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

Study on the mechanism of high energy transfer efficiency of blue light excited Cr3+, Nd3+ co-doped near infrared phosphors

Open Access Open Access

Abstract

Although Cr3+ as activator for Near infrared (NIR) phosphors has been widely studied, the peaks of Cr3+ emission spectra in most hosts are less than 1000 nm. Nd3+ as an activator in many hosts has a wide distribution of absorption peaks in the Ultraviolet-visible-Near infrared (UV-vis-NIR) band, especially in the 650-900 nm band for effective NIR to NIR Stokes luminescence (4F3/24I9/2, 4F3/24I11/2 transitions). Therefore, Cr3+, Nd3+ co-doping to achieve the emission in the NIR II region (1000-1700nm) is very meaningful. Here, we report La2CaZrO6(LCZO): Cr3+, Nd3+ NIR phosphors with emission spectra covering an ultra-wide range of 700-1400 nm and reveal their luminescence mechanism. The energy transfer efficiency of Cr3+ for Nd3+ can be as high as 88.4% under 471 nm blue light excitation. In the same case, the integrated intensity of the emission spectra of Cr3+, Nd3+ co-doped can reach 847% of that of Nd3+ alone and 204% of that of Cr3+ alone. Finally, the combination of commercial blue light chips and Cr3+, Nd3+ co-doped NIR phosphors shows great potential for applications in face recognition, night lighting, and angiography.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

NIR spectroscopy is a non-destructive, fast and convenient technique for detection and analysis [13]. In recent years, the development and application of new NIR light sources are becoming more and more abundant. With the increasing maturity of NIR phosphor-converted light emitting diodes (pc-LEDs) technology, the advantages of energy saving, environmental protection and miniaturization are gradually emerging [46]. Trivalent chromium ion (Cr3+) as the activation center is the mainstream choice for NIR phosphors recently, but in most hosts, the main peak value of its emission spectrum is less than 1000 nm [712]. It is well worthwhile to develop NIR phosphors with longer emission wavelengths. Since Nd3+ has a wide distribution of absorption peaks in the UV-vis-NIR band, it can achieve energy transfer with different ions [13]. Nd3+ co-doped phosphors with Bi3+, Ce3+, Eu3+, Ho3+, Mn4+, and Yb3+, respectively, have been investigated [1419].

Due to the overlap between the emission spectrum of Cr3+ and part of the absorption band of Nd3+, the doping of Cr3+ and Nd3+ to achieve NIR luminescence with new characteristics is a novel direction. Viktor Anselm et al. prepared SrGa12O19:Cr3+, Nd3+ phosphors and investigated the energy transfer and thermal quenching behavior [20]. Yiling Wu et al. developed La3Ga5GeO14: Cr3+, Nd3+ phosphors with NIR long persistent phosphorescence and NIR-to-NIR Stokes luminescence [21]. Abbi L. Mullins et al. synthesized LaGaO3:Cr3+, Nd3+ phosphors with high relative sensitivity by detecting the characteristic emission of Cr3+, Nd3+ at different temperatures, which can be used in the field of temperature measurement [22]. Nevertheless, the energy transfer mechanism between Cr3+ and Nd3+, Nd3+ and Nd3+ and the mutual influence in the system with Cr3+, Nd3+ co-doping as activator is still obscure.

In this work, we have synthesized La2CaZrO6:Cr3+, Nd3+ phosphors by high-temperature solid-phase method. The occupancy of Nd3+ in LCZO crystals is discussed in detail by first-principles calculations and Rietveld refinement. The X-ray photoelectron spectroscopy (XPS) shows that trivalent neodymium (Nd3+) ions are present in LCZO crystals as the main valence state, and the morphological changes of LCZO: Nd3+/Cr3+ are also analyzed. The emission spectra of LCZO:Cr3+, Nd3+ under 471 nm excitation cover an ultra-wide range of 700-1400 nm, with three distinct emission peaks at 907 nm, 1084 nm, and 1373 nm corresponding to the 4F3/24I9/2, 4F3/24I11/2, and 4F3/24I13/2 transitions of Nd3+, respectively. The concentration-dependent emission spectra of 4F3/24I9/2, 4F3/24I11/2 and 4F3/24I13/2 electronic transitions of Nd3+ ions are observed, and the concentration dependence of LCZO:Nd3+ spectra are discussed to reveal the energy transfer mechanism between Nd3+ ions. In addition, the UV–vis–NIR diffuse reflectance spectra, Kubelka-Munk (K-M) and Urbach tail theories and fluorescence lifetime analysis elucidate the interaction between Cr3+, Nd3+ and their effect on LCZO crystals. Further, the energy transfer mechanism between Cr3+ and Nd3+, Nd3+ and Nd3+ is revealed in the Cr3+, Nd3+ co-doped system. This study provides insights into the development of novel Nd3+ emission-dominated NIR phosphors and demonstrates their promising applications in several fields.

2. Experimental section

2.1. Materials and synthesis

The samples La2CaZrO6:Cr3+, Nd3+ are synthesized using the conventional high-temperature solid-phase method. Lanthanum oxide (La2O3, 99.99%, Aladdin), calcium carbonate (CaCO3, A.R.), zirconium oxide (ZrO2, 99%, Aladdin), chromium oxide (Cr2O3, 99. 95%, Aladdin), neodymium oxide (Nd2O3, 99%, Aladdin) were weighed in proportion to their chemical formula as raw materials. The raw materials were sintered at 1500°C for 8 h under a reducing atmosphere of H2 (10%) and N2 (90%) to prevent the oxidation of Cr3+. After cooling to room temperature, the samples were reground into powder for study.

2.2. Characterization and measurements

X-ray powder diffraction (XRD) were measured on a D8 X-ray diffractometer with nickel-filtered Cu Kα radiation (λ = 1.54056 Å) at 40 mA, 40 kV. The Rietveld refinement of the phase structure of the samples was done using FullProf software. The fluorescence lifetime, photoluminescence excitation (PLE) and photoluminescence (PL) spectra of the phosphors were tested with an FLS980 transient fluorescence spectrometer (Edinburgh), which uses a 450 W Xe lamp as the excitation source and an external temperature controller to test the variable high temperature spectra. The diffuse reflectance spectra (DRS) of the samples were tested using Hitachi U-4100 UV–vis–NIR spectroscopy, utilizing BaSO4 as a standard reference. The X-ray photoelectron spectra (XPS) of the samples were measured by ESCALAB 250Xi. The surface morphology of the phosphor was measured by field emission scanning electron microscopy (SEM, Nova NanoSEM450). The elemental distribution of the phosphor was tested using a high-resolution transmission electron microscope (HRTEM, JEPL-2100Plus), and the crystal plane spacing was calculated based on the obtained data using Gatan Digital Micrograph software. The external quantum yield (EQY) was measured by the Edinburgh FLS-1000. The pc-LED is prepared by using 470 nm blue light chip and phosphor LCZO:2%Cr3+, 1%Nd3+.

