, moisture resistance, thermal stability and application in warm pc-WLEDs. Dalton Trans. 2020, 49, 13805–13817. [Google Scholar] [CrossRef] [PubMed]
  44. Dat, L.Q.; Lan, N.T.; Lien, N.T.K.; Quang, N.V.; Anh, C.V.; Tinh, N.H.; Duyen, N.T.; Thoa, V.T.K.; Hanh, N.T.; Huong, P.T.L.; et al. Excellent hydrophobic property of K3AlF6: Mn4+ phosphor by coating with reduction graphene oxide on the surface of materials. Opt. Mater. 2022, 129, 112552. [Google Scholar] [CrossRef]
  45. Liu, Y.; Zhou, Z.; Huang, L.; Brik, M.G.; Si, S.; Lin, L.; Xuan, T.; Liang, H.; Qiu, J.; Wang, J. High-performance and moisture-resistant red-emitting Cs2SiF6: Mn4+ for high-brightness LED backlighting. J. Mater. Chem. C 2019, 7, 2401–2407. [Google Scholar] [CrossRef]
  46. Zhou, Y.; Song, E.; Deng, T.; Wang, Y.; Xia, Z.; Zhang, Q. Surface passivation toward highly stable Mn4+-activated red-emitting fluoride phosphors and enhanced photostability for white LEDs. Adv. Mater. Interfaces 2019, 6, 1802006. [Google Scholar] [CrossRef]
  47. Zhou, J.; Chen, Y.; Jiang, C.; Milićević, B.; Molokeev, M.S.; Brik, M.G.; Bobrikov, I.A.; Yan, J.; Li, J.; Wu, M. High moisture resistance of an efficient Mn4+-activated red phosphor Cs2NbOF5: Mn4+ for WLEDs. Chem. Eng. J. 2021, 405, 126678. [Google Scholar] [CrossRef]
  48. Beuter, A.; Kuhlmnn, W.; Sawodny, W. Vibrational spectra of hepta- and octafluorocomplexes of molybdenum (VI), tungsten (VI) and rhenium (VI). J. Fluor. Chem. 1975, 6, 367–378. [Google Scholar] [CrossRef]
  49. Peng, D.; He, S.; Zhang, Y.; Yao, L.; Nie, W.; Liao, Z.; Cai, W.; Ye, X. Blue light-induced rare-earth free phosphors for the highly sensitive and selective imaging of latent fingerprints based on enhanced hydrophobic interaction. J. Mater. 2022, 8, 229–238. [Google Scholar] [CrossRef]
  50. Wang, Z.; Yang, Z.; Yang, Z.; Wei, Q.; Zhou, Q.; Ma, L.; Wang, X. Red phosphor Rb2NbOF5: Mn4+ for warm white light-emitting diodes with a high color-rendering index. Inorg. Chem. 2019, 58, 456–461. [Google Scholar] [CrossRef]
  51. Jiang, C.; Brik, M.G.; Srivastava, A.M.; Li, L.; Peng, M. Significantly conquering moisture-induced luminescence quenching of red line-emitting phosphor Rb2SnF6: Mn4+ through H2C2O4 triggered particle surface reduction for blue converted warm white light-emitting diodes. J. Mater. Chem. C 2019, 7, 247–255. [Google Scholar] [CrossRef]
  52. Gao, X.; Song, Y.; Liu, G.; Dong, X.; Wang, J.; Yu, W. Narrow-band red emitting phosphor BaTiF6: Mn4+: Preparation, characterization and application for warm white LED devices. Dalton Trans. 2016, 45, 17886–17895. [Google Scholar] [CrossRef] [PubMed]
  53. Pang, G.; Hong, F.; Liu, X.; Zhang, Y.; Zhang, C.; Dong, T.; Fu, Z.; Liu, G.; Wang, J.; Li, D.; et al. Moisture-resistant Nb-based fluoride K2NbF7: Mn4+ and oxyfluoride phosphor K3(NbOF5)(HF2): Mn4+: Synthesis, improved luminescence performance and application in warm white LEDs. Dalton Trans. 2021, 50, 17290–17300. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD patterns of K2Nb1−xMoxF7: Mn4+ phosphors, (b) crystal structure of K2Nb1−xMoxF7: Mn4+ solid solution.
Figure 1. (a) XRD patterns of K2Nb1−xMoxF7: Mn4+ phosphors, (b) crystal structure of K2Nb1−xMoxF7: Mn4+ solid solution.
Molecules 28 04566 g001
Figure 2. SEM images of the representative K2Nb1−xMoxF7: Mn4+ (x = 0, 0.05, 0.15) samples. (a,d): x = 0; (b,e): x = 0.05; (c,f): x = 0.15.
