Yixin Sun,
Yining Wang,
Minliang Deng,
Xiaole Xing,
Yiying Zhu and
Mengmeng Shang*
Key Laboratory for Liquid-Solid Structural Evolution and Processing of Materials Ministry of Education, School of Material Science and Engineering, Shandong University, 17923 Jingshi Road, Jinan 250061, P. R. China. E-mail: mmshang@sdu.edu.cn
First published on 2nd August 2024
Near-infrared-II (NIR-II) luminescence, which serves as a promising spectral window for biomedical applications and spectral analysis, has numerous potential applications. In this work, the NIR-II luminescence is realized in Li4SrCa(SiO4)2:Cr4+ phosphors. Under 465 nm excitation, the emission peak of Li4SrCa(SiO4)2:3%Cr4+ is located at 1215 nm and the full width at half maximum (FWHM) is more than 230 nm, covering the ultra-long range of 950–1600 nm. The photoluminescence spectra, diffuse reflectance spectra, Raman spectra and time-resolved emission spectra confirm the NIR-II luminescence mechanism. Furthermore, a novel post-reduction strategy is proposed to realize the leap-forward enhancement of NIR-II luminescence from Cr4+. By mutually substituting Sr2+ and Ca2+, a series of Li4Sr1+zCa1−z(SiO4)2:3%Cr4+ (z = −1 to 1) samples are prepared to investigate the universality and applicable conditions of the post-reduction strategy. Benefitting from the NIR-II emission of Cr4+, it has great application potential in optical temperature measurement and spectral analysis.
Thereinto, lanthanide ions (Yb3+, Nd3+, Er3+, and Ho3+)10–15 and transition metal ions (Ni2+ and Cr4+)16–18 are the main doping ions that can allow NIR-II luminescence. However, the absorption and emission peaks of lanthanide ions with f–f transitions are narrow and inefficient, which is not suitable for NIR spectroscopy. The luminescence of Ni2+ ions is mostly located at 1100–1700 nm, but it is poorly matched with blue LED chips and often requires Cr3+ ions as sensitizers for energy transfer. Notably, the Cr4+ ion matches well with the blue LED chip, and its emission band is located at 1000–1700 nm (the emission peak usually exceeds 1200 nm), which seems to be the most suitable NIR-II region luminescence center.
The Cr4+ ions with 3d2 electronic configuration usually occupy the tetrahedral positions, but Cr3+ ions are inclined to occupy octahedron sites. Consequently, meticulous selection of the host material is imperative to obtain Cr4+-activated NIR phosphors. Recently, Cr4+-activated NIR phosphors have been widely studied, such as Li2ZnGeO4:Cr4+, Li2CaGeO4:Cr4+ and Ca2Ga2SiO7:Cr4+.19–21 Typically, Li2ZnGeO4 has only tetrahedral sites, and is devoid of octahedral sites, yielding an emission peak at 1218 nm in the NIR-II region when Cr4+ is doped into the Li2ZnGeO4 lattice.19 Xia's research group achieved an ultra-long wavelength NIR emission with a luminescence center at 1330 nm in CaYGaO4:Cr4+, exhibiting a FWHM of 233 nm.22 Although [YO6] and [CaO6] octahedrons are present in the CaYGaO4 crystal structure, their ionic radii significantly differ from those of Cr3+/Cr4+, leading to the absence of [CrO6] octahedrons and Cr ions occupying the [GaO4] site in the +4 valence state. Therefore, the choice of matrix is crucial to achieve effective luminescence of Cr4+.
Li4SrCa(SiO4)2 has been developed as an excellent luminescence material host for Eu2+ and Ce3+ doping,23–25 but there is no report on the doping of Cr4+ ions. In the Li4SrCa(SiO4)2 structure, the Li+ and Si4+ ions form a tetrahedron, Sr2+ ions form a polyhedron with a coordination number of 10, and Ca2+ ions form an octahedron. Owing to the significant disparity in charge and size between Ca2+ and Cr3+, [CrO6] is difficult to form in the Li4SrCa(SiO4)2 structure. The presence of [SiO4] enables the formation of [CrO4], where Cr is present in the +4 valence state. Based on the above considerations, the Li4SrCa(SiO4)2 compound is chosen as the matrix compound, and a series of Li4SrCa(SiO4)2:Cr4+ phosphors have been successfully prepared. At a 465 nm excitation wavelength, the Li4SrCa(SiO4)2:Cr4+ emission peak is located at 1215 nm, and the FWHM is more than 230 nm, covering the ultra-wide range of 950–1600 nm. An effective NIR-II luminescence is realized. Additionally, a novel post-reduction strategy is proposed to enhance NIR-II luminescence of Cr4+. Through the complete substitution between Ca2+ and Sr2+, the universality of the post-reduction strategy is verified, and the corresponding applicable conditions are concluded, which is of guiding significance. Thanks to the ultra-long wavelength emission of Li4SrCa(SiO4)2:Cr4+, this material has been verified to have great application potential in optical temperature measurement and spectral analysis.
