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

Efficient near-infrared broadband garnet phosphor for pc-LED and its application to vascular visualization and night vision

Open Access Open Access

Abstract

Ultra-broadband near-infrared (NIR) spectroscopy has unparalleled application prospects in intelligent detection and phosphor-converted light-emitting diodes (pc-LED), which are most likely to become the next generation of NIR light sources, has become a hot spot for research nowadays. To cope with the demand for more NIR spectroscopy applications, more efficient NIR phosphors need to be developed. Here, by screening the subject with a smaller band gap and by screening the suitable ion electronegativity of the lattice position where the Cr3+ is located, and then through the energy transfer, a series of Gd3Zn2GaGe2O12:xCr3+, yYb3+ (GZGG:Cr3+/Yb3+) NIR broadband garnet phosphors were found for the first time. By controlling the energy transfer process, the internal quantum yield (IQY) (54.9%), external quantum yield (EQY) (24.65%), bandwidth (260 nm), and thermal stability (60% at 150 °C) of NIR emission were substantially improved. The obtained phosphors are packaged with blue light chips into pc-LED, which can be applied in different fields such as vascular visualization and night vision.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

NIR spectroscopy is uniquely suited for applications in non-destructive testing [1], plant lighting [2], food analyzed [3], and night vision [4,5]. Especially for light sources in the 700–1100 nm spectral range, the organic composition is detected by the absorption intensity of C-H, O-H, and N-H groups in the 700–1100 nm NIR band [5]. The NIR radiation can effectively penetrate biological tissues and can also be used for nondestructive imaging of subcutaneous blood vessels, thus providing a harmless and non-invasive diagnosis of the human physiological state [6]. For multi-disciplinary applications, a wider detection range and faster response time, as well as portable NIR light sources, need to be developed. Compared with traditional NIR light sources, NIR pc-LED technology has the advantages of long operating life, small size, low power consumption, etc. [7,8], which can be achieved by integrating NIR pc-LED into portable electronic devices to realize micro NIR systems. And phosphor as the core component of pc-LED needs to provide both the NIR output of the whole system and match the corresponding LED chip.

In recent years, NIR phosphors with transition metal ions (Cr3+, Ni2+, Mn2+, etc.) [911] and rare-earth ions (Eu2+, Yb3+, etc.) [12,13] as activating ions have been well reported, e.g., (Sr,Ba)Y2O4:Eu2+ [14], Y3Al5-xGaxO12:Ni2+ [9], Li2ZnGe3O8:Mn2+ [11], AlP3O9:Cr3+ [15], LiScP2O7:Cr3+, Yb3+ [16], etc. Among the many activators, Cr3+ exhibits broadband absorption in the near-UV-Vis range with its unique 3d3 configuration, which can achieve tunable narrowband or broadband spectra from 600–1600 nm [17]. Among the many Cr3+-doped NIR phosphors, the garnet structure has been widely studied for its unique luminescence properties in the form of A3B2C3O12 [18]. However, in general, most of the phosphors with garnet structures as hosts exhibit narrow fluorescence emission ranges and full-width at half maximum (FWHM), such as CaLu2Mg2Si3O12:Cr3+ [19], Ca3Sc2Si3O12:Cr3+ [20], Ca4ZrGe3O12:Cr3+ [21], Ca2LuScGa2Ge2O12:Cr3+ [22], etc., and their emission ranges most of them do not exceed 650–1000 nm, and the FWHM does not exceed 180 nm. Garnet usually has a large band gap, which can inhibit the electron-phonon coupling (EPC) effect of Cr3+ to the host lattice, hence its emission spectrum shows a relatively narrower spectrum [23]. However, a wider spectral range can be obtained by screening garnet with a smaller band gap as a host. The luminescent properties of phosphors are determined by the electronic configuration of the activator ion, and by the coordination environment of the activator ion in the matrix [24]. The effect of the coordination environment on the luminescence properties of activator ions is mainly reflected in the nephelauxetic effect (NE) and crystal field splitting effect. The strength of the NE depends on the bonding properties of the activator ion and the ligand ion (the ratio between the ionic bonding content and the covalent bonding content) [25]. The smaller the electronegativity difference (ED) between the atoms of two elements, the stronger the covalent bonding and the stronger the NE, resulting in the redshift phenomenon of the spectrum.

