Introduction

Near-infrared (NIR) luminescent materials have attracted wide attention in emerging technology fields such as NIR spectroscopic analysis, bioimaging, and night vision1,2,3,4,5,6. To meet the current demands of portable NIR light sources, NIR-emitting phosphor-converted light-emitting diodes (pc-LEDs) have been widely developed due to their compactness and tunable broadband emission7,8,9,10. The key is to exploit blue-light excitable high-performance NIR-emitting phosphor materials. Numerous NIR-emitting phosphors have been discovered by doping rare-earth and transition metal ions, among which Cr3+-activated phosphors stand out given that they can easily generate broadband NIR emission in the range from 700 to 1100 nm11,12,13,14,15,16. This enables their applications in smart devices that are integrated with Si-based detectors17. Recently, progress has been made in realizing tunable broadband emission of Cr3+-activated phosphors18,19,20. However, the major issues still remain their unsatisfactory photoluminescence quantum efficiency and poor thermal quenching resistance, restricting their practical applications in pc-LEDs.

To address these issues, researchers have focused on the Cr3+-activated garnet-structured NIR-emitting phosphors21,22,23,24,25. The ideal garnets are assigned to the Ia-3d space group of cubic crystal systems with a chemical formula of A3B2C3O12. There are dodecahedra (A), octahedra (B), and tetrahedra (C) in garnet structures. The dodecahedra and octahedra are connected through edge-sharing, while the octahedra and tetrahedra are connected by vertex-sharing. The diversity of cation selection in garnets results in a great number of chemical variants. A3B2C3O12-typed garnets are considered promising host materials that can provide Cr3+ ions with octahedral B sites, compact coordinated environment, and tunable structures, and thereby diverse luminescence properties including the required ones can be expected26,27,28,29. For example, Gd3ZnxGa5-2xGexO12:Cr3+ show tunable broadband NIR emission via crystal field engineering; while Y3(Al,Mg)2(Al,Si)3O12:Cr3+ exhibit efficient and thermal quenching resistant NIR emission30,31. Unfortunately, there is a trade-off between the emission wavelength and quantum efficiency as well as thermal stability. That is, the highly efficient and thermally stable NIR luminescence is generally accompanied by short emission wavelength (<750 nm), which limits the applications of these garnet phosphors in some scenarios like food analysis32,33. Therefore, how to improve luminescence efficiency and thermal stability while maintaining wavelength within specific range is still a challenge for Cr3+-doped NIR-emitting garnet phosphors.

In this regard, two dominant factors affecting the luminescence of Cr3+-doped garnet phosphors should be mentioned. One is the luminescence “killer” Cr4+ that shows intensive absorption in NIR region34. The common high-temperature synthesis of the phosphors in air and the existed tetrahedral C sites in A3B2C3O12 garnets synergistically favor the formation of Cr4+ ions35,36. Although it can be partly prevented by reducing atmosphere, this way is not suitable for all host materials25. But from another perspective, this problem can be solved by

constructing more appropriate crystallographic B sites for accommodating Cr3+ ions to increase their formation competitiveness in phosphors. Another crucial factor is structural rigidity. A rigid host structure can weaken the electron-phonon coupling effect and suppress the nonradiative transition20,22.

In this work, the underlying luminescence optimization mechanism of Ca3Y2-2x(ZnZr)xGe3O12:Cr garnet system was investigated based on structural analysis and density functional theory (DFT) calculations. The coexistence of Cr3+ and Cr4+ in Ca3Y2Ge3O12:Cr was demonstrated to be responsible for its low quantum efficiency. With reconstruction of the crystallographic B sites for Cr3+ ions in Ca3Y2Ge3O12 by the designed cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+], Cr4+ luminescence killers were transformed to beneficial Cr3+ emission centers. The strategy of chemical unit cosubstitution killed two birds with one stone, which also successfully built a rigid crystal structure. These changes led to the significant improvement of photoluminescence quantum efficiency and thermal stability with only slight emission shift. Finally, information encryption inspired by chemical resistance difference was designed to achieve “burning after reading”. The great potential of the developed portable pc-LED in bio-tissue imaging and night-vision applications was also demonstrated.

Results

Crystal structure and phase identification

Figure 1a shows the crystal structure of Ca3Y2Ge3O12, which adopts a typical garnet structure with formula A3B2C3O12. Ca2+, Y3+, and Ge4+ ions are distributed at the dodecahedral A sites, octahedral B sites, and tetrahedral C sites, respectively. Considering that the crystal field stabilization energy of Cr3+ in octahedral site is eight times higher than that in tetrahedral site, Cr3+ ions preferentially occupy the octahedral Y3+ sites in Ca3Y2Ge3O12 host, while maintaining charge balance37,38. However, such a substitution can cause large lattice strain due to the obvious size-mismatch between Cr3+ and Y3+ ions (Fig. 1b and Table S1). This increases the substitution difficulty, which can be evaluated by the parameter Dr obtained from the empirical formula39:

$${D}_{r}=\frac{\left|{R}_{r}-{R}_{d}\right|}{{R}_{r}}\times 100 \%$$
(1)