Fig. 1 (a) The structural model of LSCS. (b) XRD patterns of LSCS:xCr4+ (x = 0, 1% and 5%). The XRD Rietveld refinements of (c) LSCS:0.1%Cr4+ and (d) LSCS:3%Cr4+. |
Fig. 1b presents the X-ray diffraction (XRD) patterns of LSCS:xCr4+ (x = 0, 1%, and 5%) samples prepared in an air atmosphere. All diffraction peaks match well with standard Li4SrCa(SiO4)2 (PDF # 83-0763) data. This confirms the formation of the pure LSCS phase. The XRD Rietveld refinements of LSCS:0.1%Cr4+ and LSCS:3%Cr4+ are carried out, and the refinement results are drawn in Fig. 1c and d. The residual factor values are provided in Table S1 (ESI†). The favorable fitting data not only demonstrate the reliability of the refined structure model but also confirm the high purity of the phase. The lattice parameters are calculated to be V = 693.13 Å3 for LSCS:0.1%Cr4+ and V = 693.22 Å3 for LSCS:3%Cr4+. The increased unit cell volume further indicates that Cr4+ should occupy the site of Si4+ and no existence of Cr3+. This is because there are no octahedral sites suitable for Cr3+ ions to occupy in the LSCS lattice. Even if Cr3+ enters the only [CaO6] octahedron, it would cause a reduction in the unit cell volume.
Energy-dispersive spectrometry (EDS) mapping images of LSCS:3%Cr4+ are provided in Fig. S1 (ESI†). The images reveal uniform distribution of Sr, Ca, Si, O and Cr elements throughout the sample, whereas the presence of the Li element remains undetectable due to its low energy.
Fig. 2a is the photoluminescence excitation (PLE) and photoluminescence (PL) spectra of the LSCS:3%Cr4+ sample prepared in an air atmosphere. Under excitation wavelengths of 465 and 645 nm, the sample emits NIR-II broadband emission from 950 to 1600 nm, with a peak at approximately 1215 nm, which is the characteristic of the 3T2 (3F) → 3A2 spin-allowed transition of Cr4+ ions. Additionally, a continuous PLE spectrum consisting of two bands peaking at 465 nm and 645 nm is observed when monitoring the emission at 1215 nm. These bands correspond to the 3A2 → 3T1 (3P) and 3A2 → 3T1 (3F) energy level transitions of Cr4+ ions, respectively.
Fig. 2b illustrates the PL spectra of LSCS:xCr4+ (x = 0.5–9%) samples. The FWHMs of the PL spectra for all samples are around 235 nm. When excited at a wavelength of 465 nm, both the emission peaks (1215 nm) and spectral shapes remain unchanged with increasing doping amounts. The luminescence intensity of LSCS:xCr4+ samples initially rises with increasing doping concentration, peaking at x = 3% before decreasing due to concentration quenching, as illustrated in Fig. 2c.
Based on the crystal field theory, the Cr4+ ion possesses a 3d2 electronic configuration,28 and its Tanabe–Sugano (T–S) energy level diagram resembles that of 3d8 in an octahedral environment. Therefore, the energy level splitting of LSCS:xCr4+ is delineated by the T–S diagram of the 3d8 configuration within the octahedral field, as depicted in Fig. 2d. The crystal field intensity is calculated from the ratio of the crystal field's splitting energy (Dq) to the Racah parameter (B), as shown in the following equation:19,29,30
10Dq = E(ZPL) | (1) |
(2) |
(3) |
Fig. S2 (ESI†) depicts the luminescence decay curves measured for the LSCS:xCr4+ (x = 1–11%) samples. These curves can be analyzed by using the single exponential decay model, as follows:31
It = I0 + Ae(−t/τ) | (4) |
Fig. 2e shows the PL 3D-mapping spectra of the LSCS:3%Cr4+ sample. Intuitively, in the PL 3D-mapping spectra, there is only one broadband emission at 1215 nm. And the same decay rate of the time-resolved emission (TRS) spectra (Fig. 2f) further indicates that there is only a single Cr4+ ion luminescence center in LSCS:3%Cr4+. It is worth mentioning that increasing the Cr4+ doping concentration enhances the NIR-II luminescence intensity to some extent, but the luminescence intensity of LSCS:3%Cr4+ is still low. Moreover, the internal quantum efficiency (IQE) generally diminishes with the red shift of λmax, presenting a significant challenge for NIR phosphors. Therefore, the IQE of the LSCS:3%Cr4+ sample is poor, only 2.2% (Fig. S4 (ESI†)).