Herein, through Cr3+ doping, the host band gap is indirectly reduced, so as to achieve the effect of broadband emission, a series of NIR broadband GZGG:Cr3+/Yb3+ phosphors were finally made. When x = 2%, y = 0.2%, the broadband emission of 650 nm–1200 nm was exhibited under 459 nm excitation with a FWHM of 260 nm. By co-doped Cr3+/Yb3+, the IQY increased from 47.15% to 54.9%, and at 150 °C the emission intensity remains at room temperature (RT) of 60%. Eventually, the obtained phosphor was fabricated into pc-LED for its application in night vision illumination and vascular visualization.

2. Materials and methods

Series of GZGG:xCr3+, yYb3+ phosphors were successfully synthesized by the high-temperature solid state reaction method. Gd2O3 (99.9%), ZnO (99%), Ga2O3 (99.999%), GeO2 (99.9999%), Cr2O3 (99%) and Yb2O3 (99.99%) were applied as raw material, weighed according to stoichiometric composition and no substrates in excess, then mixed and ground in an agate mortar. The mixture was then fired at 1350 °C for 5 h in air. After cooling, the mixture was ground to a powder using an agate mortar and then tested.

X-ray diffraction (XRD) patterns were recorded by a Bruker D8 Advance X-ray diffractometer at 40 KV and 40 mA with Cu-K-α (λ = 1.54056 Å) irradiation. Morphological characteristics were recorded by field emission scanning electron microscopy (Nova Nano SEM 450). Elemental composition and distribution were determined using an energy dispersive X-ray spectroscope attached to SEM. Rietveld refinement of the measured XRD data was performed using General Structure Analysis System (GSAS) software. Photoluminescence excitation (PLE) and photoluminescence emission (PL) spectra were recorded by a HORIBA FLuorolog-3 fluorescence spectrometer with a 450 W Xe lamp as the excitation source, the temperature spectra were tested with an external device, and the PL QYs were tested with Quanta-φ (HORIBA Scientific), using BaSO4 as a standard reference. The diffuse reflectance spectra (DRS) were measured by a Hitachi U-4100, using BaSO4 as a standard reference. Raman spectra were recorded with a HORIBA Raman spectrometer (LabRAM HR Evolution, Jobin Yvon) using a 532 nm Nd-YAG laser as the excitation source.

Pc-LED is a combination of 460 nm blue LED and GZGG:2%Cr3+. First, a certain amount of phosphor and silicone are fully mixed and stirred in a cup for 40 min to get the phosphors-silicone mixture coated on the blue diode chip. Then, the structure was baked in an oven at 150 °C for 1 h and then fixed on the substrate.

3. Results and discussion

GZGG belongs to a typical garnet structure with cubic crystal system $\mathrm{Ia\bar{3}d}$ space group, where Gd combines with O2- to form [GdO8] dodecahedra, Zn combines with O2- to form [ZnO6] hexahedron, Ga/Ge combines with O2- to form [Ga/GeO4] tetrahedron [26]. Figure 1(a) gives the XRD of GZGG:2%Cr3+, yYb3+ (y = 0–2%). By comparing the standard card PDF#04-007-8694, it is found that the prepared samples are pure phase. Figure 1(b) depicts the Rietveld refinement XRD pattern of GZGG and GZGG:2%Cr3+, 0.2%Yb3+ obtained by GASA software. The parameters Rp = 4.4%, Rwp = 5.9%, χ2 = 1.85% and Rp = 4.3%, Rwp = 5.6%, χ2 = 1.62% are within the confidence range, respectively. The Raman spectra of GZGG:xCr3+ revealed in Fig. 1(c), the intensity of Raman diffraction peaks decreases gradually with the doping of Cr3+, Yb3+. Figure 1(d) depicts the structure of GZGG, Cr3+ and Yb3+ generally occupy [ZnO6] six-coordination octahedron structure. Figure 1(e) portrays the SEM of GZGG:2%Cr3+, 0.2%Yb3+. The synthesized sample has a uniform particle distribution, smooth and flat surface, and a particle size of about 2 µm, which proves that the sample was synthesized very successfully.

 figure: Fig. 1.