where Dr represents the difference of ion radii between the doped ions and host cations, Rr and Rd are the radii of the host cations and the doped ions, respectively. The substitution process is believed to be possible on condition that the Dr is less than 30%. Dr for Cr3+ substituting Y3+ is 31.7%, which means the substitution is difficult. That is, it is unfavorable for Cr3+ ions to replace Y3+ ions. However, Dr for Cr4+ substituting Ge4+ is only 5.1%, indicating that chromium ions are very likely to enter the Ge4+ sites in the form of tetravalent Cr4+ ions. To maintain trivalent state of chromium ions in the host, octahedral sites of appropriate size should be constructed to accommodate Cr3+. Smaller Zn2+ and Zr4+ ions were introduced into octahedral Y sites by the designed chemical unit cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+], which was expected to reduce the size-mismatch between Cr3+ ions and Y3+ sites so as to promote the formation of Cr3+ instead of Cr4+. The valence preferences of chromium ions in Ca3Y2Ge3O12 and Ca3ZnZrGe3O12 were simulated by DFT calculations. According to the above analysis, two substitution models for the two hosts were considered, respectively. That is, one Cr3+ ion occupied one Y3+ site (Model 1) and one Cr4+ ions occupied one Ge4+ site (Model 2) in Ca3Y2Ge3O12; two Cr3+ ions occupied one Zn2+ site and one Zr4+ site (Model 3), and one Cr4+ ions occupied one Ge4+ site (Model 4) in Ca3ZnZrGe3O12. As show in Fig. 1c, the formation energy values of Model 2 and Model 3 are significantly smaller than those of Model 1 and Model 4, respectively, which indicates that chromium ions tend to exist in the form of tetravalent Cr4+ ions in Ca3Y2Ge3O12 and trivalent Cr3+ ions in Ca3ZnZrGe3O12.

Fig. 1: Structure and morphology characterizations of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1).
figure 1

a Crystal structure of Ca3Y2Ge3O12. b Ionic radii for cation ions under corresponding coordination numbers. c Formation energy (Ef) for different substitution models of Cr in Ca3Y2Ge3O12 and Ca3ZnZrGe3O12. d XRD patterns of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1). e XRD Rietveld refinement of x = 1. The inset shows the relative changes in the cell parameters (a, b, and c), volumes (V), and bond lengths (Ca–O, Y/Zn/Zr–O, and Ge–O) with x. f SEM images of x = 0 and x = 1. g Element mapping images of Ca, Zn, Zr, Ge, O, and Cr elements in x = 1

Figure 1d shows that the X-ray diffraction (XRD) patterns of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1) are consistent with the calculated pattern of Ca3Y2Ge3O12, demonstrating the successful realization of cosubstitution. In addition, the XRD peak shifts toward higher angles, which is in accordance with the fact that Zn2+ and Zr4+ ions are smaller than Y3+ ions. Rietveld refinements (Fig. 1e and Fig. S1) were conducted to identify the pure phase and reveal the crystal structure changes in depth. The standard pattern of Ca3Y2Ge3O12 used for refinements in this work is Crystallography Open Database (COD) card no. 2009001. The refined results are listed in Tables S2S3. As shown in the inset of Fig. 1e, linear changes of lattice parameters (a, b, and c) and cell volume (V) with x demonstrate that Ca3Y2-2x(ZnZr)xGe3O12:Cr solid solutions are successfully synthesized. The lattice shows shrinkage with increasing x, which corresponds to the shift of XRD peak towards higher angles. In addition, it can be seen that the Y/Zn/Zr–O bonds shorten with a more obvious change rate as x increases from 0 to 1. This is ascribed to the cosubstitution of smaller [Zn2+–Zr4+] for [Y3+–Y3+] in conjunction with the transformation from Cr4+ to Cr3+. More smaller Cr3+ ions enter in larger Y/Zn/Zr sites can further decrease the Y/Zn/Zr–O bond lengths. As for the Ca–O bonds, they show less significant shortening. This is ascribed to the large volume of [CaO8] dodecahedra, which is less affected by the cosubstitution process. While Ge–O bonds can be easily affected by Y/Zn/Zr–O bonds due to the much smaller size of [GeO4] tetrahedra. The Ge–O bonds also exhibit less shortening to reduce the lattice strain caused by the shrinkage of [CaO8] dodecahedra and [Y/Zn/ZrO6] octahedra. Figure 1f and Fig. S2 show the irregular morphologies of x = 0 and x = 1 phosphors. It is noted that x = 1 shows bulk crystals much larger than x = 0, which is conducive to luminescence. The elemental mapping images of x = 1 are given in Fig. 1g, confirming the homogeneous distribution of Ca, Zn, Zr, Ge, Cr, and O elements.

