Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Broadly tunable full-visible-spectrum ZnS/ZnO:Mn2+ composite microphosphor for warm WLED applications

Hoang Gia Chuca, Pham Minh Tria, Manh Trung Tran*a, Do Quang Trung*b, Nguyen Tub, Nguyen Van Dub, Nguyen Minh Hieua, Nguyen Duc Trung Kienc, Ta Ngoc Bachd, Le Tien Hae, Nguyen Duy Hungf, Dao Xuan Vietf, Ngo Ngoc Haa and Pham Thanh Huya
aFaculty of Materials Science and Engineering, Phenikaa School of Engineering, Phenikaa University, Yen Nghia Ha-Dong District, Hanoi, 12100, Vietnam. E-mail: trung.tranmanh@phenikaa-uni.edu.vn
bFaculty of Fundamental Science, Phenikaa University, Yen Nghia Ha-Dong District, Hanoi, 12100, Vietnam. E-mail: trung.doquang@phenikaa-uni.edu.vn
cFaculty of Electrical and Electronic Engineering, Phenikaa University, Yen Nghia, Ha-Dong District, Hanoi 10000, Vietnam
dInstitute of Materials Science, Vietnam Academy of Science and Technology, 18 Hoang Quoc Viet Street, Cau Giay District, Hanoi 10000, Vietnam
eInstitute of Science and Technology, TNU-University of Sciences, Thai Nguyen, 250000, Vietnam
fDepartment of Electronic Materials and Devices, School of Materials Science and Engineering, Hanoi University of Science and Technology (HUST), 01 Dai Co Viet Street, Hanoi 10000, Vietnam

Received 16th May 2025 , Accepted 14th July 2025

First published on 22nd July 2025


Abstract

Phosphor development for warm white LEDs (w-WLEDs) is key to enhancing light quality, energy efficiency, stability, and environmental sustainability. Mn2+-doped phosphors, known for their broad visible emission, are especially promising for UV-pumped w-WLEDs. While halide and garnet systems show potential, they suffer from toxicity and complex synthesis, making ZnO and ZnS attractive, stable alternatives. In this study, tunable full-visible-spectrum ZnS/ZnO:Mn2+ phosphors were synthesized via a simple thermal diffusion method. Structural analyses (XRD, Raman) revealed temperature-driven phase evolution, while PL and PLE measurements confirmed strong UV absorption and broad 400–700 nm emission from both host and Mn2+ centers. Emission was effectively tuned by Mn2+ concentration, with the ZnS/ZnO:0.1% Mn2+ sample emitting warm white light (CIE: x = 0.35, y = 0.37; CCT: 4920 K; decay time: 0.35 ms; activation energy: 0.31 eV). A phosphor-coated LED prototype demonstrated high luminous efficacy (127.7 lm per W), CIE coordinates (x, y) = (0.4974, 0.3405), CRI of 72, and a correlated color temperature (CCT) of 2506 K, suitable for warm lighting applications.


1 Introduction

Light is a crucial factor affecting the survival of living organisms. Finding an artificial light source that closely mimics sunlight remains a significant challenge, yet it represents an area of immense research potential. With this goal in mind, humans have developed successive generations of lighting technologies such as fluorescent lamps, compact fluorescent lamps, and now, LEDs and WLEDs. These technologies are widely adopted due to advantages such as low power consumption, high luminous efficiency, and the absence of mercury, allowing them to dominate the lighting market compared to other options.1,2 The current method of producing WLEDs involves combining blue LED chips with the commercial yellow phosphor YAG:Ce3+ (Y3Al5O12:Ce3+),3,4 which offers good conversion efficiency. However, the blue light emitted from the blue chip can suppress melatonin—a hormone responsible for regulating circadian rhythms—potentially leading to sleep disturbances. Additionally, the emission spectrum of commercial WLEDs lacks green and red regions, which reduces color rendering accuracy.

Thanks to their broad-band emission, which surpasses that of many traditional rare-earth ions (Eu2+, Ce3+, etc.) and other transition metal ions (Bi2+, Bi3+, Ni2+, etc.), Mn2+-doped phosphors hold strong potential for UV-driven tunable warm WLED applications.5 Notable examples include (C6H18N2O2)PbBr4:Mn2+,6 BaAl12O19:Mn2+, SrAl12O19:Mn4+,7 Na3LuSi2O7:Eu2+,Mn2+.8 However, these materials often face challenges such as moisture sensitivity, lead toxicity, limited thermal stability, complex synthesis, doping imbalance, and performance inconsistency due to sensitivity to doping and thermal quenching under high-power operation. Several studies have shown that wide bandgap semiconductors like ZnO and ZnS are promising alternatives to complex oxides as host matrices,9–11 offering strong UV resistance and excellent compatibility with transition metal ions.12–14 Recently, ZnS/ZnO-based materials, including ZnS/ZnO phosphors15 and ZnS/ZnO:Mn2+ nanobelts,13 have been successfully fabricated, exhibiting full-visible-spectrum emission. However, their practical application remains limited by complex fabrication processes and poor scalability. To overcome these issues, surface diffusion has emerged as an efficient doping technique, enabling scalable synthesis of phosphors for warm white LED applications.16

Herein, we developed a ZnS/ZnO:Mn2+ material using a surface diffusion method. ZnS/ZnO samples doped with 0.1% Mn2+ were annealed in an argon atmosphere for 2 hours at various temperatures. These samples have dual emission bands, including 450–550 nm from the ZnS/ZnO host and a strong 560–700 nm band from Mn2+ ions. The influence of annealing temperature on crystal phase formation was also systematically investigated.