Fig. 1. (a) X-ray diffraction pattern of GZGG:2%Cr3+, yYb3+ (y = 0-2%). (b) Rietveld refinement of GZGG and GZGG:2%Cr3+, 0.2%Yb3+. (c) Raman spectra of GZGG:xCr3+, yYb3+. (d) Structure of GZGG. (e) SEM image of GZGG:2%Cr3+, 0.2%Yb3+.

Download Full Size | PDF

Figure 2(a) shows the PLE spectra of GZGG:2%Cr3+ monitored at 855 nm and PL spectra monitored at 459 nm of GZGG:2%Cr3+, 0.2%Yb3+. The excitation spectra consisted of 4A24T1(4P) transition at 240 nm-380 nm, 4A24T1(4F) transition at 380 nm-570 nm and 4A24T2(4F) transition at 570 nm-750 nm [27]. Under the 459 nm excitation, the phosphor demonstrates an ultra-broadband NIR emission of 650-1200 nm with a FWHM of 218 nm, which is generated by the spin-allowed transition of 4T2(4F)→4A2 [28]. When co-doped Yb3+, the spectrum is widened to 260 nm. The PL spectra of GZGG:xCr3+ are gradually redshifted with increasing Cr3+ concentration and redshifted by 50 nm (Fig. 2(b)). The emission intensities reached a maximum at x = 2%, followed by concentration quenching and a slight broadening of the spectrum was obtained (see Supplement 1, Fig. S1).

 figure: Fig. 2.

Fig. 2. PLE and PL spectra of GZGG:2%Cr3+ and GZGG:2%Cr3+, 0.2%Yb3+. (b) PL spectra of GZGG:xCr3+. (c) Schematic diagram of the process of reducing electronegativity difference due to Cr3+ replacement of Zn2+. (d) Diffuse reflection spectrum of GZGG:2%Cr3+, 2%Yb3+ and PL spectrum of GZGG:2%Cr3+. (e) PL spectra of GZGG:2%Cr3+, yYb3+.

Download Full Size | PDF

Coordination chemistry theory is used to explain this redshift phenomenon. In Fig. 2(c), a schematic diagram of the process of Cr3+ substitution into the [ZnO6] six-coordination octahedron were displayed, where Cr3+ has an electronegativity of 1.66, which is bigger than that of Zn2+. And Cr3+ enters [ZnO6], replacing Zn2+ to reconstruct the [Zn/CrO6] polyhedron, and the previous Zn-O bond gradually turns into Cr-O bond, and the ED of Zn-O bond is 1.79, while the ED of Cr-O bond is 1.78, which is less than Zn-O bond, and both are bigger than 1.7, hence the ionic bond is formed, but the ionicity gradually decreases. With the doping of Cr3+, the ionicity of [ZnO6] gradually decreases, while the covalency gradually increases, leading to the bond length of Zn/Cr-O bond gradually becoming shorter and the NE getting enhanced, which eventually leads to the redshift phenomenon of the spectrum, and the shortening of the bond length also matches with the results obtained from the refinement (see Supplement 1, Fig. S2). At the same time, we also did a comparative experiment and chose Mg2+ as its central ion (R = 0.72 Å), which is close to the radius of Zn2+, and then doped with equal amount of Cr3+. It was found that the spectrum of Gd3Mg2GaGe2O12:2%Cr3+ (GMGG) was blue shifted toward the short band by 30 nm compared with GZGG:2%Cr3+, while the electronegativity of Mg was 1.31, and the ED of the Mg-O bond (2.13) was much larger than that of the ED of the Cr-O bond, hence this is the reason for the blueshift of the spectrum (see Supplement 1, Fig. S3a, b).