Photoluminescence properties

The optimal doping concentration of Cr in Ca3Y2Ge3O12 host was determined to be 0.01 (Fig. S3). Thereafter, the luminescence properties of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1) were regulated by the designed cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+]. As shown by the diffuse reflectance (DR) spectra in Fig. 2a, both x = 0 and x = 1 exhibit typical absorption peaks of Cr3+ at about 470 and 660 nm. However, Cr4+ ions are also observed in x = 0 noting an additional absorption signal at 1100 nm, which nearly disappears in x = 1. Moreover, the absorption peak of Cr3+ at 470 nm for x = 1 increases compared with x = 0, indicating the increased Cr3+ concentration in x = 1. The Cr K-edge X-ray absorption near-edge structure (XANES) spectra were measured to further determine the valence state of chromium. As shown in Fig. 2b, the main edge peaks of x = 0 (6007.4 eV) and x = 1 (6008.5 eV) are close to that of the standard Cr2O3, which suggests that chromium elements predominantly present the +3 oxidation state in the phosphors. However, it can be observed that the absorption edges of x = 0 and x = 1 are located between the absorption step of Cr2O3 and CrO2, which suggests the presence of both trivalent Cr3+ and tetravalent Cr4+ in the phosphors. Moreover, compared to x = 1, the absorption edge of x = 0 is closer to the absorption edge of CrO2, which means that there are more tetravalent Cr4+ ions in x = 0 than in x = 1. This result is consistent with the result shown by DRS. These results suggest the reduction of Cr4+ to Cr3+ with x increasing from 0 to 1, which supports the assumption that the cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+] contributes to the formation of trivalent Cr3+ ions. As the absorption spectrum of Cr4+ partially overlaps the emission spectrum of x = 0 (inset of Fig. 2a), the existence of Cr4+ can certainly reduce the photoluminescence quantum efficiency of Cr3+. The photoluminescence excitation (PLE) spectra of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1) phosphors in Fig. S4a correspond to those absorption peaks in DR spectra, which shows optimal absorption and excitation in the blue-light region, matching with commercial blue LED chips. The photoluminescence (PL) spectra of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1) phosphors excited under 470 nm are shown in Fig. 2c, exhibiting broadband emission originating from 4T2g4A2g transition in the range of 650–1100 nm. As x increases from 0 to 1, the PL intensity is increased by 2.7 times accompanied by a slight blue shift of the emission peak from 812 to 795 nm (Fig. S4b, c). The transformation from Cr4+ to Cr3+ certainly plays a positive role in the luminescence enhancement. The blue shift of d–d transition for Cr3+ was understood by the spectroscopic parameters Dq and B. Fig. S5 shows that Dq/B value increases with increasing x, which indicates the strengthened crystal field, as interpreted by Tanabe–Sugano diagram (details are provided in the Supporting Information, Fig. S6 and Table S4). Thus, the PL spectra show reasonable blue shift. Figure 2d shows the contour maps of the normalized PLE spectra and PL spectra of x = 0 measured under different conditions. No visible change in spectral profile and peak position in the PLE and PL spectra proves only one Cr3+ luminescent center in x = 0 phosphor. As shown in Fig. S7a, the PL spectrum of x = 0 phosphor measured under 7 K shows an almost symmetric emission peak at 805 nm without any new emission peaks appearing. Moreover, it can be seen that the two PLE spectra in Fig. S7b monitored at 750 and 880 nm completely overlap. These low-temperature results further demonstrate one luminescent center in x = 0, which is believed to be Cr3+ substituting Y3+ in the Ca3Y2Ge3O12 host.

Fig. 2: Photoluminescence properties of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1).
figure 2

a Diffuse reflectance spectra of x = 0 and x = 1. The inset shows the absorption spectrum of Cr4+ and emission spectrum of Cr3+ in x = 0. b Normalized Cr K-edge XANES spectra of Ca3Y2-2x(ZnZr)xGe3O12:0.2Cr (x = 0 and 1) and reference compounds Cr2O3 and CrO2. c PL spectra of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1). d Contour maps of normalized PLE spectra and PL spectra of x = 0 measured under different monitoring and excitation wavelengths, respectively. e Raman spectra of x = 0 and x = 1. f PL decay curves of x = 0 and x = 1 measured under 7 K, monitoring emission at 805 nm for x = 0 and 785 nm for x = 1. The excitation wavelength is 470 nm. g Lifetimes and fitted R2 values of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1). h Schematic ion distribution in x = 0 and x = 1. The black arrows represent energy transfer. i Local structure change in x = 0 and x = 1. The large and small purple arrows represent the large and small lattice strain caused by doping, respectively. The thick blue lines represents the rigid bond with weak vibration. The black arrows represent energy transfer