2.3 Computational details

All calculations were performed with the CASTEP code [23,24]. The relaxation and formation energy (Eform) of LCZO was calculated using generalized density functional theory and generalized gradient modified Perdew-Burke-Ernzerhof (PBE) [25]. All structures were converged to 0.01 eV/Å.

3. Results and discussion

3.1. Phase and morphology analysis of phosphors

The double perovskite oxide LCZO belongs to the monoclinic crystal system with the space group P21/n as shown in Fig. 1(a). Under the condition of charge conservation, Nd3+ (coordination number CN = 8, 1.109 Å) tends to replace La3+ (CN = 8, 1.16 Å) due to their similar ionic radii. The formation energy for the three possible occupancies of Nd3+ (replacing La or Ca or Zr) is calculated using the following equation [26,27]:

$${E_f} = {E_{LCZO:Nd}} - {E_{LCZO}} - \; {\mu _{Nd}} + \; {\mu _s}$$
where ELCZO: Nd denotes the total energy of LCZO: Nd3+, ELCZO represents the total energy of LCZO, and µNd and µS imply the chemical potentials of the Nd atom and the substituted atom, respectively. The results in Table 1 show that Nd3+ is most likely to replace La3+ due to its minimal formation energy (-8.072 eV).

 figure: Fig. 1.

Fig. 1. (a) Crystal structure of LCZO. (b) Cr3+, Nd3+ substitution schematic. (c) Rietveld refinement of XRD of LCZO: 2% Cr3+, 4% Nd3+. (d), (e) XRD patterns of LCZO: x Nd3+ (x = 0.1%∼8%). (f), (g) XRD patterns of LCZO: 2% Cr3+, y Nd3+ (y = 0%∼8%).

Download Full Size | PDF

Tables Icon

Table 1. The calculated formation energies (Ef) of Nd3+ occupying different sites in LCZO

Cr3+ ions tend to co-substitute Ca2+ and Zr4+ because Cr3 + (0.615 Å) and Zr4 + (0.72 Å) have similar radii and must satisfy the condition of charge conservation. The schematic diagram of Cr3+, Nd3+ substitution and the mechanism of their interactions are given in Fig. 1(b). Further, Fig. 1(c) depicts the Rietveld refinement results of the XRD of LCZO: 2% Cr3+, 4% Nd3+. The refinement parameters (Rp = 5.50%, Rwp = 9.54%, and Rexp = 3.70%) are less than 10%, which ensures the confidence of the refinement results. The atomic occupation information obtained from the XRD refinement results is shown in Table 2, which indicates that the Nd3+ and La3+ ions occupy the same lattice sites, and the lattice sites of the two Cr3+ ions match with Ca2+ and Zr4+, respectively.

Tables Icon

Table 2. Atomic occupancy information of LCZO: 2% Cr3+, 4% Nd3+

Fig. 1(d),1(f) show the XRD patterns of LCZO: x Nd3+ (x = 0.1%∼8%) and LCZO: 2% Cr3+, y Nd3+ (y = 0%∼8%), respectively, and the results show good synthesis compared with the calculated reference XRD and refinement results. As shown in Fig. 1(e), 1(g), Nd3+ doping in LCZO does not cause a significant shift in the XRD peak due to the close ionic radius of Nd3+ and La3+.

In addition, Fig. S1 shows the XRD patterns of LCZO: 4% Nd3+, zCr3+ (z = 0 ∼ 6%). With the increase of Cr3+ concentration, the position of the XRD peak shifts considerably toward the large angle direction (Fig. S1(b)), which implies that the crystal plane spacing will become smaller.

The XPS spectra of LCZO: 2% Cr3+, 4% Nd3+ samples are plotted in Fig. 2(a), where the peaks of elements La, Ca, Zr, O, Cr and Nd could be identified. The two peaks at 576.3 eV and 586.3 eV correspond to 2p3/2 and 2p1/2 of the Cr3+ ion, respectively, as shown in the inset in the upper left corner of Fig. 2(a) [28]. And, the peak is located at 982.9 eV corresponding to the major 3d5/2 orbital of the Nd3+ ion as shown in the upper right inset of Fig. 2(a) [29].

 figure: Fig. 2.

Fig. 2. (a) XPS spectra of LCZO: 2% Cr3+, 4%Nd3+. (b), (c) The HRTEM image of phosphor LCZO:2% Cr3+, 4% Nd3+ particles and elemental distribution maps.

Download Full Size | PDF

The above results show that Cr3+, Nd3+ exist in phosphor LCZO as the main valence state. Furthermore, Fig. S2 shows the SEM images of phosphors LCZO: 2%Nd3+, 2%Cr3+, which shows that the phosphors have irregular particle shapes but the samples are relatively uniform in size. The size distribution of the phosphor LCZO: 2%Nd3+, 2%Cr3+ samples ranged from 0-30 µm, with an average particle size of 9.0 µm. Fig. 2(b) shows the high-resolution TEM (HR-TEM) images of LCZO: 2% Cr3+, 4% Nd3+ samples, the results of which show (221) crystal plane spacing of 0.139 nm. Fig. 2(c) shows the elemental distribution of LCZO: 2% Cr3+, 4% Nd3+ phosphor particles, in which La, Ca, Zr, O, Cr and Nd are uniformly distributed.

3.2. Photoluminescence properties and energy transfer mechanism of Cr3+, Nd3+ doped in LCZO

Fig. 3(a) depicts the UV-vis-NIR diffuse reflectance spectra of LCZO:Nd3+ in the range of 320∼1200 nm, where the absorption peak of Nd3+ at 360 nm corresponds to the 4I9/24D3/2 transition, 471 nm to the 4I9/24G9/2 transition, 530 nm to the 4I9/24G7/2 transition, 590 nm to the 4I9/24G5/2 transition, 687 nm to the4I9/24F9/2 transition, 750 nm to the 4I9/24F7/2 transition, 808 nm to the 4I9/24F5/2 transition, and 880 nm corresponds to 4I9/24F3/2 transition(Fig. 3(a), 3(b)). The absorption peaks gradually rise with the increase of Nd3+ concentration. Further, Fig. 3(b) also depicts the diffuse reflectance spectra of LCZO: 2% Cr3+, y Nd3+ (y = 0 ∼ 8%), where the absorption peaks of both Cr3+, Nd3+ are present. The diffuse reflectance spectra of LCZO:Cr3+ show absorption peaks at 471 nm and 687 nm corresponding to the 4A2(F)→4T1(F) and 4A2(F)→4T2(F) transitions of Cr3+, respectively. With the increase of Nd3+ concentration, the intensity of Cr3+ dominated absorption bands (4A2(F)→4T1(F) and4A2(F)→4T2(F) transitions) remain basically unchanged, while the Nd3+ dominated absorption bands are gradually enhanced on the basis of Cr3+.

 figure: Fig. 3.