Figure S4a presents the DRS of GZGG:xCr3+ under the RT. The absorption gradually increases with the increase of concentration. When x = 2%, the calculated band gap is 3.63 eV, which is smaller than the band gap of GZGG (Eg = 4.15 eV). In Supplement 1, Fig. S4b, with the increase of concentration, the band gap gradually decreases. Smaller band gaps exhibit weaker ionic properties, which will lead to increased covalency in the octahedral coordination environment in which Cr3+ resides, resulting in increased the NE and ultimately causing redshift [29,30].

Figure 2(d) tested the DRS of GZGG:2%Cr3+, 2%Yb3+, and the absorption of Yb3+ appeared around 950 nm, which partially overlapped with the emission of GZGG:2%Cr3+, indicating that ET became possible. Figure 2(e) demonstrates the PL spectral of GZGG:2%Cr3+, yYb3+, which broadens to 260 nm under the excitation of 459 nm at y = 0.2%. With the increase of Yb3+ concentration, the emission intensity of Yb3+ gradually increases while the emission intensity of Cr3+ gradually decreases, indicating the ET from Cr3+ to Yb3+.

The thermal stability of the phosphors is extremely important to the pc-LED. Figure 3(a) depict the temperature-dependent PL spectra of GZGG:2%Cr3+ from 25 °C to 200 °C. With the increase of temperature, the emission intensity decreases gradually, while the spectral shape hardly changes, indicating that the crystal structure does not change with the change of temperature. At 100 °C, the emission intensity is 66.7% of 25 °C, and 41.2% can be maintained even at 150 °C (Fig. 3(b)). The Huang-Rhys factor (S) can reflect how strongly electrons couple to phonons and can be obtained by fitting the temperature- dependent full width at FWHM of photoluminescence peaks using the following equation [31]

$$FWHM = 2.36\sqrt S h\omega \sqrt {\coth \left( {\frac{{h\omega }}{{2kT}}} \right)} $$

Based on the temperature-dependent PL spectra, Fig. 3(c) demonstrates the fitting results of square of FWHM (FWHM2) as a function of 1/2kT, the result explains that the values of ħω and S are 0.0514 eV and 4.14, respectively. The relatively strong S indicates the strong EPC effect in GZGG:2%Cr3+. Stronger EPC effect results in a wider Cr3+ emission width, which in turn generally results in a more severe thermal quenching effect. In order to further improve the stability of GZGG:2%Cr3+ as well as to broaden its spectral range, Yb3+ was chosen to be co-doped with Cr3+. The temperature-dependent PL spectra of GZGG:2%Cr3+, 0.2%Yb3+ were next tested (Fig. 3(d)), and the intensity of the spectra gradually decreased with increasing temperature, and at 100 °C, the emission intensity was calculated by the integrated area to be about 77.9% of that at RT, while it was also maintained at 60% at 150 °C, and the stability was greatly improved (Fig. 3(e)). To better understand thermal quenching process, the activation energy (Ea) could be obtained by Arrhenius equation [32]

$${I_T} = \frac{{{I_0}}}{{1 + c\textrm{exp} \left( { - \frac{{Ea}}{{kT}}} \right)}}$$
where IT and I0 are the integrated intensity at certain temperature and RT, respectively. c represents a constant, k stands for the Boltzmann constant. Figure 3(f) displays the plotting of ln [(I0/It)−1] versus 1/(kT). The corresponding Ea was obtained as the slope of fitting line and equals to 0.304 eV.

 figure: Fig. 3.