To further explore the influence of the designed cosubstitution of on the local structures and PL properties, Raman spectra of x = 0 and x = 1 are compared in Fig. 2e. These Raman peaks at 100−280 and 280−1000 cm−1 are assigned to the [CaO8] and [GeO4] sites, respectively (Table S5)40. Apparently, the intensity of these Raman signals becomes much lower after the cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+], which means weaker vibrations of Ca–O and Ge–O bonds in x = 141. In addition, the vibration peaks broaden greatly for x = 1, indicating changes in crystal environment. CaO8 and GeO4 are surrounded by YO6 in x = 0, while surrounded by ZnO6 and ZrO6 in x = 1. The more uneven environment surrounding CaO8 and GeO4 in x = 1 can lead to broader Raman peaks. It should be noted that Ca2+ and Ge4+ ions construct the second coordination shell of Cr3+ ions. As displayed in the inset of Fig. 2e, one [CrO6] octahedra is edge-shared with six [CaO8] dodecahedra and vertex-shared with six [GeO4] tetrahedra. On the one hand, the shorter Cr–O bonds caused by cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+] indicate more strengthening bond force between Cr3+ and O2- ions, enabling a more rigid [CrO6] octahedra. On the other hand, weaker vibrations of the second coordination shells means that they have less influence on the vibrations of [CrO6] octahedra and offer a more rigid secondary coordination environment for Cr3+. The Debye temperature (ΘD) is regarded as an indicator of structural rigidity, which was theoretically calculated to be 495 K for Ca3Y2Ge3O12 and 564 K for Ca3ZnZrGe3O1242,43. The higher ΘD of Ca3ZnZrGe3O12 demonstrates the fact that it provides a more rigid structure for Cr3+. This is conducive to its radiative recombination, consistent with the enhanced NIR emission.

The luminescence decay behaviors of x = 0 and x = 1 phosphors under 7 K were studied (Fig. 2f). The decay curves were measured under such low temperature, which minimize the influence of temperature on luminescence decay and can better reflect the intrinsic lifetime of the phosphors. The influence of lattice thermal vibration can be ignored. The mono-exponential (n = 1) and bi-exponential (n = 2) functions were used to fit their decay curves:2

$$I=\mathop{\sum }\limits_{i=1}^{n}{A}_{{\rm{i}}}\exp \left(-\frac{t}{{\tau }_{{\rm{i}}}}\right),\left(n=1,2,\cdots \right)$$
(2)

where I is the emission intensity at time t, Ai is fitting constants, and τi is the lifetime for different components. The decay curve of x = 1 is well fitted by both mono- and bi-exponential function, and the two fitting curves almost overlap. It is worth noting that only one type of Cr3+ luminescent center existed in x = 0, while its decay curve deviates from mono-exponential fitting and is well fitted by the bi-exponential function. This mismatch between the numbers of luminescent centers and fitting index is due to the additional nonradiative transitions2. One common explanation is energy migration between activator ions44, which is very likely to occur in this work. The results of DFT calculations, DR spectra, and XANES spectra demonstrate that Cr3+ and Cr4+ coexist in the synthesized phosphors, and the designed cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+] played a role as reductant to promote the valence reduction of Cr4+ to Cr3+ (from x = 0 to x = 1). The overlap between the PL spectrum of Cr3+ and the absorption spectrum of Cr4+ (inset of Fig. 2a) suggests the possibility of energy transfer in x = 0. According to the structural analysis, the YO6 octahedra are separated by GeO4 and CaO8 polyhedra. In x = 0, the distance between Y3+ and Ge4+ (3.58 Å) is much smaller than that between Y3+ and Y3+ (5.55 Å), which indicates that the distance between Cr3+ and Cr4+ is much smaller than that between Cr3+ and Cr3+. That is, the energy transfer can occur more easily between Cr3+ and Cr4+ rather than between Cr3+ and Cr3+. Thus, it is reasonable to ascribe the bi-exponential fitting for x = 0 phosphor to additional energy decay path caused by energy migration between Cr3+ and Cr4+. In x = 1, the concentration of Cr4+ is decreased, while that of Cr3+ is increased. This means the decreased Cr3+–Cr4+ energy transfer in x = 1. In addition, the distance between Cr3+ and Cr3+ is 5.44 Å, which exceeds the exchange interaction distance (<5 Å)45. Thus, the energy transfer process (from Cr3+ to Cr4+ and from Cr3+ to Cr3+) in x = 1 is limited to some extent. This could explain why the decay curve of x = 1 can be well-fitted by the mono-exponential function. The room temperature PL decay curves of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1) phosphors were also recorded (Fig. S8a). Interestingly, the PL decay rate changes non-monotonically with x, but first exhibits a decreasing trend until x = 0.4 and then increase (Fig. S8b). To explain this, the decay curves were fitted by mono-exponential function. The resulted R-Squared (R2) and corresponding lifetime are shown in Fig. 2g. The R2 value first increases and then tends to be unchanged. The increased R2 means better mono-exponential fitting. On the one hand, the energy transfer from Cr3+ to Cr4+ decreases since Cr4+ concentration is lowered after the designed cosubstitution. On the other hand, the structure rigidity increases after cosubstitution, which can decrease the nonradiative transition from the lattice vibration at room temperature. These two aspects should result in the increased R2 value. However, the increased Cr3+ concentration as well as the shortened Cr3+–Cr3+ distance can increase the energy transfer between Cr3+ ions, as illustrated by Fig. 2h. Thus, the R2 value no longer increases at last. These results are consistent with the previous analysis about the reduction of Cr4+ to Cr3+ and improvement of structural rigidity through simple cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+]. As the energy transfer between Cr3+ and Cr4+ is more easily to occur than that between Cr3+ and Cr3+, the lifetime should increase with x as Cr4+ ions are transformed to Cr3+. However, this is not the case. The gradually introduced Zn2+ and Zr4+ ions coexist with Y3+ ions, which causes uneven crystal field around Cr3+, facilitating the breaking of d–d forbidden transition. This explains the shortened lifetime with x from 0 to 0.4. While the lifetime prolongation after x > 0.4 is ascribed to the gradually restored even field as well as the suppressed nonradiation transition by the increased structure rigidity. Figure 2i draws systematical models illustrating the effect of cosubstitution on chromium valence and crystal structure.