2 Experimental

2.1 Synthesis process of Mn2+ doped ZnS/ZnO materials

High-purity starting materials, including ZnS (99.9%, Merck), ZnO (99.9%, Merck), and MnCl2·4H2O (99.9%, Merck), were used to synthesize the ZnS/ZnO:Mn2+ phosphors via a facile thermal diffusion method.16,17 ZnS and ZnO were first ball-milled at 500 rpm for 30 minutes, then stirred in 200 mL of deionized water at room temperature for 1 hour to form a uniform ZnS/ZnO mixture. A 1 mol per L MnCl2·4H2O solution was gradually added, allowing Mn2+ ions to diffuse into the host lattice. The resulting white mixture was dried at 120 °C for 10 hours (Fig. 1a), then sintered in argon gas at temperatures ranging from 600 to 1200 °C for 2 hours to obtain the final Mn2+-doped ZnS/ZnO powder (Fig. 1b). The thermal diffusion mechanism is illustrated in Fig. 1c.16
image file: d5ra03465e-f1.tif
Fig. 1 (a and b) Schematic of the ZnS/ZnO:Mn2+ fabrication process; (c) illustration of the Mn2+ diffusion mechanism.

2.2 Characterization

The phase compositions and crystalline structures of all samples were determined using an X-ray powder diffractometer (XRD D8 advanced) with CuKα radiation (λ = 0.1541 Å). Surface morphology and chemical composition were examined through FESEM images and EDS spectra based on JSM-7600F equipment. Raman spectra were recorded using an XploRA Plus Raman microscope (Horiba, France) with a 532 nm laser excitation source. The binding energies of Zn, S, O, and Mn were analyzed with an Al Kα X-ray photoelectron spectrometer (XPS, Thermo Scientific, USA). The optical properties of the samples were measured using a photoluminescence spectrophotometer (Nanolog Horiba) with excitation from a 450 W xenon discharge lamp. The optical properties of the LED device were assessed using a Gamma Scientific GS-1290 spectroradiometer (RadOMA).

3 Results and discussion

Fig. 2a and b show the XRD patterns of ZnS/ZnO:0.1% Mn2+ powers sintered in argon gas at temperatures ranging from 600 to 1200 °C for 6 hours. At 600 °C, the cubic ZnS phase predominates, displaying characteristic peaks at 2θ = 28.54°, 33.32°, 47.53°, and 56.31°, corresponding to the (111), (200), (220), and (311) planes, respectively.18–20 As the temperature increases to between 800 and 1200 °C, the hexagonal ZnS phase begins to appear, with new peaks at 2θ = 26.92°, 28.65°, 30.50°, and 39.50°, which are associated with the (100), (002), (101), and (102) diffraction planes of hexagonal ZnS (#36-1450).13,21,22 Concurrently, the hexagonal ZnO phase becomes more apparent, indicated by peaks at 2θ = 31.80°, 34.40°, and 36.23°, corresponding to the (100), (002), and (101) diffraction planes.13,21,23–25 The intensities of these peaks gradually increase with higher sintering temperatures, suggesting an improvement in the crystalline quality of both ZnS and ZnO.
image file: d5ra03465e-f2.tif
Fig. 2 (a and b) X-ray diffraction (XRD) patterns of the ZnS/ZnO:0.1% Mn2+ sample sintered at 600–1200 °C. (c and d) XRD patterns of ZnS/ZnO:x% Mn2+ samples (x = 0.1–0.7) sintered at 1000 °C in an argon gas atmosphere.

Fig. 2c and d illustrate the XRD patterns of ZnS/ZnO:x% Mn2+ (x = 0.1–0.7) powders sintered at 1000 °C for 6 hours in argon. As shown in Fig. 2c, XRD patterns indicate that the structure of both ZnS and ZnO remains unchanged as the concentration of Mn2+ increases. However, the diffraction peak shifts towards a smaller 2θ angle with increasing Mn2+ concentration (Fig. 2d), indicating the substitution of larger-radius ions for smaller-radius ions in the ZnS/ZnO lattice.26