Figure 2. SEM images of the representative K2Nb1−xMoxF7: Mn4+ (x = 0, 0.05, 0.15) samples. (a,d): x = 0; (b,e): x = 0.05; (c,f): x = 0.15.
Molecules 28 04566 g002
Figure 3. Excitation spectra (a), emission spectra (b), CIE chromaticity coordinates (c), and decay curves (d) of K2Nb1−xMoxF7: Mn4+ phosphors.
Figure 3. Excitation spectra (a), emission spectra (b), CIE chromaticity coordinates (c), and decay curves (d) of K2Nb1−xMoxF7: Mn4+ phosphors.
Molecules 28 04566 g003
Figure 4. Temperature-dependent emission spectra of K2Nb1−xMoxF7: Mn4+ with x = 5% (KNMF) (a); Normalized emission intensity of the red emission peak at 627 nm as the function of temperature for KNMF (b); Plot of ln(I0/I-1) versus 1/kT for KNMF (c); Configurational coordinate diagram of Mn4+ in the K2Nb1−xMoxF7: Mn4+ phosphors (d).
Figure 4. Temperature-dependent emission spectra of K2Nb1−xMoxF7: Mn4+ with x = 5% (KNMF) (a); Normalized emission intensity of the red emission peak at 627 nm as the function of temperature for KNMF (b); Plot of ln(I0/I-1) versus 1/kT for KNMF (c); Configurational coordinate diagram of Mn4+ in the K2Nb1−xMoxF7: Mn4+ phosphors (d).
Molecules 28 04566 g004
Figure 5. Photographs of K2NbF7: Mn4+ (KNF) and K2Nb1−xMoxF7: Mn4+ with x = 0.05 (KNMF) phosphors after immersing into deionized water for different times under the sunlight (a) and blue light excitation at 470 nm (b). The normalized intensity of the strongest peak at 627 nm of the KNF and KNMF phosphors as a function of immersion time (c).
Figure 5. Photographs of K2NbF7: Mn4+ (KNF) and K2Nb1−xMoxF7: Mn4+ with x = 0.05 (KNMF) phosphors after immersing into deionized water for different times under the sunlight (a) and blue light excitation at 470 nm (b). The normalized intensity of the strongest peak at 627 nm of the KNF and KNMF phosphors as a function of immersion time (c).
Molecules 28 04566 g005
Figure 6. Electroluminescence spectra (a) and CIE chromaticity diagram (b) of the WLEDs were packaged by employing a blue chip (InGaN) with yellow YAG: Ce3+ phosphor—A, blue chip (InGaN) with the mixtures of YAG: Ce3+ and KNF—B, blue chip (InGaN) with the mixtures of YAG: Ce3+ and KNMF—C. Insets show the photographs of the fabricated WLEDs.
Figure 6. Electroluminescence spectra (a) and CIE chromaticity diagram (b) of the WLEDs were packaged by employing a blue chip (InGaN) with yellow YAG: Ce3+ phosphor—A, blue chip (InGaN) with the mixtures of YAG: Ce3+ and KNF—B, blue chip (InGaN) with the mixtures of YAG: Ce3+ and KNMF—C. Insets show the photographs of the fabricated WLEDs.
Molecules 28 04566 g006
Table 1. XRD refinement results and lattice parameters of the K2Nb1−xMoxF7: Mn4+ (0≤ x ≤ 0.15) crystals.
Table 1. XRD refinement results and lattice parameters of the K2Nb1−xMoxF7: Mn4+ (0≤ x ≤ 0.15) crystals.
K2Nb1−xMoxF7: Mn4+x = 0x = 3%x = 5%x = 7%x = 10%x = 15%
Rwp (%)7.107.336.957.586.835.47
Rp (%)5.015.235.015.275.104.10
χ23.6123.8893.5374.3603.4292.193
a (Å)5.84525.84665.84695.84805.84765.8470
b (Å)12.692212.691312.691012.691112.692412.6947
c (Å)8.51388.51348.51388.51388.51448.5156
Cell volume (Å3)631.624631.697631.752631.871631.930632.067
Table 2. The actual doping concentrations of Mo6+ ions and Mn4+ ions in the K2Nb1−xMoxF7: Mn4+ crystals.
Table 2. The actual doping concentrations of Mo6+ ions and Mn4+ ions in the K2Nb1−xMoxF7: Mn4+ crystals.
The Designed K2Nb1−xMoxF7: Mn4+
Samples
Actual Molar Ratio of Mo6+ in
K2Nb1−xMoxF7: Mn4+ Crystals
(%)
Actual Molar Ratio of Mn4+ in