Fig. 3. (a) Diffuse reflectance spectra of LCZO: x Nd3+ (x = 0 ∼ 8%). (b) Diffuse reflectance spectra of LCZO: 2%Cr3+, y Nd3+ (y = 0% ∼ 8%). (c) PL and PLE spectra of LCZO: Cr3+, LCZO: Nd3+. (d) PL and PLE spectra of LCZO: 2%Cr3+, 0.5%Nd3+.

Download Full Size | PDF

The optical band gap can be calculated using UV-vis-NIR absorption spectral data according to the Kubelka-Munk equations [3032]:

$$F(R )= \; \frac{{{{({1 - R} )}^2}}}{{2R}}$$
$${({F(R )h\nu } )^n} = A({h\nu - {E_g}} )$$
where A, hv, and Eg denote the absorption constant, photon energy, and optical band gap, respectively, n = 2 for direct transition, 1/2 for indirect transition, R represents the reflection coefficient, and F(R) is the absorption coefficient. Fig. S3 shows that LCZO is a direct band gap material. Fig. S4 (a) shows that the optical band gap Eg of LCZO: x Nd3+ (x = 0%∼8%) decreases from 3.59 eV to 3.35 eV with increasing Nd3+ concentration, which is consistent with the variation of LCZO: 2% Cr3+, y Nd3+ (y = 0%∼8%) (Fig. S4(b)). This reduction in the band gap can be attributed to the appearance of new unoccupied electronic states in the gap below the conduction band edge due to the substitution of Nd3+ ions for La3+. The introduction of Cr3+ significantly enhances the absorption coefficients of Nd3+ at various energy levels as shown in the Kubelka-Munk (K-M) absorption spectra (Fig. S4(a), S4(b)).

Fig. 3(c) shows the excitation and emission spectra of LCZO: Cr3+ and LCZO: Nd3+ simultaneously. Monitoring 861 nm, the excitation spectrum of phosphor LCZO:Cr3+ shows peaks of 471 nm, 651 nm corresponding to the 4A2(F)→4T1(F), 4A2(F)→4T2(F) transitions of Cr3+, respectively. Moreover, the LCZO: Cr3+ shows an ultra-broadband emission spectrum covering 700 to 1400 nm under 471 nm excitation. In addition, the emission peaks of LCZO: Nd3+ at 907 nm, 1084 nm and 1373 nm correspond to the electronic transitions of 4F3/24I9/2, 4F3/24I11/2 and 4F3/24I13/2 of Nd3+ at 822 nm excitation, respectively. The 4f electrons of Nd3+ are shielded by the outer 5s2 and 5p6 electrons, and the overall energy levels of Nd3+ ions do not change greatly in different hosts, but the individual energy levels of Nd3+ can be divided into 2J + 1 sub-energy levels by the surrounding local crystal field. Compared with previous reports (Table S1), the wavelengths of the emission peaks at 1084 nm and 1373 nm corresponding to 4F3/24I11/2 and 4F3/24I13/2 of LCZO:Nd3+ are longer in the case of high concentration of Nd3+ doping (discussed in the later section).

Monitoring 1084 nm, a series of excitation peaks with 880 nm as the strongest peak can be observed, among which 880 nm, 894 nm, 902 nm and 915 nm correspond to the 4I9/24F3/2 transition of Nd3+. The excitation spectrum of Nd3+ and the emission spectrum of Cr3+ effectively overlap, which provides the possibility of energy transfer from Cr3+ to Nd3+.

Fig. 3(d) depicts the PL and PLE spectra of phosphor LCZO: 2% Cr3+, 0.5% Nd3+. The emission spectrum of LCZO: 2% Cr3+, 0.5% Nd3+ is dominated by the emission of Nd3+ at 471 nm excitation, which indicates that the energy transfer between Cr3+, Nd3+ occurs. The excitation spectrum of LCZO:2%Cr3+, 0.5%Nd3+ at 500∼1000 nm is obtained by monitoring at 1084 nm, which is basically consistent with the excitation spectrum of LCZO: Nd3+. The excitation spectrum of LCZO: 2% Cr3+, 0.5% Nd3+ at 890 nm is monitored and shows the characteristic excitation peaks of Nd3+ at 360 nm, 530 nm, 590 nm, 750 nm, 808 nm for Nd3+ and 471 nm for Cr3+, which can be perfectly matched with the absorption peaks in the DRS spectra.

Fig. 4 exhibits a schematic diagram of the energy levels of Nd3+ in LCZO, where detailed information can be seen in the fine emission spectrum of LCZO: 4% Nd3+ labeled in Fig. S5 and Table S2. The 695 nm, 890 nm, and 907 nm peaks in the emission spectra of LCZO: Nd3+ correspond to the 4F9/24I9/2 transition, 4F3/2(R1)→4I9/2(Z2) transition, and 4F3/2(R1)→4I9/2(Z4) transition, respectively. Fig. 5(a) shows the emission spectra of LCZO: x Nd3+ (x = 0.1 ∼ 8%) under 471 nm excitation. The integrated intensity of the emission spectra of LCZO: x Nd3+ increases with the concentration of Nd3+ and reaches a maximum at x = 6%. Further, the effect of Nd3+ concentration on the intensity of the electronic transitions from 4F3/24IJ/J’ can be expressed in terms of the fluorescence intensity branching ratio βJ/J' [13]:

$$\beta J/J' = \; \frac{{{A_{J - J'}}}}{{\mathop \sum \nolimits_{A_{J - J'}}}}$$
where AJ-J’ denotes the emission intensity of the 4F3/24IJ/J’ transition, and the denominator is the sum of the emission intensities of the 4F3/2 radiation reduction to all energy levels. Fig. 5(b) depicts the effect of concentration on the branching ratio of fluorescence intensity, where β9/2 shows a decreasing trend, β11/2 shows an increasing trend, while β13/2 remains essentially unchanged. The rapid diffusion of the excitation energy corresponding to the resonant 4F3/2 energy levels between the Nd3+ ions (Purple arrow in process (a) of Figure 4) leads to the variations in the emission intensity branching ratio. Fig. S6 (a), (c), (e) show the variation of the relative intensity of LCZO: xNd3+ (x = 0.1% ∼ 8%) at 907 nm (I907) and 890 nm (I890), at 1084 nm (I1084) and 1070 nm (I1070), and at 1373 nm and 1348 nm. The results in Fig. S6(b), (d), (f) reveal that the emission intensity ratios (I907/I890, I1084/I1070, I1373/I1348) increase with the rise of Nd3+ concentration. The relative intensities of the short wavelength portions of the 4F3/24I9/2, 4F3/24I11/2, 4F3/24I13/2 transition bands become weaker with increasing Nd3+ doping concentration, which can be attributed to the enhanced cross-relaxation {4F3/2,4I9/2}⇔{4I15/2,4I15/2} of Nd3 + (blue arrow in process (b) of Figure 4) [33]. In addition, the re-absorption process between Nd3+ ions may also cause changes in the emission intensity branching ratio due to the relatively high absorption cross section of 4I9/24F3/2 (Fig. 3(a)) and the overlap between the excitation spectrum monitored at 1084 nm (4F3/24I11/2) and the emission spectrum produced by 4F3/24I9/2 (Fig. 3(c)).

 figure: Fig. 4.

Fig. 4. Schematic diagram of the energy levels of Nd3+ in LCZO.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. (a) Emission spectra of LCZO: x Nd3+ (x = 0.1 ∼ 8%). (b) Branching ratio βJ/J’ as a function of concentration of Nd3+. (c) Emission spectra of LCZO: 2%Cr3+, y Nd3+ (y = 0%∼8%). (d) Fitted curves of Is0/Is versus Cn/3(n = 6, 8, 10) associated with LCZO: 2% Cr3+, y Nd3+ (y = 0% ∼ 8%).

Download Full Size | PDF

Fig. S7 shows the emission spectra of LCZO: 4% Nd3+, zCr3+ (z = 0∼6%) under 471 nm excitation. The results show that the introduction of Cr3+ does not change the shape and peak of the emission spectra, and the best emission intensity is obtained with 2% Cr3+ doping. Fig. 5(c) depicts the emission spectra of LCZO: 2% Cr3+, y Nd3+ (y = 0% ∼ 8%), the intensity of the emission spectrum reaches a maximum at 1% Nd3+ doping. In addition, the fluorescence intensity at 861 nm for LCZO: 2% Cr3+, y Nd3+ (y = 0 ∼ 8%) gradually decreases with increasing Nd3+ concentration as shown in Fig. S8, which implies that energy transfer between Cr3+ and Nd3+ occurs. The emission spectra of LCZO: 2% Cr3+, y Nd3+ (y = 0% ∼ 8%) at 471 nm excitation does not show an emission peak at 695 nm, which implies that the Cr3+ ions absorb the main excitation energy.

The central distance between ions will be less than the critical distance of ions at a high doping concentration of Nd3+ ions, and the critical distance of energy transfer can be calculated by the equation proposed by Blasse as follows [34]:

$${R_c}\; \approx 2{\left[ {\frac{{3V}}{{4\pi {x_c}N}}} \right]^{1/3}}$$
where V (V = 288.125 Å3) refers to the cell volume of the sample material, Rc represents the critical distance between Nd3+, xc(xc = 0.06 for LCZO: Nd3+, xc = 0.01 for LCZO: 2%Cr3+, yNd3+) represents the critical concentration, and N is the cation number in the unit cell (N = 3). The critical distance of LCZO: 2%Cr3+, yNd3+ (Rc = 26.37 Å) is larger than that of LCZO: Nd3+ (Rc = 7.26 Å), which can be seen from the fact that the introduction of Cr3+ leads to smaller grain plane spacing and as well as looser atomic spacing.

The interaction between Cr3+ and Nd3+ can be explained by Reisfeld's approximation and Dexter's multipole interaction theory. The type of interaction between the donor and acceptor ions can be determined according to the following equation [35]:

$$\frac{{{I_{{s_0}}}}}{{{I_S}}}\; \alpha \; {C^{\frac{n}{3}}}$$
where Is0 and Is denote the integrated intensity of the emission spectra in the presence and absence of Nd3+, respectively; C is the critical doping concentration; and n = 6, 8, and 10 represent dipole-dipole (d-d), dipole-quadrupole (d-q), and quadrupole-quadrupole (q-q) interactions, respectively. As shown in Fig. 5(d), the best fit line (R2 = 98.7%) is obtained for n = 6, which indicates that the d-d interaction of electrons plays a dominant role between Cr3+ and Nd3+.

To further investigate the effect of Cr3+ introduction on LCZO: Nd3+ phosphors, LCZO: 4% Nd3+, LCZO: 4% Nd3+, 2% Cr3+ are prepared under the same conditions, and their DRS spectra are shown in Fig. 6(a). The introduction of 2% Cr3+ ions significantly enhance the absorption of the phosphor in the UV-vis-NIR band, and the characteristic absorption of Cr3+ and Nd3+ coexists. The difference in the radii of Cr3+ ions with Ca2+ ions and Zr4+ ions and the charge compensation mechanism can cause defects centers in LCZO crystals.

 figure: Fig. 6.

Fig. 6. (a) Diffuse reflectance spectra of LCZO: 4% Nd3+, LCZO: 4% Nd3+, 2% Cr3+. (b) Diagram of Urbach tail in the energy band. (c) K-M function plot of reflection spectra of LCZO: 4% Nd3+, LCZO: 4% Nd3+, 2% Cr3+. (d) Plot of Ln(F(R)) versus hν for the width of Urbach tail determined by the slope.

Download Full Size | PDF

As shown in Fig. 6(b), these defects generate strong potential gradients in the LCZO host, leading to quantum mechanical penetration of the localized states into the forbidden band gap [36]. The relationship between [F(R)hν]2 and energy can be obtained by substituting the data in Fig. 6(a) into Equation (3) as in Fig. 6(c). The doping of Cr3+ ions lead to a slight increase in the optical band gap of LCZO (3.56eV→3.58 eV), which is due to the Burstein-Moss effect [37].