The crystallite size can be estimated using the Scherrer eqn (1):27

 
image file: d5ra03465e-t1.tif(1)
where D is the crystal size, λ is the X-ray wavelength (in this study, Cu Kα radiation, λ = 0.154 nm); θ is the Bragg angle (in degrees), k is the shape factor (0.94), β is the full width at half maximum (FWHM) of the diffraction peak. The FWHM value of the ZnS peak at approximately 28.6° decreases as the sintering temperature increases from 600 °C to 1200 °C, indicating larger and more refined ZnS crystals at higher temperatures. Using the Scherrer eqn (1), the average crystal sizes are estimated at approximately 29.34 nm for ZnS and 14.99 nm for ZnO. Additionally, Rietveld refinement analysis of ZnS/ZnO:0.1% Mn2+ samples sintered at 600–1200 °C was presented in Fig. 3a. The phase proportions, lattice parameters and unit cell volumes of the ZnS and ZnO phases were listed in Table 1.28 The low value of χ2 indicates a strong agreement between the experimental and the calculated data. At 600 °C, two distinct phases – cubic ZnS and hexagonal ZnO – were identified in a 7[thin space (1/6-em)]:[thin space (1/6-em)]3 ratio. As the temperature increases, the hexagonal ZnS phase emerges, and the crystallinity of both ZnS and ZnO phases becomes more refined.


image file: d5ra03465e-f3.tif
Fig. 3 Rietveld refinement analysis of (a–d) ZnS/ZnO:0.1% Mn2+ samples sintered at 600–1200 °C and (e–h) of ZnS/ZnO:x%Mn2+ samples (x = 0.1–0.7) sintered at 1000 °C.
Table 1 The fitted values from Rietveld refinement of ZnS/ZnO:0.1% Mn2+ samples sintered at 600–1200 °C
Temperature (°C) Phase Lattice parameters (Å) Lattice volume, V3) Phase fraction (%)
a c
600 ZnS (cubic) 5.39832   157.157 ∼67.33
ZnO 3.25022 5.20539 47.593 ∼32.67
Rp = 14.3% Rexp = 15.81% χ2 = 1.48  
800 ZnS (hex) 3.82515 6.26245 79.355 23.79
ZnS (cubic) 5.41190   158.508 41.07
ZnO 3.25259 5.20815 47.717 35.14
Rp = 13.6% Rexp = 14.86% χ2 = 1.58  
1000 ZnS (hex) 3.82531 6.26532 79.397 27.67
ZnS (cubic) 5.41238   158.549 36.95
ZnO 3.25778 5.21191 47.904 35.68
Rp = 15.1% Rexp = 14.57% χ2 = 1.99  
1200 ZnS (hex) 3.81957 6.25292 79.095 ∼58.20
ZnS (cubic) 5.40373   157.790 5.62
ZnO 3.25556 5.20908 47.871 ∼36.18
Rp = 12.6% Rexp = 14.81% χ2 = 1.31  


Rietveld refinement analysis of ZnS/ZnO:x% Mn2+ samples (x = 0.1–0.7) sintered at 800 °C in an argon atmosphere were shown in Fig. 3b. All fitted values including lattice parameters, volume, phase composition and standard deviations were illustrated in Table 2. As shown in Table 2, the phosphors contain both ZnS and ZnO phases, confirming the formation of a composite structure. With increasing Mn doping from 0.1% to 0.7%, the lattice constants and unit cell volumes of both phases increase, suggesting the substitution of larger Mn2+ (r = 0.83 Å) ions for smaller Zn2+ (r = 0.74 Å) ions in the ZnS/ZnO lattice.26

Table 2 The fitted values from Rietveld refinement of the XRD data for the ZnS/ZnO:x% Mn2+ (x = 0.1–0.7) samples sintered at 1000 °C
Concentration (%) Phase Lattice parameters (Å) Lattice volume, V3)
a c
0.1 ZnS (hex) 3.82269 6.25952 79.215
ZnS (cubic) 5.40726   158.100
ZnO 3.25041 5.20473 47.622
Rp = 13.6% Rexp = 14.70% χ2 = 1.52
0.3 ZnS (hex) 3.82452 6.26202 79.323
ZnS (cubic) 5.41123   158.448
ZnO 3.25404 5.20901 47.767
Rp = 15.8% Rexp = 16.18% χ2 = 1.59
0.5 ZnS (hex) 3.82404 6.26282 79.313
ZnS (cubic) 5.41008   158.348
ZnO 3.25807 5.21341 47.926
Rp = 15.1% Rexp = 15.12% χ2 = 1.67
0.7 ZnS (hex) 3.82628 6.26476 79.431
ZnS (cubic) 5.41283   158.589
ZnO 3.26127 5.21733 48.056
Rp = 14.0% Rexp = 15.87% χ2 = 1.37


Fig. 4a–d presents SEM images of ZnS/ZnO:0.1% Mn2+ phosphors sintered at temperatures from 600 °C to 1200 °C, showing clear microstructural evolution. At 600 °C, particles appear small, irregular, and highly agglomerated, indicating incomplete crystallization. By 800 °C, the particles become more defined with reduced agglomeration, indicating the onset of significant grain growth. At 1000 °C, the crystallites enlarge, surfaces smooth out, and overall uniformity improves, reflecting enhanced crystallinity. Finally, at 1200 °C, particles grow significantly larger and faceted, with clear grain boundaries due to sintering and densification. This temperature-driven evolution closely aligns with the crystallinity improvements observed in the XRD results.


image file: d5ra03465e-f4.tif
Fig. 4 (a–d) SEM images of ZnS/ZnO:0.1% Mn2+ samples sintered for 2 h in the argon at various temperatures of 600–1200 °C, (e) EDS elements and (f–i) EDS mapping spectra of ZnS/ZnO:0.1% Mn2+ sample sintered in argon at 1000 °C for 2 h.