As analyzed above, cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+] in this system not only favor the formation of trivalent Cr3+, but also increases the lattice rigid. As a result, the NIR emission of Cr3+ is enhanced as expected. After cosubstitution, the internal quantum efficiency (IQE) is increased from 25% to 96%, and the external quantum efficiency (EQE) is increased from 6% to 20% (Fig. S9). The previously reported Ca3Y2Ge3O12:Cr3+ showed higher IQE (81%) and EQE (10%) than our Ca3Y2Ge3O12:Cr3+46, which is ascribed to the use of higher sintering temperature (1450 °C) and flux (LiF). It should be noted that using flux is also one of the methods to improve luminescence.

As the high structural rigidity of the host means less soft phonon modes that cause nonradiative relaxation process, excellent PL thermal stability of the phosphors can be expected47. The thermal quenching resistance of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1) is investigated (Fig. 3a and Fig. S10), which is greatly improved after the cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+]. As shown in Fig. 3b, x = 1 phosphor shows nearly zero-thermal quenching in the temperature range of 298–373 K. As the temperature elevates to 423 K, the emission intensity still keeps 89% of that at 298 K, which is much higher than that of 59% for x = 0. This confirms the positive effect of the designed cosubstitution on improving structure rigidity. The thermal quenching behavior can be expressed by the configuration coordinate diagram in Fig. 3c. Take x = 0 as an example, O→A and B→C’ represent the normal excitation and emission processes. The excited electrons are thermal activated by the elevated temperature to reach the intersection of the 4T2g and 4A2g curves (B→C) and then return to the 4A2g ground state through nonradiative transition (C→O), leading to the so-called thermal quenching. The activation energy (ΔEa) indicates the difficulty of this thermal activated process, which can be determined by the Arrhenius formula48. The activation energy of x = 1 (0.40 eV) is higher than that of x = 0 (0.29 eV), in consistent with the better thermal stability of x = 1. The valence electrons of Cr3+ bared in outer layer have strong interactions with the lattice environment, which causes large lattice relaxation, display a horizontal displacement (ΔR) of the parabola in configuration coordinate diagram. It is clear that small displacement means large activation energy, which is related to rigid bond49,50. The rigid Cr–O bonds and second coordination shells cause the large activation energy for x = 1, increasing its thermal disturbance resistance. In addition, Huang’s theory points out that the activation energy is dependent on the strength of electron-phonon coupling, which can be reflected by the Huang–Rhys factor (S)51,52. The less spectral broadening with increasing temperature of x = 1 indicates its weaker electron-phonon coupling (Fig. S11). As shown in Fig. 2d, the factor S can be obtained by fitting the full-width at half maximum (FWHM) values of temperature-dependent PL spectra using the following equation:51

$${\rm{FWHM}}=2.36\sqrt{S}\hbar \omega \sqrt{\coth \left(\frac{\hbar \omega }{2{kT}}\right)}$$
(3)

where ω is the phonon energy, k is the Boltzmann constant, and T is the temperature. As a result, the phonon energy values for x = 0 and x = 1 are 0.048 and 0.052 eV, respectively. S values are obtained to be 3.26 for x = 0 and 2.80 for x = 1. The smaller S value for x = 1 implies that it has larger activation energy and higher thermal quenching temperature52. This also demonstrates the dependency of electron–phonon coupling and thermal quenching on structural rigidity. Structural rigidity could be further improved by adopting smaller host cations due to the shorter and more rigid bonds. However, a stronger crystal field would be introduced that generates shorter-wavelength emission. Hence, appropriately sized host cations are essential for high-performance long-wavelength NIR emission.

Fig. 3: Temperature-dependent PL properties of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1).
figure 3

a Temperature-dependent PL spectral profiles of x = 0 and x = 1. b Normalized emission intensity as a function of temperature of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1). c Configuration coordinate curves illustrating the thermal quenching of Cr3+ luminescence. d Fitting results of FWHM as a function of T for x = 0 and x = 1. e Temperature-dependent PL decay times of x = 0 and x = 1. f TL curves of x = 0 and x = 1 measured in the temperature range of 295–550 K. The inset shows the charge transfer process from defects to emission centers. g Ratio of emission intensity under 423 and 298 K (I423/I298) and IQE for some Cr3+-doped garnet phosphors with emission peaking at 770–820 nm. The number 1–11 garnet phosphors are Ca2LaHf2Al3O12:Cr3+, La3Sc2Ga3O12:Cr3+, Gd3MgScGa2SiO12:Cr3+, Lu2CaMg2Ge3O12:Cr3+, Ca2LaZr2Ga2.8Al0.2O12:Cr3+, Ca2LuZr2Al3O12:Cr3+, Ca2YHf2Al3O12:Cr3+, Gd3Zn0.8Ga3.4Ge0.8O12:5%Cr3+, Na3In2Li3F12:Cr3+, Gd3In2Ga3O12:Cr3+, and Ca3Sc2Si3O12:Cr3+