The Urbach tail (Eu) width can be obtained from the absorption coefficient (α) versus the photon energy () as follows [38]:

$$\alpha = \; {\alpha _0}\; \textrm{exp}\left( {\frac{{h\nu }}{{{E_u}}}} \right)$$
$$\ln (\alpha )= \ln ({{\alpha_0}} )+ \frac{{h\nu }}{{{E_u}}}\; $$
$$\ln ({F(R )} )= \ln ({{R_0}} )+ \; \frac{{h\nu }}{{{E_u}}}$$
where α0 is a constant, F(R) is proportional to α, and Equations (8) and (9) can be derived by taking the logarithm of equation (7). Fig. 6(d) depicts the plot between ln(F(R)) and energy for LCZO: 4% Nd3+, LCZO: 4% Nd3+, 2% Cr3+, where the value of Urbach tail (Eu) can be obtained from the inverse of the slope of the straight line after linear fitting of the data points. The results show that the Urbach energy value of LCZO: 4% Nd3+, 2% Cr3+ (Eu = 308 meV) is higher than that of LCZO: 4% Nd3+ (Eu = 302 meV). The introduction of Cr3+ ions lead to an increase in disordered atoms and an increase in structural bonding defects, which can produce localized states at or near the conduction band level, leading to an increase in Urbach tail (Eu) [39,40].

Fig. S9(a), (b) show the fluorescence lifetimes monitored at 861 nm for Cr3+ luminescence-dominated and 1084 nm for Nd3+ luminescence-dominated, respectively. With the increase of Nd3+ concentration, the fluorescence lifetime at 861 nm becomes smaller, while the fluorescence lifetime at 1084 nm shows a trend of first increasing and then decreasing, which implies the process of energy transfer from Cr3+ to Nd3+.

The energy transfer efficiency can be calculated by the reduction of the emission intensity of Cr3+ or the shortening of its lifetime as in the following equation [41,42]:

$${\eta _{ET}} = 1 - \; \frac{{{I_{Cr}}}}{{{I_{C{r_0}}}}}$$
$${\eta _{ET}}^{\prime} = 1 - \; \frac{{{\tau _{Cr}}}}{{{\tau _{C{r_0}}}}}$$
where ICr and ICr0 are the integrated intensities of the emission spectra of Cr3+ in the presence and absence of Nd3+, respectively, and τCr and τCr0 are the corresponding fluorescence lifetimes.

Fig. 7(a) shows that the energy transfer efficiencies calculated by the two methods are similar, while it can be seen that the energy transfer efficiency increases with increasing Nd3+ concentration, reaching 88.4% at 6% Nd3+ concentration according to the variation of Cr3+ emission spectral intensity. Fig. 7(b) exhibits the emission spectra of LCZO:1% Nd3+, LCZO: 6% Nd3+, LCZO: 2% Cr3+, LCZO: 2% Cr3+, 1%Nd3+ and their integrated intensity comparison at 471 nm excitation. The emission spectra of LCZO: 2% Cr3+, 1% Nd3+ have an integrated intensity of 847% relative to LCZO: 1% Nd3+, 457% relative to LCZO: 6% Nd3+, and 204% relative to LCZO: 2% Cr3+, and have a longer wavelength dual-peak emission. Moreover, Fig. S10 shows the external quantum yield of 27.8% for LCZO: %Cr3+,1%Nd3+.

 figure: Fig. 7.

Fig. 7. (a) Variation of emission spectral intensity with Nd3+ concentration, energy transfer efficiency calculated based on Cr3+ emission intensity, lifetime. (b) The emission spectra of LCZO:1% Nd3+, LCZO: 6% Nd3+, LCZO: 2% Cr3+, LCZO: 2% Cr3+, 1%Nd3+ at 471 nm excitation.

Download Full Size | PDF

Fig. 8 depicts the schematic diagram of the energy levels in LCZO: Cr3+, Nd3+ and the corresponding energy transfer processes. The phosphor LCZO: Cr3+, Nd3+ is excited at 471 nm (21231 cm-1), and the Cr3+ ions absorb the main excitation energy and radiate a broadband emission spectrum of 700 (14285 cm-1)-1400 nm (7142 cm-1) as shown in process (1). The process (2) shows that electrons from the 4T1 and 4T2 excited states relaxing to the 2E metastable state can be transferred to the 4F7/2 or 4S3/2, 4F5/2 or 2H9/2, 4F3/2 of Nd3+ and subsequently relax to the lower 4F3/2 excited state of Nd3+ by a resonant energy transfer process. At 471 nm excitation, a small fraction of the ground state 4I9/2 electrons of the Nd3+ ions are pumped to the 4G9/2 excited state, and then they rapidly nonradiative relax to the 4F3/2 energy level of Nd3+ as depicted by the red dashed arrow in process (3). As discussed earlier, there is rapid diffusion and reabsorption between the 4F3/24I9/2 energy levels of the Nd3+ ions (process (4) in the Fig. 8). The electrons on the 4F3/2 excited state of the Nd3+ ions eventually return to the 4I9/2, 4I11/2, 4I13/2 ground state, producing enhanced near-infrared emission.

 figure: Fig. 8.

Fig. 8. Schematic energy level diagrams of Cr3+, Nd3+ in LCZO, illustrating the involved energy transfer process.

Download Full Size | PDF

In addition, the temperature-dependent PL spectra of LCZO: 2% Cr3+, 1% Nd3+ are shown in Fig. S11(a). The intensity of the emission spectra decreases with increasing temperature, and the intensity of the emission spectrum at 373 K is 67% of that at room temperature. As the temperature increases, the emission peak at 822 nm is gradually shown, which arises from the 4F5/24I9/2 electron transition of Nd3 +  [43]. Fig. S11(b) depicts the emission peak intensity ratios at 907 nm and 890 nm for LCZO: 2% Cr3+, 1% Nd3+, where the value of I907/I890 decreases with increasing temperature. Fig. S11(c) shows β5/2, β9/2, β11/2, and β13/2 versus temperature for LCZO: 2% Cr3+, 1% Nd3+. The value of β5/2 gradually rises with increasing temperature, while the values of β9/2, β11/2, and β13/2 show a decreasing trend with increasing temperature. According to Boltzmann's distribution law, the population of (4F5/2) states increase with temperature compared to the 4F3/2 level [44]. The activation energy ΔE of the non-radiative relaxation of the phosphor can be obtained from the following equation [45]:

$${I_T} = \; \frac{{{I_0}}}{{1 + c\; \textrm{exp}\left( {\frac{{\Delta E}}{{kT}}} \right)}}$$
where IT is the intensity of the emission spectrum at a certain temperature, I0 is the intensity of the emission spectrum at room temperature, c is a constant, and ΔE and k are the activation energy and Boltzmann constant, respectively. According to the plot of Ln(I0/IT-1) versus 1/kT in Fig. S11(d), the activation energy of LCZO: 2% Cr3+, 1%Nd3+ is 0.290 eV.