K2Nb1−xMoxF7: Mn4+ Crystals
(%)
x = 001.21
x = 3%0.241.19
x = 5%0.391.19
x = 7%0.911.21
x = 10%1.051.24
x = 15%1.461.48
Table 3. CIE chrominance coordinates and CCT values of the K2Nb1−xMoxF7: Mn4+ phosphors.
Table 3. CIE chrominance coordinates and CCT values of the K2Nb1−xMoxF7: Mn4+ phosphors.
K2Nb1−xMoxF7: Mn4+CIE (x, y)Color Purity (%)CCT (K)
x = 0(0.6759, 0.3223)93.603533
x = 3%(0.6755, 0.3226)93.503513
x = 5%(0.6779, 0.3206)94.113645
x = 7%(0.6770, 0.3214)93.883592
x = 10%(0.6771, 0.3213)93.903599
x = 15%(0.6775, 0.3210)94.003619
Table 4. Comparison of the thermal stability between K2Nb1−xMoxF7: Mn4+ (x = 0.05) (KNMF) phosphor with the reported some typical Mn4+-activated fluoride phosphors.
Table 4. Comparison of the thermal stability between K2Nb1−xMoxF7: Mn4+ (x = 0.05) (KNMF) phosphor with the reported some typical Mn4+-activated fluoride phosphors.
SamplesPL Normalized IntensityActivation Energy (eV)Ref.
K2Nb1−xMoxF7: Mn4+ (x = 0.05)69.95%@353 K0.74This work
K2NbF7: Mn4+70%@348 K0.66[4]
K2TiF6: Mn4+/0.70[4]
K2LiAlF6: Mn4+/0.62[4]
K2NaGaF6: Mn4+76%@398 K/[17]
BaTiF6: Mn4+44%@425 K0.628[20]
K0.07Ba0.965TiF6: Mn4+60%@425 K0.940[20]
Cs2KAlF6: Mn4+59.8%@423 K/[27]
Cs2RbAlF6: Mn4+72.1%@423 K/[27]
K2LiAlF6: Mn4+51.5%@423 K/[27]
(NH4)2TiF6: Mn4+50%@343 K0.3123[30]
(NH4)2SiF6: Mn4+64%@323 K0.4619[30]
K2TaF7: Mn4+70.9%@343 K/[40]
BaTiF6: Mn4+70%@425 K0.84[52]
Table 5. Comparison of moisture resistance between K2Nb1−xMoxF7: Mn4+ (x = 0.05) (KNMF) phosphor with the reported typical Mn4+-activated fluoride phosphors.
Table 5. Comparison of moisture resistance between K2Nb1−xMoxF7: Mn4+ (x = 0.05) (KNMF) phosphor with the reported typical Mn4+-activated fluoride phosphors.
SamplesImmersion TimeIntensityReference
K2Nb1-xMoxF7: Mn4+ (x = 0.05)1440 min86.37%This work
K2NbF7: Mn4+1440 min69.88%This work
K2TiF6: Mn4+120 min6.8%[1]
K2TiF6: Mn4+@CaF2120 min86.4%[1]
BaGeF6: Mn4+7200 min15%[10]
BaGeF6: Mn4+@PPG7200 min35%[10]
K3RbGe2F12: Mn4+600 min40%[13]
K2NaGaF6: Mn4+(2.04 at%)420 min83%[17]
K2SiF6: Mn4+(3.31 at%)420 min23%[17]
K2TiF6: Mn4+150min20%[20]
BaTiF6: Mn4+150 min50%[20]
K0.07Ba0.965TiF6: Mn4+150 min65%[20]
Cs2KAlF6: Mn4+4320 min65%[27]
Na2SiF6: 0.06Mn4+300 min32%[43]
Na2GeF6: 0.06Mn4+300 min33%[43]
Na2Si0.5Ge0.5F6: 0.06Mn4+300 min71%[43]
Rb2SnF6: Mn4+30 min10.3%[51]
K3(NbOF5)(HF2): Mn4+360 min74%[53]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gao, Y.; Feng, L.; Wang, L.; Zheng, J.; Ren, F.; Liu, S.; Ning, Z.; Zhou, T.; Wu, X.; Lai, X.; et al. Novel Mn4+-Activated K2Nb1−xMoxF7 (0 ≤ x ≤ 0.15) Solid Solution Red Phosphors with Superior Moisture Resistance and Good Thermal Stability. Molecules 2023, 28, 4566. https://doi.org/10.3390/molecules28114566

AMA Style

Gao Y, Feng L, Wang L, Zheng J, Ren F, Liu S, Ning Z, Zhou T, Wu X, Lai X, et al. Novel Mn4+-Activated K2Nb1−xMoxF7 (0 ≤ x ≤ 0.15) Solid Solution Red Phosphors with Superior Moisture Resistance and Good Thermal Stability. Molecules. 2023; 28(11):4566. https://doi.org/10.3390/molecules28114566

Chicago/Turabian Style

Gao, Yuhan, Lei Feng, Linglin Wang, Jun Zheng, Feiyao Ren, Siyu Liu, Zhanglei Ning, Ting Zhou, Xiaochun Wu, Xin Lai, and et al. 2023. "Novel Mn4+-Activated K2Nb1−xMoxF7 (0 ≤ x ≤ 0.15) Solid Solution Red Phosphors with Superior Moisture Resistance and Good Thermal Stability" Molecules 28, no. 11: 4566. https://doi.org/10.3390/molecules28114566

Article Metrics

Back to TopTop