Fig. 4e presents the EDS spectrum of the ZnS/ZnO:0.1% Mn sample sintered in argon at 1000 °C for 2 h, confirming the presence of Zn (49.91 at%), S (30.38 at%), O (19.01 at%), and Mn (0.69 at%) with no detectable impurities, indicating high purity. The corresponding EDS elemental maps (Fig. 4f–i) further demonstrate the uniform distribution of all constituent elements across the material.

Fig. 5 illustrates the Raman spectra of ZnS/ZnO:x% Mn2+ (x = 0.1–0.7) samples sintered at 1000 °C for 2 hours in argon gas, measured using a 785 nm laser across 100800 cm−1.26,29 As shown in Fig. 5a, ZnS exhibits characteristic peaks at 157, 217, 268, 348, and 412 cm−1,30,31 while ZnO features are observed at 98, 375, 437 and 535 cm−1.29,32 In Fig. 5b, increasing Mn2+ content leads to reduced intensity and a blue shift in Raman peaks, likely due to the substitution of larger Mn2+ (r = 0.83 Å) ions for smaller-radius Zn2+ (r = 0.74 Å) ions in the ZnS/ZnO lattice.33,34


image file: d5ra03465e-f5.tif
Fig. 5 (a) Raman spectra of ZnS/ZnO:x% Mn2+ (x = 0.1–0.7) samples sintered at 1000 °C for 2 h in argon; (b) enlarged view of the ZnS vibrational mode at 348 cm−1.

Fig. 6a shows the XPS survey spectrum of the ZnS/ZnO:0.1% Mn2+ sample sintered at 1000 °C for 2 hours in argon, calibrated using the binding energy of carbon (C 1s) with a reference peak at 284.6 eV.15,35 Fig. 6b, the Zn 2p3/2 peak is observed at approximately 1021.6 eV, consistent with the binding energy of Zn2+ in ZnS/ZnO.15 Fig. 6c shows S 2p peaks at ∼161.9 eV and ∼163.2 eV, corresponding to the S 2p3/2 và S 2p1/2 states.21 The O 1s spectrum reveals two component peaks: one at around 530.5 eV, attributed to O–Zn bonds in the host lattice,13 and another at approximately 532.0 eV, likely associated with adsorbed H2O or O2 on the surface.35 In Fig. 6d, the Mn 2p spectrum shows a peak at ∼640.0 eV, confirming the presence of Mn2+ ions in the ZnS and ZnO matrix.33,36


image file: d5ra03465e-f6.tif
Fig. 6 XPS spectra of (a) survey spectrum, (b) Zn 2p, (c) S 2p, (d) O 1s and (e) Mn 2p of ZnS/ZnO:0.1% Mn2+ samples sintered at 1000 °C for 2 hours in argon.

Fig. 7a and b show the PLE spectra mentioned at 466 nm and 586 nm of ZnS/ZnO:0.1% Mn2+ samples sintered at 600–1200 °C for 2 hours in argon. A strong absorption band appears at 333–339 nm, attributed to ZnS band-to-band transitions,21,24,37 and a weaker peak at 372 nm corresponds to ZnO transitions.25,38 Fig. 7c and d show the PL spectra and responding to the normalized PL spectra of undoped ZnS/ZnO and ZnS/ZnO:0.1% Mn2+ samples sintered at different temperatures in the range of 600–1200 °C. The undoped ZnS/ZnO sample exhibits a strong blue-green emission peaking at 495 nm, linked to defect-related emissions.39–42 In contrast, the ZnS/ZnO:0.1% Mn2+ samples show an additional broad emission centered at 586 nm, attributed to the 4T16A1 transition of Mn2+ ions in the ZnS/ZnO lattice.13,22,43 The emission intensity ratio between the yellow-orange and blue-green regions varies significantly with sintering temperature. Yellow-orange emission peaks at 800 °C, while blue-green emission is strongest at 1000 °C, likely due to competition between Mn2+ transitions and intrinsic defect emissions in ZnS/ZnO.44


image file: d5ra03465e-f7.tif
Fig. 7 PLE spectra mentioned at 466 nm (a) and 586 nm (b) of ZnS/ZnO:0.1% Mn2+ samples sintered at 1000 °C for 2 hours in argon. PL spectra excited at 333 nm (c) and corresponding to the normalized PL spectra (d) of ZnS/ZnO:0.1% Mn2+ samples sintered at 1000 °C for 2 hours in an argon gas environment.

To further elucidate this observation, the PL spectra of ZnS/ZnO:0.1% Mn2+ samples sintered at 600 to 1200 °C were deconvoluted into three bands corresponding to the blue, green, and yellow regions, as illustrated in Fig. 8(a–d).


image file: d5ra03465e-f8.tif
Fig. 8 Deconvoluted PL spectra of ZnS/ZnO:0.1% Mn2+ samples sintered for 2 hours in argon at (a) 600 °C; (b) 800 °C; (c) 1000 °C; and (d) 1200 °C.