Fig. 3. (a) Temperature-dependent PL spectra of GZGG:2%Cr3+. (b) PL intensity variation. (c) Fitting results of square of FWHM2 as a function of 1/2kT. (d) Temperature-dependent PL spectra of GZGG:2%Cr3+, 0.2%Yb3+. (e) The total integral intensity, peak intensity of Cr3+ and peak intensity of Yb3+ at different temperatures of GZGG:2%Cr3+, 0.2%Yb3+. (f) Plotting of ln[(I0/I)-1] versus 1/(kT) of optimal GZGG:2%Cr3+, 0.2%Yb3+.

Download Full Size | PDF

The QY of the phosphor was improved by Cr3+/Yb3+ co-doping. In Supplement 1, Fig. S5, the fluorescence QYs as well as the absorption efficiency of GZGG:2%Cr3+ and GZGG:2%Cr3+, 0.2%Yb3+ were tested, respectively. The IQY increased from 47.15% to 54.9% and the EQY also increased from 22.77% to 24.65%. Through the way of ET, the QY of phosphors is improved, which will provide a theoretical basis for the application of phosphors.

To verify the application value of the phosphor, it was packaged with a blue light chip as pc-LED. Figure 4(a) portrays the PL spectrum of the pc-LED at RT, the spectrum covers the emission range of 750-1200 nm, and in the inset pictures of pc-LED without and with the current. The temperature of the pc-LED at different operating currents is also recorded in Fig. 4(b). With the increase of the operating current, the temperature of pc-LED gradually increases and reaches 89.3 °C at 300 mA of current. Due to the unique spectral characteristics of the NIR light source, including invisibility to the naked eye, excellent penetration capability, and low biological tissue-related absorption and scattering, the manufactured pc-LED can be used for night vision surveillance and biological tissue penetration applications. Figure 4(c) displays a photograph taken by a NIR camera using a pc-LED as a light source penetrating a finger in darkness, where the blood vessels of the finger can be clearly seen. The encapsulated pc-LED was simulated for nighttime illumination, and the glass bottle was photographed at 0.5 m using the NIR pc-LED as the light source, and a clearly visible photograph was obtained using the NIR camera, indicating the commercial value of the manufactured phosphor for night vision.

 figure: Fig. 4.

Fig. 4. (a) The output spectrum of fabricated NIR pc-LED packaged by combining a 460 nm chip with GZGG:Cr3+/Yb3+. The inset shows the fabricated LED devices. (b) Thermal images of the pc-LED at different operating currents over 0-300 mA. (c) Photograph of human fingers, glass stopper, and glass bottle after irradiating by the NIR pc-LED.

Download Full Size | PDF

4. Conclusions

A series of efficient and stable NIR broadband garnet GZGG:Cr3+/Yb3+ phosphors were synthesized by actively screening of host with smaller band gaps and selection of suitable electronegativity of the central coordination ions in combination with NE. Through the Cr3+/Yb3+ co-doping, GZGG: Cr3+/Yb3+ exhibits broadband emission of 650-1200 nm with an FWHM of 260 nm, while the IQY and EQY of the phosphors are also well enhanced, and reach 54.9% and 24.65%, respectively. The luminescence stability was also greatly enhanced by NE, and at 150 °C, the emission intensity was 60% of that at RT. The produced phosphors were combined with a blue light chip to fabricate pc-LED and tested for their operating temperature at different operating currents. The phosphors were used for vascular visualization and night vision, and the experimental results revealed their application prospects.

Funding

Central government to guide local scientific and Technological Development (206Z1102G, 216Z1101G); Personnel training project of Hebei Province (A201902005); Natural Science Foundation of Hebei Province (E2019201223); National Natural Science Foundation of China (51902080).