In order to further understand the different thermal quenching behaviors of x = 0 and x = 1, the temperature-dependent PL decay curves were measured (Fig. S12). Both of their PL decay rates accelerate with elevated temperature, which is attributed to the thermal-related nonradiative processes. It is observed from Fig. 3e that x = 1 shows slower decay rate than x = 0 with increasing temperature. In addition to the aforementioned influence of structural rigidity, the influence of thermally activated de-trapping of trapped electrons is also considered. XPS spectra of O 1 s show that oxygen vacancy defects exist in x = 0 and x = 1 (details are available in Supporting Information, Fig. S13), which form shallow traps below the conduction band. The trap depths are estimated to be 0.68 eV for x = 0 and 0.69 eV for x = 1 by the formula of ΔEt = Tm/500, where Tm represents the temperature of the peak center in thermoluminescence (TL) curves plotted in Fig. 3f53,54. At high temperature, the captured electrons in traps are released and charge transfer processes from defects to emission centers occur subsequently (see inset of Fig. 3f), which can compensate partial thermal loss of luminescence. The XPS results show that there are more oxygen vacancy defects in x = 1 (Fig. S13), which results in its higher TL intensity. The more traps in x = 1 can cause more emission compensation and observed better thermal quenching resistance. As shown in Fig. 3g, the excellent thermal stability and high IQE of x = 1 almost surpass all of the Cr3+-doped garnet phosphors within similar emission region (details are listed in Table S6)24,25,30,55,56,57,58,59,60,61,62. Overall, the EQE values of these Cr3+-doped garnet phosphors still need to be improved. Although the as-synthesized phosphors do not show advantage in EQE, this work achieved an improvement in photoluminescence quantum efficiency by the designed chemical unit cosubstitution. This demonstrates the feasibility of the idea about valence conversion and site reconstruction in improving luminescence for Cr3+-doped garnet phosphors.

Chemical stability and information encryption

Previous works have focused on the thermal stability of phosphors, but have paid little attention to their chemical stability, which is also important to practical applications. Here, the influence of water, NaOH (aq.), and HCl (aq.) on x = 0 and x = 1 phosphors were investigated (details are described in Supporting Information, Fig. S14). Both x = 0 and x = 1 show good stability against water and strong base. However, x = 0 and x = 1 show extremely different chemical resistance to HCl (aq.). It was seen that x = 0 quickly decomposed in several seconds, while x = 1 showed better resistance (Fig. S15). As shown in Fig. 4a, b, x = 1 maintains the original phase and PL intensity after soaking in HCl (aq.) for 30 min. Even if it was placed in HCl (aq.) for 12 h, 64% PL intensity could still remain (Fig. S14d). The PL intensity loss is due to the partial destruction of phase by acid, as evidenced by Fig. S14e. This observation suggests more rigid covalent structure of x = 1 phosphor.

Fig. 4: Information encryption and pc-LED applications.
figure 4

a XRD patterns and b PL spectra of x = 1 before (pristine) and after soaking in HCl (aq.) for 30 min. c Schematic diagram of the designed digital information encryption. The white “388” and “301” represent the NIR luminescent images. d Illustration of information encryption of “LINE” letters based on Morse code. The white dots represent the NIR luminescent images. e Practical demonstration of information encryption designed in Fig. 4d. The NIR luminescent images were captured by NIR camera. f Emission spectra and g NIR output power and photoelectric conversion efficiency under driving current in 20–300 mA of the as-fabricated pc-LED using x = 1 phosphor. The insets in Fig. 4f are photos of the as-fabricated pc-LED captured by visible and NIR camera, respectively. h Photos of the NIR pc-LED light penetrating palm, wrist, and fist captured by NIR camera. Photos of plants and fruits captured under i natural light by visible camera and j NIR pc-LED light by NIR camera

Inspired by the different chemical stability of phosphors in HCl (aq.), digital information encryption is designed. As shown in Fig. 4c, the real information is hidden in number “888” consist of by x = 0 phosphor, x = 1 phosphor, and x = 1 host. Under blue-light illumination, wrong information “388” is read out. After adding HCl (aq.) to the number “888”, the parts written by x = 0 phosphor disappear. Therefore, another interference information “801” is displayed. Only with the help of subsequent blue-light illumination, the real information “301” can be read. In addition, a coding method according to Morse code was developed to realize multiple information storage, as shown in Fig. 4d. The Morse code “dot” or “dash” are arranged as the designed coding pattern. After steps (i) – (iii), the information “LINE” can be decrypted. To demonstrate this, hosts x = 0, 1 and phosphors x = 0, 1 were mixed with proper amount of ethyl alcohol to form a slurry as inks, respectively. The Morse code “L” was written by the inks on a slide (Fig. 4e). After first blue-light illumination (step (i)), only two luminescent dots (Morse code “I”) were read. On step (ii), the two dots written by x = 0 host and phosphor disappeared in HCl (aq.). The rest one dash and one dot comprised the Morse code “N”. With blue-light illumination once again (step (iii)), only last one luminescent dot was seen. That is the Morse code “E”. This decryption process of information “LINE” can be completed within 1 min. Further, When HCl (aq.) was sprayed by simply heating, the initial Morse code “L” reappeared (Fig. S16). This reversible signal-switching process enables enhanced security for information encryption and achieves the purpose of “burning after reading”.