Table 3 presents the yellow-orange to blue-green emission intensity ratios for samples dried at various temperatures. The sample annealed at 1000 °C shows the most optimal ratio for full-spectrum WLED applications, making it the chosen condition for subsequent experiments.

Table 3 Emission origin in ZnS/ZnO:Mn2+ samples
Intensity (a.u)
Radiation center Temperature (°C)
600 800 1000 1200
ZnS (vacancy S) 1.65 × 105 3.38 × 106 6.56 × 106 1.53 × 106
ZnO (vacancy O) 6.11 × 104 3.04 × 106 3.89 × 106 1.80 × 106
Mn (4T16A1) 2.78 × 106 1.73 × 107 1.28 × 107 1.12 ×107


Fig. 9a and b illustrates the PLE and PL spectra excited at 330 nm for ZnS/ZnO:x Mn2+ (x = 0.1–0.7%) sintered at 1000 °C. All spectra feature a prominent yellow-orange emission band attributed to the 4T16A1 transition of Mn2+ ions and a weaker blue-green emission related to intrinsic defects in the ZnS and ZnO hosts.


image file: d5ra03465e-f9.tif
Fig. 9 (a) PLE spectra, (b) PL spectra & ratio of emission intensity at peak 466 nm and 586 nm (c) CIE chromaticity coordinate of Zn/ZnO:x Mn2+ (x = 0.1–0.7%) and (d) PL decay curves of ZnS/ZnO:x% Mn2+ (x = 0.1–0.7) phosphors sintered for 2 hours at 1000 °C in argon.

For the yellow-orange emission band, the PL intensity initially increased with Mn2+ concentration from 0.1% to 0.5%, reaching a maximum at 0.5% Mn concentration before declining at higher concentrations – a typical luminescence quenching behavior (Fig. 9b). Meanwhile, the blue-green emission steadily weakened with increased Mn2+ content, likely due to competition between the 4T16A1 transition of Mn2+ ions and defect-related emissions from ZnS/ZnO.

Using ColorCalculator software, the chromaticity coordinates of the samples were analyzed before and after doping. As shown in Fig. 9c, the undoped ZnS/ZnO sample exhibits coordinates of (0.2013, 0.3765), falling in the cyan region, whereas the ZnS/ZnO: 0.1% Mn2+ sample shifts to (0.3490, 0.3693), positioning it in the white light zone. The PL spectral differences across varying Mn2+ concentrations (x = 0–0.7%) indicate 0.1% as the optimal doping level for warm-white light emission. A performance comparison with previously reported materials, detailed in Table 4, underscores the suitability of this phosphor for WLED applications.

Table 4 Emission and excitation peak of various powder manufactured on ZnS, ZnO-based materials
Materials Fabrication method Emission peaks (nm) Excitation peak (nm) Ref.
ZnS:Mn nanobelts Thermal evaporation 440, 510, and 577 325 13
ZnS/ZnO:Mn2+ nanobelts Thermal evaporation     13
ZnO:Mn2+ nanopowder Microwave-assisted hydrothermal 620 405 23
ZnS/ZnO micropyramid Thermal evaporation 500 341 24
ZnS:Mn2+ nanoparticles Co-precipitation 445 and 580 320 45
ZnO:Mn2+ nanorods Thermal evaporation 444 and 575 350 46
ZnS/ZnO: Mn2+composite microphosphors Thermal diffusion 466 and 586 333 This work


Fig. 9d illustrates the PL decay curves of ZnS/ZnO:0.1% Mn2+ samples at various doping concentrations, excited at 333 nm and monitored at 586 nm. The measured lifetimes, ranging from 0.78 ms to 0.56 ms for 0.3–0.7% Mn2+, reflect the characteristic long-lived emission of Mn2+ ions. At lower doping levels, Mn2+ ions tend to remain isolated, resulting in longer lifetimes. However, increased concentrations promote Mn–Mn interactions, which enhance non-radiative processes and reduce lifetimes.47 Interestingly, the 0.1% Mn2+ sample exhibits a shorter lifetime of 0.35 ms, likely due to additional emissions from defect states associated with sulfur and oxygen in the host lattice.

Temperature-dependent PL spectra of the ZnS/ZnO:0.1% Mn2+ sample (annealed at 1000 °C and excited at 333 nm) are shown in Fig. 10a. As temperature rises from 27 °C to 210 °C, an apparent decrease in emission intensity is observed. As shown in Fig. 10b, the PL intensity retains 24.94% of its initial value at 150 °C, indicating moderate thermal stability. This performance is comparable to that of NaGa11O17:Mn2+, which maintains 26.9% under similar conditions48 and is notably higher than that of Mn-doped Gd3Ga5O12 phosphor, which retains only 21.9%.49


image file: d5ra03465e-f10.tif
Fig. 10 (a) The temperature-dependent PL spectra excited at 333 nm of ZnS/ZnO:0.1% Mn2+ phosphors, (b) the corresponding integral normalized PL intensity as a function of temperature. (c) EL spectra of the NUV chip (366 nm) without and with coating a layer of the ZnS/ZnO:0.1% Mn2+ phosphor and (d) color coordinates of the fabricated LED.