Disclosures

The authors declare that they have no conflict of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. E. J. Marques, S. T. de Freitas, M. F. Pimentel, and C. Pasquini, “Rapid and non-destructive determination of quality parameters in the ‘Tommy Atkins’ mango using a novel handheld near infrared spectrometer,” Food Chem. 197(Part B), 1207–1214 (2016). [CrossRef]  

2. M.-H. Fang, G. N. A. De Guzman, Z. Bao, N. Majewska, S. Mahlik, M. Grinberg, G. Leniec, S. M. Kaczmarek, C.-W. Yang, K.-M. Lu, H.-S. Sheu, S.-F. Hu, and R.-S. Liu, “Ultra-high-efficiency near-infrared Ga2O3:Cr3+ phosphor and controlling of phytochrome,” J. Mater. Chem. C 8(32), 11013–11017 (2020). [CrossRef]  

3. W. Wang and J. Paliwal, “Near-infrared spectroscopy and imaging in food quality and safety,” Sens. & Instrumen. Food Qual. 1(4), 193–207 (2007). [CrossRef]  

4. J. Zhong, Y. Zhuo, F. Du, H. Zhang, W. Zhao, and J. Brgoch, “Efficient and tunable luminescence in Ga2-xInxO3:Cr3+ for near-infrared imaging,” ACS Appl. Mater. Interfaces 13(27), 31835–31842 (2021). [CrossRef]  

5. Y. Wang, Z. Wang, G. Wei, Y. Yang, S. He, J. Li, Y. Shi, R. Li, J. Zhang, and P. Li, “Ultra-Broadband and high efficiency Near-Infrared Gd3ZnxGa5-2xGexO12:Cr3+ (x = 0-2.0) garnet phosphors via crystal field engineering,” Chem. Eng. J. 437(1), 135346 (2022). [CrossRef]  

6. L. Zhang, D. Wang, Z. Hao, X. Zhang, G. H. Pan, H. Wu, and J. Zhang, “Cr3+-doped broadband NIR garnet phosphor with enhanced luminescence and its application in NIR spectroscopy,” Adv. Opt. Mater. 7(12), 1900185 (2019). [CrossRef]  

7. A. Zampetti, A. Minotto, and F. Cacialli, “Near-infrared (NIR) organic light-emitting diodes (OLEDs): challenges and opportunities,” Adv. Funct. Mater. 29(21), 1807623 (2019). [CrossRef]  

8. D. Hayashi, A. M. van Dongen, J. Boerekamp, S. Spoor, G. Lucassen, and J. Schleipen, “A broadband LED source in visible to short-wave-infrared wavelengths for spectral tumor diagnostics,” Appl. Phys. Lett. 110(23), 233701 (2017). [CrossRef]  

9. L. Yuan, Y. Jin, D. Zhu, Z. Mou, G. Xie, and Y. Hu, “Ni2+-doped yttrium aluminum gallium garnet phosphors: Bandgap engineering for broad-band wavelength-tunable shortwave-infrared long-persistent luminescence and photochromism,” ACS Sustainable Chem. Eng. 8(16), 6543–6550 (2020). [CrossRef]  

10. J. Xu, J. Ueda, Y. Zhuang, B. Viana, and S. Tanabe, “Y3Al5-xGaxO12:Cr3+:A novel red persistent phosphor with high brightness,” Appl. Phys. Express 8(4), 042602 (2015). [CrossRef]  

11. J. Cheng, P. Li, Z. Wang, Y. Sun, Q. Bai, Z. Li, M. Tian, C. Wang, and Z. and Yang, “Synthesis, structure and luminescence properties of novel NIR luminescent materials Li2ZnGe3O8:xMn2+,” J. Mater. Chem. C 5(1), 127–133 (2017). [CrossRef]  

12. J. Qiao, G. Zhou, Y. Zhou, Q. Zhang, and Z. Xia, “Divalent europium-doped near-infrared-emitting phosphor for light-emitting diodes,” Nat. Commun. 10(1), 5267 (2019). [CrossRef]  

13. E. T. Basore, H. Wu, W. Xiao, G. Zheng, X. Liu, and J. Qiu, “High-power broadband NIR LEDs enabled by highly efficient blue-to-NIR conversion,” Adv. Opt. Mater. 9(7), 2001660 (2021). [CrossRef]  

14. Z. Yang, Y. Zhao, Y. Zhou, J. Qiao, Y. C. Chuang, M. S. Molokeev, and Z. Xia, “Giant red-shifted emission in (Sr,Ba)Y2O4:Eu2+ phosphor toward broadband near-infrared luminescence,” Adv. Funct. Mater. 32(1), 2103927 (2022). [CrossRef]  