LED package and applications

The optimal Ca3ZnZrGe3O12:Cr phosphor was coated on a blue chip (460 nm) to manufacture a NIR-emitting pc-LED. Figure 4f shows its driving current-dependent PL spectra. The emission peaks at around 460 and 800 nm is from the blue LED chip and the phosphor, respectively. It is seen from Fig. 4g that the output powers of total radiance, NIR, and blue light all increase with the increase of current in the range 20–300 mA. More details are listed in Table S7. The NIR output powers reach 34 and 80 mW under 100 and 300 mA, respectively. The conversion efficiency from blue LED to NIR light (ηNIR/blue) decreases from 26% to 23% with current from 20 to 140 mA. The ηNIR/blue values under current beyond 140 mA are absent since the output power of blue LED is out of maximum range of the instrument. The conversion efficiency from input electrical power to NIR light (ηNIR/input) drops from 14% to 7% with current from 20 to 300 mA, which is much lower than ηNIR/blue. This should be ascribed to the unsatisfied photoelectric conversion efficiency from input electrical power to blue light (Table S8). The as-fabricated pc-LED still show comparable photoelectric performance to the reported (Table S6). Once the blue chips become efficient, the ηNIR/input and NIR output power can be further enhanced. As show in Fig. S17a, the emission intensity of the pc-LED only slightly decreases after it being placed in air for 6 months. In addition, after continuous measurement at 100 mA for 12 h, the NIR emission intensity of the pc-LED remains almost unchanged (Fig. S17b). These two results indicate the remarkable stability of the fabricated pc-LED. Figure 4h shows the images of the pc-LED light penetrating palm, wrist, and fist captured by NIR camera. Blood vessels still can be clearly observed when the penetrate depth is 5 cm, which indicates that the as-fabricated NIR-emitting pc-LED is promising for nondestructive examination of deep bio-tissues. The application ability of this pc-LED in night vision is also investigated. Figure 4i shows the photographs of plants and fruits taken by visible camera under natural light. Figure 4j shows the corresponding clear photographs taken by NIR camera under the NIR pc-LED light. Even the small black dot and gap on the leaves, rhizomes of the plants, black spot on the banana peel, and the blue Chinese characters on the plastic bag can be distinguished. These results demonstrate that our NIR pc-LED has great application potential in night vision techniques.

Discussion

In summary, NIR-emitting Ca3Y2-2x(ZnZr)xGe3O12:Cr garnet solid-solution system was established by the designed chemical unit cosubstitution of [Zn2+–Zr4+] for [Y3+–Y3+], which significantly improved the IQE from 25% to 96% and thermal stability (I423 K/I298 K) from 59% to 89%, with only slight emission shift from 812 to 795 nm. The overlarge size-mismatch between Cr3+ and Y3+ contributed to the coexistence of Cr3+ and Cr4+ in Ca3Y2Ge3O12:Cr, which quenched the luminescence of Cr3+. As suggested by DFT calculations, the introduction of smaller [Zn2+–Zr4+] units to Y sites could reconstruct the octahedral sites for Cr3+ ions to promote their formation in Ca3ZnZrGe3O12:Cr, which was responsible for the enhanced luminescence. Moreover, the reconstructed rigid structure also contributed to the high photoluminescence quantum efficiency and thermal stability. Benifing from the weakened electron-phonon coupling effect and the luminescence compensation by traps, Ca3ZnZrGe3O12:Cr exhibited nearly zero-thermal quenching at 373 K and 11% thermal quenching loss at 423 K. In addition, Both Ca3Y2Ge3O12:Cr and Ca3ZnZrGe3O12:Cr showed outstanding water and base resistance, but showed different acid resistance. Information storage was accordingly designed inspired by this difference. Finally, the pc-LED fabricated by Ca3ZnZrGe3O12:Cr demonstrated great potential in bio-tissue imaging and night-vision technologies. This work provides a new perspective of luminescence optimization by chemical unit cosubstitution and could motivate further exploration of high-performance Cr3+-doped garnet phosphors and other types of NIR-emitting phosphor materials.