3.3. Preparation of NIR pc-LED and its multifunctional application

Face recognition is a common mode in biometrics and is used in public safety. In the field of face recognition, NIR light is more suitable for face recognition than UV and visible light due to its non-destructive characteristics to human body. Fig. S12 shows that the sample LCZO: 2% Cr3+, 1% Nd3+ can be maintained at 373 K at 67% of room temperature, which is higher than 64% of LCZO: 2% Cr3+, 2% Nd3+ and 46% of LCZO: 2% Cr3+, 4% Nd3+, so LCZO: 2% Cr3+, 1% Nd3+ is chosen as the NIR light source. Fig. S13 shows the thermograms of pc-LEDs prepared using LCZO:2%Cr3+,1%Nd3+ at currents ranging from 0 ∼ 300 mA, where the temperature of pc-LEDs rises with increasing current and reaches 99.3 °C at 300 mA. Fig. 9(a), (b) shows the photos of the NIR camera under NIR light and blue light irradiation, where the faces under NIR light can be clearly recognized, while the photos under blue light are very blurred. Fig. 9(c) shows that the visible light camera does not take good pictures in backlight conditions. Fig. 9(d), (e) shows the photographs of orange, pear, grape and dragon fruit taken by the visible camera under white fluorescent light irradiation and those taken by the NIR camera under NIR pc-LED irradiation, respectively. Fig. 9(f) exhibits a photograph of a NIR pc-LED applied to finger penetration, where blood vessels can be clearly identified, indicating its potential application in the field of biological tissue penetration imaging as well. The above applications show that NIR light sources with dual emission peaks have great potential for applications in several fields.

 figure: Fig. 9.

Fig. 9. (a), (b) NIR pc-LED for face recognition. (c) Photo taken by visible light camera in backlight environment. (d), (e)NIR pc-LED for night lighting of oranges, pears, grapes and dragon fruit. (f) NIR light source for angiography.

Download Full Size | PDF

4. Conclusion

In summary, we explore LCZO: Cr3+, Nd3+ phosphors with emission spectra covering the range of 700-1400 nm under 471 nm blue light excitation. It is important to note that the energy transfer efficiency of Cr3+ for Nd3+ is as high as 88.4%. The spectral integrated intensity of LCZO: Cr3+, Nd3+ at 471 nm excitation is 847% of LCZO: Nd3+ and 204% of LCZO: Cr3+. This work provides a new vision for the development of novel NIR phosphors with Nd3+ as the dominant activator. Finally, LCZO: Cr3+, Nd3+ phosphor and 470 nm InGaN chips were combined to make NIR pc-LED. The NIR light sources are used in face recognition, night illumination and angiography, showing great potential for applications.

Funding

National Natural Science Foundation of China (62175075, 62075055); Natural Science Foundation of Hebei Province (2020201030); Central Project Guide local science and technology for development of Hebei Province (216Z1104G, 226Z1103G).

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. M.-H. Fang, Z. Bao, W.-T. Huang, and R.-S. Liu, “Evolutionary Generation of Phosphor Materials and Their Progress in Future Applications for Light-Emitting Diodes,” Chem. Rev. 122(13), 11474–11513 (2022). [CrossRef]  

2. A. Guelpa, F. Marini, A. du Plessis, R. Slabbert, and M. Manley, “Verification of authenticity and fraud detection in South African honey using NIR spectroscopy,” Food Control 73, 1388–1396 (2017). [CrossRef]  

3. S. Pradhan, M. Dalmases, and G. Konstantatos, “Solid-State Thin-Film Broadband Short-Wave Infrared Light Emitters,” Adv. Mater. 32(45), 2003830 (2020). [CrossRef]  

4. E. F. Schubert and J. K. Kim, “Solid-State Light Sources Getting Smart,” Science 308(5726), 1274–1278 (2005). [CrossRef]  

5. N. Tessler, V. Medvedev, M. Kazes, S. Kan, and U. Banin, “Efficient Near-Infrared Polymer Nanocrystal Light-Emitting Diodes,” Science 295(5559), 1506–1508 (2002). [CrossRef]  

6. Y. Zhuo and J. Brgoch, “Opportunities for Next-Generation Luminescent Materials through Artificial Intelligence,” J. Phys. Chem. Lett. 12(2), 764–772 (2021). [CrossRef]  

7. 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, 131332 (2021). [CrossRef]  

8. D. C. Huang, X. G. He, J. R. Zhang, J. Hu, S. S. Liang, D. J. Chen, K. Y. Xu, and H. M. 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]  

9. C. Yuan, R. Li, Y. Liu, L. Zhang, J. Zhang, G. Leniec, P. Sun, Z. Liu, Z. Luo, R. Dong, and J. Jiang, “Efficient and Broadband LiGaP2O7:Cr3+ Phosphors for Smart Near-Infrared Light-Emitting Diodes,” Laser Photonics Rev. 15(11), 2100227 (2021). [CrossRef]  

10. F. He, E. Song, Y. Zhou, H. Ming, Z. Chen, J. Wu, P. Shao, X. Yang, Z. Xia, and Q. Zhang, “A General Ammonium Salt Assisted Synthesis Strategy for Cr3+-Doped Hexafluorides with Highly Efficient Near Infrared Emissions,” Adv. Funct. Mater. 31(36), 2103743 (2021). [CrossRef]  

11. 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]  

12. D. W. Wen, H. M. Liu, Y. Guo, Q. G. Zeng, M. M. Wu, and R. S. Liu, “Disorder-Order Conversion-Induced Enhancement of Thermal Stability of Pyroxene Near-Infrared Phosphors for Light-Emitting Diodes,” Angew. Chem., Int. Ed. 61(28), 8 (2022). [CrossRef]  

13. G. S. Maciel and N. Rakov, “Thermometric analysis of the near-infrared emission of Nd3+ in Y2SiO5 ceramic powder prepared by combustion synthesis,” Ceram. Int. 46(8), 12165–12171 (2020). [CrossRef]  

14. Y. V. Baklanova, A. N. Enyashin, L. G. Maksimova, A. P. Tyutyunnik, A. Y. Chufarov, E. V. Gorbatov, I. V. Baklanova, and V. G. Zubkov, “Sensitized IR luminescence in Ca3Y2Ge3O12: Nd3+, Ho3+ under 808 nm laser excitation,” Ceram. Int. 44(6), 6959–6967 (2018). [CrossRef]  

15. J. De Anda, E. F. Huerta, J. U. Balderas, G. C. Righini, and C. Falcony, “The effect of Li+ incorporation in Yb3+-Nd3+ co-doped CaF2 phosphors over the NIR photoluminescence emission excited under visible light,” Ceram. Int. 47(4), 4694–4701 (2021). [CrossRef]  