15. D. Huang, X. He, J. Zhang, J. Hu, S. Liang, D. Chen, K. Xu, and H. Zhu, “Efficient and thermally stable broadband near-infrared emission from near zero thermal expansion AlP3O9:Cr3+ phosphors,” Inorg. Chem. Front. 9(8), 1692–1700 (2022). [CrossRef]  

16. L. Yao, Q. Shao, S. Han, C. Liang, J. He, and J. Jiang, “Enhancing near-infrared photoluminescence intensity and spectral properties in Yb3+ codoped LiScP2O7:Cr3+,” Chem. Mater. 32(6), 2430–2439 (2020). [CrossRef]  

17. L. You, R. Tian, T. Zhou, and R.-J. Xie, “Broadband near-infrared phosphor BaMgAl10O17:Cr3+ realized by crystallographic site engineering,” Chem. Eng. J. 417(1), 129224 (2021). [CrossRef]  

18. D. Huang, S. Liang, D. Chen, J. Hu, K. Xu, and H. Zhu, “An efficient garnet-structured Na3Al2Li3F12:Cr3+ phosphor with excellent photoluminescence thermal stability for near-infrared LEDs,” Chem. Eng. J. 426(15), 131332 (2021). [CrossRef]  

19. R. Li, Y. Liu, C. Yuan, G. Leniec, L. Miao, P. Sun, Z. Liu, Z. Luo, R. Dong, and J. Jiang, “Thermally stable CaLu2Mg2Si3O12:Cr3+ phosphors for NIR LEDs,” Adv. Opt. Mater. 9(16), 2100388 (2021). [CrossRef]  

20. Z. Jia, C. Yuan, Y. Liu, X. J. Wang, P. Sun, L. Wang, H. Jiang, and J. Jiang, “Strategies to approach high performance in Cr3+-doped phosphors for high-power NIR-LED light sources,” Light: Sci. Appl. 9(1), 86 (2020). [CrossRef]  

21. J. Xiang, J. Zheng, X. Zhao, X. Zhou, C. Chen, M. Jin, and C. Guo, “Synthesis of broadband NIR garnet phosphor Ca4ZrGe3O12:Cr3+, Yb3+ for NIR pc-LED applications,” Mater. Chem. Front. 6(4), 440–449 (2022). [CrossRef]  

22. B. Bai, P. Dang, D. Huang, H. Lian, and J. Lin, “Broadband near-infrared emitting Ca2LuScGa2Ge2O12:Cr3+ phosphors: Luminescence properties and application in light-emitting diodes,” Inorg. Chem. 59(18), 13481–13488 (2020). [CrossRef]  

23. F. Zhao, Z. Song, J. Zhao, and Q. Liu, “Double perovskite Cs2AgInCl6:Cr3+: broadband and near-infrared luminescent materials,” Inorg. Chem. Front. 6(12), 3621–3628 (2019). [CrossRef]  

24. H. Ji, “Analysis of site-occupation of activator in phosphors,” Chin. J. Lumin. 43(01), 26–41 (2022). [CrossRef]  

25. L. Wang, R. J. Xie, T. Suehiro, T. Takeda, and N. Hirosaki, “Down-conversion nitride materials for solid state lighting: Recent advances and perspectives,” Chem. Rev. 118(4), 1951–2009 (2018). [CrossRef]  

26. X. Zou, X. Wang, H. Zhang, Y. Kang, X. Yang, X. Zhang, M. S. Molokeev, and B. Lei, “A highly efficient and suitable spectral profile Cr3+-doped garnet near-infrared emitting phosphor for regulating photomorphogenesis of plants,” Chem. Eng. J. 428(15), 132003 (2022). [CrossRef]  