The activation energy (ΔE) for thermal quenching was determined using the equation50:

 
image file: d5ra03465e-t2.tif(2)
where I0 is the initial PL intensity, I is the intensity at a given temperature, A is a constant, kB is the Boltzmann constant, and T is the absolute temperature. The inset in Fig. 10b shows a linear fit of ln(I0/I − 1) versus 1/T, from which ΔE was calculated to be approximately 0.31 eV. This value is higher than that of ZnS/ZnO phosphor (0.26 eV)15 and aligns well with reported activation energies for other Mn-doped systems, such as MnS/ZnS core–shell (0.393 eV),51 ZnSe:Mn2+ (0.37 eV),52 ZnS:Mn2+ nanocrystalline (0.62 eV).50

Fig. 10c shows the electroluminescence (EL) spectra of a 366 nm NUV chip driven at 200 mA, featuring a sharp emission at 366 nm from the chip and a broad visible emission from the ZnS/ZnO:0.1% Mn2+ phosphor coating. The 366 nm excitation, slightly red-shifted from the optimal ∼333 nm PLE peak (Fig. 9a), was selected for its commercial availability and ease of integration, despite 333 nm providing stronger photoluminescence but posing challenges in LED availability and stability. The broad emission confirms the phosphor's effective contribution to the visible range. The quantum efficiency (QE), a key performance metric, was calculated using the following equation:25

 
image file: d5ra03465e-t3.tif(3)

Based on this, the ZnS/ZnO:0.1% Mn phosphor annealed at 1000 °C achieved a QE of approximately 29.17%.

In addition, the luminous efficacy (lm per W), K, of the LED can be determined using the following formula:53

 
image file: d5ra03465e-t4.tif(4)
where Φ is the luminous flux (41.56 lm), P is the measured electrical power (0.33 W). Based on these values, the luminous efficacy was calculated to be approximately 127.7 lm per W. This is relatively high and comparable to white light emission technologies using micro-LEDs with RGB quantum dots (165 lm per W).54 It also surpasses the performance reported in several other studies, such as blue LED chips coated with an epoxy:yellow phosphor mixture (97[thin space (1/6-em)]:[thin space (1/6-em)]3 wt%), which achieved 23.2 lm per W,55 or UV LEDs combined with PFO polymer and CdSe/ZnS quantum dots, yielding only 13 lm per W.56

The corresponding CIE color coordinates of the fabricated LED are (x, y) = (0.4974, 0.3405), placing it within the warm white region (Fig. 10d), with a correlated color temperature (CCT) of 2506 K. These results highlight the potential of ZnS/ZnO:0.1% Mn2+ phosphor as an efficient and promising phosphor for warm-white WLED applications.

4 Conclusions

In this work, the full-visible-spectrum ZnS/ZnO:Mn2+ phosphors were successfully produced by using a thermal diffusion method. The phosphors exhibit strong absorption in the NUV/UV region, with prominent peaks at 333 nm and 372 nm, and emits a broad emission spanning 400–700 nm. Optimization of Mn2+ concentration and annealing temperature identified ZnS/ZnO:0.1% Mn2+, annealed at 1000 °C for 2 hours in argon, as the optimal composition. This sample shows a relatively long decay time of 0.35 ms, moderate thermal stability with 24.94% PL retention at 150 °C, and an activation energy of 0.31 eV. When applied to a 366 nm LED chip, it achieves a quantum efficiency of 29.17%, a CRI of 72, a luminous efficacy of 127.7 lm per W, and warm white emission with CIE coordinates (x, y) = (0.4974, 0.3405) and CTT of 2506 K. These results highlight its strong potential for high-performance warm WLED applications.

Data availability

The data supporting this article have been included as part of the ESI.†

Conflicts of interest

The authors declare no conflicts of interest.