16. S. X. Guo, S. A. Zhang, Z. F. Mu, F. G. Wu, X. Feng, Q. T. Zhang, J. Q. Feng, D. Y. Zhu, and Q. P. Du, “Enhanced near infrared luminescence of Lu2GeO5: Nd3+ by the co-doping of Bi3+,” J. Lumin. 206, 278–283 (2019). [CrossRef]  

17. K. Lio and R. Van Deun, “Novel Intense Emission-Tunable Li1.5La1.5WO6:Mn4+,Nd3+,Yb3+ Material with Good Luminescence Thermal Stability for Potential Applications in c-Si Solar Cells and Plant-Cultivation Far-Red-NIR LEDs,” ACS Sustainable Chem. Eng. 7(19), 16284–16294 (2019). [CrossRef]  

18. J. Liu, G. Li, H. J. Guo, D. W. Liu, P. Feng, and Y. H. Wang, “Design, synthesis and characterization of a novel bluish-green long-lasting phosphorescence phosphor BaLu2Si3O10:Eu2+, Nd3+,” RSC Adv. 8(19), 10246–10254 (2018). [CrossRef]  

19. A. A. Pathak, R. A. Talewar, C. P. Joshi, and S. V. Moharil, “NIR emission and Ce3+-Nd3+ energy transfer in LaCaAl3O7 phosphor prepared by combustion synthesis,” J. Lumin. 179, 350–354 (2016). [CrossRef]  

20. V. Anselm and T. Justel, “On the photoluminescence and energy transfer of SrGa12O19:Cr3+,Nd3+ microscale NIR phosphors,” J. Mater. Res. Technol. 11, 785–791 (2021). [CrossRef]  

21. Y. L. Wu, Y. Li, X. X. Qin, R. C. Chen, D. K. Wu, S. J. Liu, and J. R. Qiu, “Dual mode NIR long persistent phosphorescence and NIR-to-NIR Stokes luminescence in La3Ga5GeO14: Cr3+, Nd3+ phosphor,” J. Alloys Compd. 649, 62–66 (2015). [CrossRef]  

22. A. L. Mullins, A. Ciric, I. Zekovic, J. A. G. Williams, M. D. Dramicanin, and I. R. Evans, “Dual-emission luminescence thermometry using LaGaO3:Cr3+, Nd3+ phosphors,” J. Mater. Chem. C 10(28), 10396–10403 (2022). [CrossRef]  

23. M. C. Payne, M. P. Teter, D. C. Allan, T. A. Arias, and J. D. Joannopoulos, “Iterative minimization techniques for ab initio total-energy calculations: molecular dynamics and conjugate gradients,” Rev. Mod. Phys. 64(4), 1045–1097 (1992). [CrossRef]  

24. M. D. Segall, J. D. L. Philip, M. J. Probert, C. J. Pickard, P. J. Hasnip, S. J. Clark, and M. C. Payne, “First-principles simulation: ideas, illustrations and the CASTEP code,” J. Phys.: Condens. Matter 14(11), 2717–2744 (2002). [CrossRef]  

25. J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized Gradient Approximation Made Simple,” Phys. Rev. Lett. 77(18), 3865–3868 (1996). [CrossRef]  

26. S. Guo, Y. Wang, C. Wang, Z. Tang, and J. Zhang, “Large spin-orbit splitting in the conduction band of halogen (F, Cl, Br, and I) doped monolayer WS2 with spin-orbit coupling,” Phys. Rev. B 96(24), 245305 (2017). [CrossRef]  

27. Y. Li, C. Zhang, X. Zhang, D. Huang, Q. Shen, Y. Cheng, and W. Huang, “Intrinsic point defects in inorganic perovskite CsPbI3 from first-principles prediction,” Appl. Phys. Lett. 111(16), 162106 (2017). [CrossRef]  

28. 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, 129224 (2021). [CrossRef]  

29. J. Wang, L. Liang, L. Zhang, L. Sun, and S. Hirano, “Deduction of the chemical state and the electronic structure of Nd2Fe14B compound from X-ray photoelectron spectroscopy core-level and valence-band spectra,” J. Appl. Phys. 116(16), 163917 (2014). [CrossRef]  

30. T. Lang, M. Cai, S. Fang, T. Han, S. He, Q. Wang, G. Ge, J. Wang, C. Guo, L. Peng, S. Cao, B. Liu, V. I. Korepanov, A. N. Yakovlev, and J. Qiu, “Trade-off Lattice Site Occupancy Engineering Strategy for Near-Infrared Phosphors with Ultrabroad and Tunable Emission,” Adv. Opt. Mater. 10(2), 2101633 (2022). [CrossRef]  

31. H. S. Lokesha, K. R. Nagabhushana, F. Singh, S. H. Tatumi, A. R. E. Prinsloo, and C. J. Sheppard, “Unraveling the Charge State of Oxygen Vacancies in Monoclinic ZrO2 and Spectroscopic Properties of ZrO2:Sm3+ Phosphor,” J. Phys. Chem. C 125(49), 27106–27117 (2021). [CrossRef]  

32. R. S. Monika, A. Yadav, S. B. Bahadur, and Rai, “Near-Infrared Light Excited Highly Pure Green Upconversion Photoluminescence and Intrinsic Optical Bistability Sensing in a Ho3+/Yb3+ Co-Doped ZnGa2O4 Phosphor through Li+ Doping,” J. Phys. Chem. C 124(18), 10117–10128 (2020). [CrossRef]  

33. L. Marciniak, W. Strek, Y. Guyot, and D. Hreniak, “Synthesis and luminescent properties of La1-xNdxP5O14 nanocrystals,” Phys. Chem. Chem. Phys. 16(33), 18004–18009 (2014). [CrossRef]  

34. A. Huang, Z. Yang, C. Yu, Z. Chai, J. Qiu, and Z. Song, “Tunable and White Light Emission of a Single-Phased Ba2Y(BO3)2Cl:Bi3+,Eu3+ Phosphor by Energy Transfer for Ultraviolet Converted White LEDs,” J. Phys. Chem. C 121(9), 5267–5276 (2017). [CrossRef]  

35. G. Xing, Y. Feng, M. Pan, Y. Wei, G. Li, P. Dang, S. Liang, M. S. Molokeev, Z. Cheng, and J. Lin, “Photoluminescence tuning in a novel Bi3+/Mn4+ co-doped La2ATiO6:(A = Mg, Zn) double perovskite structure: phase transition and energy transfer,” J. Mater. Chem. C 6(48), 13136–13147 (2018). [CrossRef]  