27. X. Xu, Q. Shao, L. Yao, Y. Dong, and J. Jiang, “Highly efficient and thermally stable Cr3+-activated silicate phosphors for broadband near-infrared LED applications,” Chem. Eng. J. 383(1), 123108 (2020). [CrossRef]  

28. S. Miao, Y. Liang, Y. Zhang, D. Chen, and X. J. Wang, “Broadband short-wave infrared light-emitting diodes based on Cr3+-doped LiScGeO4 phosphor,” ACS Appl. Mater. Interfaces 13(30), 36011–36019 (2021). [CrossRef]  

29. Y. Su, L. Yuan, H. Liu, G. Xiong, H. Wu, Y. Hu, X. Cheng, and Y. Jin, “Multi-site occupation of Cr3+ toward developing broadband near-infrared phosphors,” Ceram. Int. 47(16), 23558–23563 (2021). [CrossRef]  

30. W. Nie, L. Yao, G. Chen, S. Wu, Z. Liao, L. Han, and X. Ye, “A novel Cr3+-doped Lu2CaMg2Si3O12 garnet phosphor with broadband emission for near-infrared applications,” Dalton Trans. 50(24), 8446–8456 (2021). [CrossRef]  

31. E. Song, H. Ming, Y. Zhou, F. He, J. Wu, Z. Xia, and Q. Zhang, “Cr3+-doped Sc-based fluoride enabling highly efficient near infrared luminescence: A case study of K2NaScF6:Cr3+,” Laser Photonics Rev. 15(2), 2000410 (2021). [CrossRef]  

32. Q. Wu, P. Li, Z. Ye, X. Huo, H. Yang, Y. Wang, D. Wang, J. Zhao, Z. Yang, and Z. Wang, “Near-infrared emitting phosphor LaMg0.5(SnGe)0.5O3:Cr3+ for plant growth applications: Crystal structure, luminescence, and thermal stability,” Inorg. Chem. 60(21), 16593–16603 (2021). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supporting information

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

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

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. (a) X-ray diffraction pattern of GZGG:2%Cr3+, yYb3+ (y = 0-2%). (b) Rietveld refinement of GZGG and GZGG:2%Cr3+, 0.2%Yb3+. (c) Raman spectra of GZGG:xCr3+, yYb3+. (d) Structure of GZGG. (e) SEM image of GZGG:2%Cr3+, 0.2%Yb3+.
Fig. 2.
Fig. 2. PLE and PL spectra of GZGG:2%Cr3+ and GZGG:2%Cr3+, 0.2%Yb3+. (b) PL spectra of GZGG:xCr3+. (c) Schematic diagram of the process of reducing electronegativity difference due to Cr3+ replacement of Zn2+. (d) Diffuse reflection spectrum of GZGG:2%Cr3+, 2%Yb3+ and PL spectrum of GZGG:2%Cr3+. (e) PL spectra of GZGG:2%Cr3+, yYb3+.
Fig. 3.
Fig. 3. (a) Temperature-dependent PL spectra of GZGG:2%Cr3+. (b) PL intensity variation. (c) Fitting results of square of FWHM2 as a function of 1/2kT. (d) Temperature-dependent PL spectra of GZGG:2%Cr3+, 0.2%Yb3+. (e) The total integral intensity, peak intensity of Cr3+ and peak intensity of Yb3+ at different temperatures of GZGG:2%Cr3+, 0.2%Yb3+. (f) Plotting of ln[(I0/I)-1] versus 1/(kT) of optimal GZGG:2%Cr3+, 0.2%Yb3+.
Fig. 4.
Fig. 4. (a) The output spectrum of fabricated NIR pc-LED packaged by combining a 460 nm chip with GZGG:Cr3+/Yb3+. The inset shows the fabricated LED devices. (b) Thermal images of the pc-LED at different operating currents over 0-300 mA. (c) Photograph of human fingers, glass stopper, and glass bottle after irradiating by the NIR pc-LED.

Equations (2)

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

F W H M = 2.36 S h ω coth ( h ω 2 k T )
I T = I 0 1 + c exp ( E a k T )
Select as filters


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