References

  1. T. Pulli, T. Dönsberg, T. Poikonen, F. Manoocheri, P. Kärhä and E. Ikonen, Light:Sci. Appl., 2015, 4, 1–7 CrossRef.
  2. M. Bliss, T. R. Betts and R. Gottschalg, Reliab. Photovoltaics Cells Modules Comps., Syst., 2008, 7048, 704807 CrossRef.
  3. K. H. Loo, Y. M. Lai, S. C. Tan and C. K. Tse, IEEE Trans. Power Electron., 2012, 27, 974–984 Search PubMed.
  4. L. Chen, C. C. Lin, C. W. Yeh and R. S. Liu, Materials, 2010, 3, 2172–2195 CrossRef CAS.
  5. Y. Li, S. Qi, P. Li and Z. Wang, RSC Adv., 2017, 7, 38318–38334 RSC.
  6. G. Zhou, X. Jiang, M. Molokeev, Z. Lin, J. Zhao, J. Wang and Z. Xia, Chem. Mater., 2019, 31, 5788–5795 CrossRef CAS.
  7. Y. Zhu, C. Li, D. Deng, B. Chen, H. Yu, H. Li, L. Wang, C. Shen, X. Jing and S. Xu, J. Alloys Compd., 2021, 853, 157262 CrossRef CAS.
  8. K. Li, M. Xu, J. Fan, M. Shang, H. Lian and J. Lin, J. Mater. Chem. C, 2015, 3, 11618–11628 RSC.
  9. K. K. Dubey, V. Nayar and P. S. Choudhary, Mater. Sci. Res. India, 2010, 7, 195–200 CAS.
  10. X. Fang, Y. Bando, U. K. Gautam, T. Zhai, H. Zeng, X. Xu, M. Liao and D. Golberg, Crit. Rev. Solid State Mater. Sci., 2009, 34(3–4), 190–223 CrossRef CAS.
  11. F. Li, Y. Jiang, L. Hu, L. Liu, Z. Li and X. Huang, J. Alloys Compd., 2009, 474, 531–535 CrossRef CAS.
  12. D. Peipei, L. Dongjie, L. Guogang, A. A. Al Kheraif and L. Jun, Adv. Opt. Mater., 2020, 8(16), 19011993 Search PubMed.
  13. D. Q. Trung, M. T. Tran, N. D. Hung, Q. Nguyen Van, N. T. Huyen, N. Tu and H. Pham Thanh, Opt. Mater., 2021, 121, 111587 CrossRef CAS.
  14. R. Mahajan and R. Prakash, J. Mater. Sci.:Mater. Electron., 2022, 33, 25491–25517 CrossRef CAS.
  15. D. Q. Trung, M. T. Tran, N. Van Du, N. D. Hung, T. Thi, H. Tam, N. Tu, P. M. Tri, H. G. Chuc, N. Duc, T. Kien, N. Van Quang and P. T. Huy, RSC Adv., 2025, 616–627 RSC.
  16. M. T. Tran, D. Quang Trung, N. Tri Tuan, N. T. Tuan, N. Tu, N. Van Du, N. Duy Hung, N. Van Quang, T. Thi Hao Tam, N. D. Trung Kien, N. M. Hieu and P. T. Huy, Dalton Trans., 2023, 52, 12704–12716 RSC.
  17. M. Pudukudy and Z. Yaakob, Appl. Surf. Sci., 2014, 292, 520–530 CrossRef CAS.
  18. Y. Cheng, R. Chen, H. Feng, W. Hao, H. Xu, Y. Wang, J. Li, F. Li, S. Zhang and H. Sun, Phys. Chem. Chem. Phys., 2014, 16, 4544 RSC.
  19. M. Sundararajan, P. Sakthivel and A. C. Fernandez, J. Alloys Compd., 2018, 768, 553–562 CrossRef CAS.
  20. D. Son, D. R. Jung, J. Kim, T. Moon, C. Kim and B. Park, Appl. Phys. Lett., 2007, 90, 88–91 Search PubMed.
  21. B. Zhu, Q. Zhang, X. Li and H. Lan, Phys. Status Solidi A, 2018, 215, 1–6 Search PubMed.
  22. B. S. R. Devi, R. Raveendran and A. V. Vaidyan, Pramana - J. Phys., 2007, 68, 679–687 CrossRef CAS.
  23. F. C. Romeiro, J. Z. Marinho, A. C. A. Silva, N. F. Cano, N. O. Dantas and R. C. Lima, J. Phys. Chem. C, 2013, 117, 26222–26227 CrossRef CAS.
  24. D. Q. Trung, M. T. Tran, N. Tu, L. T. H. Thu, N. T. Huyen, N. D. Hung, D. X. Viet, N. D. T. Kien and P. T. Huy, Opt. Mater., 2022, 125, 1–11 CrossRef.
  25. D. Q. Trung, N. V. Quang, M. T. Tran, N. V. Du, N. Tu, N. D. Hung, D. X. Viet, D. D. Anh and P. T. Huy, Dalton Trans., 2021, 50, 9037–9050 RSC.
  26. M. A. Avilés, J. M. Córdoba, M. J. Sayagués and F. J. Gotor, J. Mater. Sci., 2020, 55, 1603–1613 CrossRef.
  27. F. I. H. Rhouma, F. Belkhiria, E. Bouzaiene, M. Daoudi, K. Taibi, J. Dhahri and R. Chtourou, RSC Adv., 2019, 9, 5205–5217 RSC.
  28. P. Schröer, P. Krüger and J. Pollmann, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 6971–6980 CrossRef PubMed.