36. D. Redfield, “Effect of Defect Fields on the Optical Absorption Edge,” Phys. Rev. 130(3), 916–918 (1963). [CrossRef]  

37. A. Mondal, S. Das, and J. Manam, “Hydrothermal synthesis, structural and luminescent properties of a Cr3+ doped MgGa2O4 near-infrared long lasting nanophospor,” RSC Adv. 6(86), 82484–82495 (2016). [CrossRef]  

38. A. K. Srivastava and J. Kumar, “Effect of precursor solvent on the opto-electrical properties of spin coated transparent conducting ZnO: Ga thin films,” Mater. Chem. Phys. 162, 436–441 (2015). [CrossRef]  

39. S. Jana, A. Mondal, J. Manam, and S. Das, “Pr3+ doped BaNb2O6 reddish orange emitting phosphor for solid state lighting and optical thermometry applications,” J. Alloys Compd. 821, 153342 (2020). [CrossRef]  

40. A. S. Hassanien and A. A. Akl, “Influence of composition on optical and dispersion parameters of thermally evaporated non-crystalline Cd50S50−xSex thin films,” J. Alloys Compd. 648, 280–290 (2015). [CrossRef]  

41. M. Inokuti and F. Hirayama, “Influence of Energy Transfer by the Exchange Mechanism on Donor Luminescence,” J. Chem. Phys. 43(6), 1978–1989 (1965). [CrossRef]  

42. P. I. Paulose, G. Jose, V. Thomas, N. V. Unnikrishnan, and M. K. R. Warrier, “Sensitized fluorescence of Ce3+/Mn2+ system in phosphate glass,” J. Phys. Chem. Solids 64(5), 841–846 (2003). [CrossRef]  

43. M. Suta, Ž Antić, V. Ðorđević, S. Kuzman, M. D. Dramićanin, and A. Meijerink, “Making Nd3+ a Sensitive Luminescent Thermometer for Physiological Temperatures—An Account of Pitfalls in Boltzmann Thermometry,” Nanomaterials 10(3), 543 (2020). [CrossRef]  

44. G. Jiang, X. Wei, S. Zhou, Y. Chen, C. Duan, and M. Yin, “Neodymium doped lanthanum oxysulfide as optical temperature sensors,” J. Lumin. 152, 156–159 (2014). [CrossRef]  

45. Z. Wu, X. Han, Y. Zhou, K. Xing, S. Cao, L. Chen, R. Zeng, J. Zhao, and B. Zou, “Efficient broadband near-infrared luminescence of Cr3+ doped fluoride K2NaInF6 and its NIR-LED application toward veins imaging,” Chem. Eng. J. 427, 131740 (2022). [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 (9)

Fig. 1.
Fig. 1. (a) Crystal structure of LCZO. (b) Cr3+, Nd3+ substitution schematic. (c) Rietveld refinement of XRD of LCZO: 2% Cr3+, 4% Nd3+. (d), (e) XRD patterns of LCZO: x Nd3+ (x = 0.1%∼8%). (f), (g) XRD patterns of LCZO: 2% Cr3+, y Nd3+ (y = 0%∼8%).
Fig. 2.
Fig. 2. (a) XPS spectra of LCZO: 2% Cr3+, 4%Nd3+. (b), (c) The HRTEM image of phosphor LCZO:2% Cr3+, 4% Nd3+ particles and elemental distribution maps.
Fig. 3.
Fig. 3. (a) Diffuse reflectance spectra of LCZO: x Nd3+ (x = 0 ∼ 8%). (b) Diffuse reflectance spectra of LCZO: 2%Cr3+, y Nd3+ (y = 0% ∼ 8%). (c) PL and PLE spectra of LCZO: Cr3+, LCZO: Nd3+. (d) PL and PLE spectra of LCZO: 2%Cr3+, 0.5%Nd3+.
Fig. 4.
Fig. 4. Schematic diagram of the energy levels of Nd3+ in LCZO.
Fig. 5.
Fig. 5. (a) Emission spectra of LCZO: x Nd3+ (x = 0.1 ∼ 8%). (b) Branching ratio βJ/J’ as a function of concentration of Nd3+. (c) Emission spectra of LCZO: 2%Cr3+, y Nd3+ (y = 0%∼8%). (d) Fitted curves of Is0/Is versus Cn/3(n = 6, 8, 10) associated with LCZO: 2% Cr3+, y Nd3+ (y = 0% ∼ 8%).
Fig. 6.
Fig. 6. (a) Diffuse reflectance spectra of LCZO: 4% Nd3+, LCZO: 4% Nd3+, 2% Cr3+. (b) Diagram of Urbach tail in the energy band. (c) K-M function plot of reflection spectra of LCZO: 4% Nd3+, LCZO: 4% Nd3+, 2% Cr3+. (d) Plot of Ln(F(R)) versus hν for the width of Urbach tail determined by the slope.
Fig. 7.
Fig. 7. (a) Variation of emission spectral intensity with Nd3+ concentration, energy transfer efficiency calculated based on Cr3+ emission intensity, lifetime. (b) The emission spectra of LCZO:1% Nd3+, LCZO: 6% Nd3+, LCZO: 2% Cr3+, LCZO: 2% Cr3+, 1%Nd3+ at 471 nm excitation.
Fig. 8.
Fig. 8. Schematic energy level diagrams of Cr3+, Nd3+ in LCZO, illustrating the involved energy transfer process.
Fig. 9.
Fig. 9. (a), (b) NIR pc-LED for face recognition. (c) Photo taken by visible light camera in backlight environment. (d), (e)NIR pc-LED for night lighting of oranges, pears, grapes and dragon fruit. (f) NIR light source for angiography.

Tables (2)

Tables Icon

Table 1. The calculated formation energies (Ef) of Nd3+ occupying different sites in LCZO

Tables Icon

Table 2. Atomic occupancy information of LCZO: 2% Cr3+, 4% Nd3+

Equations (12)

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

E f = E L C Z O : N d E L C Z O μ N d + μ s
F ( R ) = ( 1 R ) 2 2 R
( F ( R ) h ν ) n = A ( h ν E g )
β J / J = A J J A J J
R c 2 [ 3 V 4 π x c N ] 1 / 3
I s 0 I S α C n 3
α = α 0 exp ( h ν E u )
ln ( α ) = ln ( α 0 ) + h ν E u
ln ( F ( R ) ) = ln ( R 0 ) + h ν E u
η E T = 1 I C r I C r 0
η E T = 1 τ C r τ C r 0
I T = I 0 1 + c exp ( Δ E 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.