Materials and methods

Materials synthesis

The designed composition of Ca3Y2-2x(ZnZr)xGe3O12:Cr (x = 0–1) phosphors is Ca3Y1.99-2x(ZnZr)xGe3O12:0.01Cr, which were synthesized by a high-temperature solid-state reaction method. Calcium carbonate (CaCO3, 99.99%), yttrium(III) oxide (Y2O3, 99.99%), zirconium dioxide (ZrO2, 99.99%), germanium dioxide (GeO2, 99.99%), and chromium sesquioxide (Cr2O3, 99.95%) were obtained from Aladdin Reagent Co., Ltd. Zinc oxide (ZnO, A. R.) was acquired from Beijing Chemical Works. Stoichiometrical raw materials were weighed and thoroughly ground in agate mortars for 20 min. Then the mixtures were put into alumina crucibles and sintered at 1673 K for 6 h in a box furnace. The resulting products were slowly cooled down to room temperature and ground again for further characterization.

LED fabrication

The optimal Ca3ZnZrGe3O12:Cr NIR-emitting phosphor was firstly mixed with epoxy resins A and B (mass ratio of 1:1) and then coated on the 460 nm blue chips. The mixtures were cured at 100 °C for 1 h to form the final LED devices.

Characterization

The XRD patterns of the phosphors were measued on a Bruker D8 ADVANCE powder diffractometer using Cu Kα radiation (λ = 1.54 Å) in the 2θ range of 10°−120°. XRD Rietveld refinements were conducted using the GSAS program to reveal the crystal structure and phase purity. The morphology and elemental mapping were obtained by the field-emission scanning electron microscope (FE-SEM, S-4800, Hitachi) equipped with an Energy Dispersive Spectrometer (EDS). Raman spectra were measured by a Raman spectrometer (Model T64000, Horiba JobinYvon, France) using a 512 nm laser. The diffuse reflectance (DR) spectra were recorded on a UV−vis−NIR spectrophotometer (UV-3600 plus, Shimadzu, Japan). The photoluminescence excitation (PLE) and photoluminescence (PL) spectra were measured by the fluorescence spectrometer (Edinburgh Instruments FLSP-920) with a 450 W xenon lamp as excitation source. The light detector R5509-72 photomultiplier (PMT) was used. The PL decay curves were measured by an Edinburgh Instruments FLSP-920 fluorescence spectrometer with a μF2 lamp as excitation source. The temperature-dependent PL spectra and decay curves were also recorded by Edinburgh Instruments FLSP-920 fluorescence spectrometers equipped with a temperature controller. The quantum efficiency measurement was carried out on the measurement system (C9920-02, Hamamatsu photonics K. K., Japan) equipped with a 150 W Xe lamp. The X-ray absorption experiments were carried out at the XAS station (BL14W1) of the Shanghai Synchrotron Radiation Facility. The X-ray photoelectron spectroscopy (XPS) spectra were measured using the Thermo SCIENTIFIC ESCALAB 250Xi with an Al Kα source. The thermoluminescence (TL) spectra were obtained from an LTTL-3DS measurement with a heating rate of 2 K/s. Before TL measurement, each sample was pre-irradiated with a UV lamp (254 nm) for 5 min. The emission spectrum of the as-fabricated pc-LED was measured on the HAAS 2000 photoelectric measuring system from EVERFINE.

Theoretical calculations

Theoretical calculations were performed using density functional theory (DFT) implemented in Vienna Ab-initio Simulation Package (VASP)63. The generalized gradient approximation (GGA) with the PBE functional was applied. The cutoff energy and k-point mesh were set to 450 eV and the 3 × 3 × 3 Monkhorst-Pack grid, respectively. The convergence criterion for the electronic energy was 10−5 eV and the structures were relaxed until the Hellmann–Feynman forces were smaller than 0.02 eVÅ−1.

The Debye temperature ΘD was calculated based on the harmonic Debye model, relying on the bulk modulus and Poisson ratio:64

$${\varTheta }_{{\rm{D}}}=\frac{\hbar }{{k}_{{\rm{B}}}}{\left[6{\pi }^{2}{V}^{1/2}N\right]}^{1/3}\sqrt{\frac{{B}_{{\rm{H}}}}{M}}f\left(v\right)$$
(4)
$$f\left(v\right)={\left\{{\left[{2\left(\frac{2}{3}\cdot \frac{1+v}{1-2v}\right)}^{3/2}+{\left(\frac{1}{3}\cdot \frac{1+v}{1-v}\right)}^{3/2}\right]}^{-1}\right\}}^{1/3}$$
(5)

where kB and are the Boltzmann constant and Planck constant, respectively. BH represents the adiabatic bulk modulus of the crystal, M is the molecular mass of the unit cell, N means the number of atoms per unit cell, V is the unit cell volume, and v is the Poisson ratio.

The formation energy (Ef) of the Cr3+ and Cr4+ doped Ca3Y2Ge3O12 and Ca3ZnZrGe3O12 can be calculated by63:

$${E}_{{\rm{f}}}=E\left({\rm{doped}}\right)-E\left({\rm{perfect}}\right)-\sum {n}_{{\rm{i}}}{{\rm{\mu }}}_{{\rm{i}}}$$
(6)

where E(doped) and E(perfect) are the total energy of the doped and undoped crystal, respectively. The ni and μi represent the chemical potential and the number of the added (ni > 0) or removed (ni < 0) i-type atoms, respectively.