  29. N. X. Sang, N. T. P. Anh, N. D. Dung, N. H. C. Huy, N. H. Tho and L. T. L. Anh, VNU J. Sci. Math.–Phys., 2019, 35, 69–76 Search PubMed.
  30. X. Wang, J. Shi, Z. Feng, M. Li and C. Li, Phys. Chem. Chem. Phys., 2011, 13, 4715–4723 RSC.
  31. U. P. Gawai, U. P. Deshpande and B. N. Dole, RSC Adv., 2017, 7, 12382–12390 RSC.
  32. Ü. Özgür, Y. I. Alivov, C. Liu, A. Teke, M. A. Reshchikov, S. Doǧan, V. Avrutin, S. J. Cho and H. Mork[o with combining cedilla], J. Appl. Phys., 2005, 98, 1–103 CrossRef.
  33. M. A. Kamran, A. Majid, T. Alharbi, M. W. Iqbal, M. W. Amjad, G. Nabi, S. Zou and B. Zou, J. Mater. Chem. C, 2017, 5, 8749–8757 RSC.
  34. C. Bi, L. Pan, M. Xu, J. Yin, Z. Guo, L. Qin, H. Zhu and J. Q. Xiao, Chem. Phys. Lett., 2009, 481, 220–223 CrossRef CAS.
  35. X. Liu, X. Chen, X. Cui and R. Yu, Ceram. Int., 2014, 40, 13847–13854 CrossRef CAS.
  36. Y. Tong, F. Cao, J. Yang, P. Tang and M. Xu, Mater. Lett., 2013, 94, 124–127 CrossRef CAS.
  37. J. Cao, J. Yang, Y. Zhang, Y. Wang, L. Yang and D. Wang, Opt. Mater., 2010, 32, 643–647 CrossRef CAS.
  38. E. A. Batista, A. C. A. Silva, T. K. de Lima, E. V. Guimarães, R. S. da Silva and N. O. Dantas, J. Alloys Compd., 2021, 850, 156611 CrossRef CAS.
  39. D. Q. Trung, N. T. Tuan, H. V. Chung, P. H. Duong and P. T. Huy, J. Lumin., 2014, 153, 321–325 CrossRef CAS.
  40. A. Kanti Kole, C. Sekhar Tiwary and P. Kumbhakar, J. Mater. Chem. C, 2014, 2, 4338–4346 RSC.
  41. N. Tu, D. Q. Trung, N. D. T. Kien, P. T. Huy and D. H. Nguyen, Phys. E, 2017, 85, 174–179 CrossRef CAS.
  42. P. C. Patel, S. Ghosh and P. C. Srivastava, Mater. Res. Bull., 2016, 81, 85–92 CrossRef CAS.
  43. S. Adachi, J. Lumin., 2023, 263, 119993 CrossRef CAS.
  44. T. Chanier, F. Virot and R. Hayn, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79(20), 205204 CrossRef.
  45. G. Murugadoss, B. Rajamannan and V. Ramasamy, J. Lumin., 2010, 130, 2032–2039 CrossRef CAS.
  46. Y. Tong, F. Cao, J. Yang, P. Tang and M. Xu, Mater. Lett., 2013, 94, 124–127 CrossRef CAS.
  47. T. W. Kang, K. W. Park, G. Deressa and J. S. Kim, J. Lumin., 2018, 194, 551–556 CrossRef CAS.
  48. G. Lu, Y. Wang, K. Ma, X. Chen, W. Geng, T. Liu, S. Xu, J. Zhang and B. Chen, Ceram. Int., 2024, 50, 16190–16200 CrossRef CAS.
  49. K. T. Thu, V. D. Huan, N. Tu, D. Q. Trung, N. Van Du, N. Van Quang, T. N. Bach, L. T. Ha, N. D. Hung, D. X. Viet, N. T. Tuan, N. M. Hieu, M. T. Tran and P. T. Huy, RSC Adv., 2025, 15, 8275–8286 RSC.
  50. J. F. Suyver, S. F. Wuister, J. J. Kelly and A. Meijerink, Nano Lett., 2001, 1, 429–433 CrossRef CAS.
  51. D. F. Fang, X. Ding, R. C. Dai, Z. Zhao, Z. P. Wang and Z. M. Zhang, Chin. Phys. B, 2014, 23(12), 127804 CrossRef.
  52. X. Yuan, J. Zheng, R. Zeng, P. Jing, W. Ji, J. Zhao, W. Yang and H. Li, Nanoscale, 2014, 6, 300–307 RSC.
  53. Z. P. Rimbawati, H. Alam and M. Irwanto, J. Phys.: Conf. Ser., 2021, 1811, 012106 CrossRef.
  54. H. V. Han, H. Y. Lin, C. C. Lin, W. C. Chong, J. R. Li, K. J. Chen, P. Yu, T. M. Chen, H. M. Chen, K. M. Lau and H. C. Kuo, Opt. Express, 2015, 23(25), 32504–32515 CrossRef CAS PubMed.
  55. C. S. Son, H. J. Chang, K. H. Jaekal, Y. C. Chang and S. W. Lee, Mater. Sci. Forum, 2006, 510, 106–109 Search PubMed.
  56. K. J. Chen, Y. C. Lai, B. C. Lin, C. C. Lin, S. H. Chiu, Z. Y. Tu, M. H. Shih, P. Yu, P. T. Lee, X. Li, H. F. Meng, G. C. Chi, T. M. Chen and H. C. Kuo, Opt. Express, 2015, 23(7), A204–A210 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.