The diffuse reflectance (DR) spectra test is conducted to confirm the existence of Cr6+. Fig. S5a (ESI†) illustrates the DR spectra of LSCS:xCr4+ (x = 0, 0.5% and 3%) samples. Besides the absorption peak of Cr4+ in the 420–1000 nm range, additional absorption peaks at 280 nm and 375 nm are also observed, which are consistent with the absorption peak of Cr6+ reported in the literature.20,33 The Cr6+ has ground state of (3p6 3d0) which transforms into Cr5+ (3p6 3d1) by promoting an electron from 2p6 of O2− to 3d0 of Cr6+ through charge transfer [Cr6+ O2− (3d0 2p6) → Cr5+ O− (3d1 2p5)].34–37 Therefore, the transitions at 280 nm and 375 nm in DR spectra are resulting from the splitting of this charge transfer state in the crystal field. At the same time, the optical band gap (Eg) calculated according to the Tauc and Kubelka–Munk function formulas (eqn (S3) and (S4), (ESI†)) is shown in Fig. S5b (ESI†). As the Cr ion concentration increases, Eg decreases from 5.7 eV to 4.5 eV, indicating the successful doping of Cr. Subsequently, the PLE spectrum of the extended scan range is then re-examined and shows an additional excitation peak at 375 nm (Fig. S5c (ESI†)), consistent with the results of DR spectra. Moreover, Fig. S6 (ESI†) presents X-ray photoelectron spectroscopy (XPS) test results of the LSCS:3%Cr4+ sample, detecting signal peaks for Li 1s, Si 2p, Sr 3d, Ca 2p, O 1s, and Cr 2p. In the high-resolution Cr 2p XPS spectrum, four signal peaks at 574.2 eV, 585.7 eV, 578.9 eV and 590.4 eV are observed, corresponding to the 2p3/2 and 2p1/2 states of Cr4+ and Cr6+, respectively.33,38–40 Therefore, all these results prove the coexistence of Cr4+ and Cr6+ in the sample.
During the initial air sintering process, Cr ions are allowed to be incorporated into the lattice in a hexavalent (+6) state. If a reducing atmosphere is directly adopted during the initial sintering process, it is easy to reduce Cr6+ ions to Cr3+ and even to Cr2+ with lower valence states, and thus it is difficult to obtain Cr4+-activated NIR-II phosphors. So, a post-reduction strategy is used to optimize the luminescence performance of LSCS:Cr4+. That is, after initial air sintering, a H2 reduction atmosphere is introduced during the secondary sintering process to reduce the Cr6+ ions. Fig. 3a illustrates that the luminous intensity of the sample obtained through the post-reduction strategy is significantly improved, approximately 8 times higher than that of the sample sintered in air, whereas the luminous intensity of the sample directly sintering in a reducing (H2) atmosphere during the initial sintering process is the lowest. This is because Cr6+ ions are easily reduced to Cr3+ or even lower valence Cr2+ when sintered only in a reducing (H2) atmosphere. However, Cr3+ ions lack appropriate lattice sites for occupation, thereby hindering the incorporation of Cr3+ ions into the lattice. The post-reduction strategy overcomes this issue by initially sintering in an air atmosphere. The Cr6+ ions are allowed to enter the radius-matched Si4+ site. This process is often accompanied by the creation of cation vacancies or interstitial oxygen defects. Subsequently, secondary sintering in a reducing H2 atmosphere reduces the Cr6+ ions at the Si4+ sites, resulting in efficient Cr4+ luminescence. Simultaneously, the reduction of cation vacancies and interstitial oxygen defects reduces non-radiative transitions, thereby enhancing the luminescence intensity. In addition, the emission spectra sintered under different atmospheres have no obvious changes except the intensity, indicating that the properties of the luminescent center remain unchanged, displaying Cr4+ luminescence.
The PL spectra of post-reduction-Li4SrCa(SiO4)2 (labeled as LSCSH):yCr4+ (y = 1–9%) samples are shown in Fig. S7 (ESI†). The emission peaks are still located at 1215 nm, and the luminescence intensity is improved. Fig. S8 (ESI†) presents the XRD patterns of LSCSH:yCr4+ (y = 0, 1%, and 5%) samples. All diffraction peaks match well with standard Li4SrCa(SiO4)2 (PDF # 83-0763) data. This suggests that the LSCSH:Cr4+ prepared using the post-reduction strategy maintains a pure phase. Fig. S9a and b (ESI†) show the scanning electron microscopy (SEM) images of LSCS:3%Cr4+ and LSCSH:3%Cr4+, respectively. The results show that after the post-reduction strategy, the surface and particle size of the samples remain basically unchanged, indicating that secondary sintering has little effect on the morphology. Furthermore, to eliminate the influence of secondary sintering on luminescence enhancement, three kinds of luminescent materials are prepared at the same sintering temperature, time and sintering times, and only the sintering atmosphere (air + air, air + H2, air + N2) is changed. The emission spectra are shown in Fig. S10 (ESI†). It can be seen that the samples obtained through the post-reduction strategy still exhibit the highest luminescence intensity, followed by those sintered under air + N2 conditions, and the weakest under air + air conditions. Therefore, the enhancement in luminescence intensity achieved by the post-reduction strategy is not due to an increase in crystallinity.
As discussed earlier, the DR spectra (blue curve in Fig. 3b) of the LSCS:0.5%Cr4+ sample reveal distinct absorption peaks attributed to Cr6+ at 280 nm and 375 nm.20,33,41 After applying the post-reduction strategy, there is a notable increase in the absorption peaks (red curve in Fig. 3b) that correspond to the Cr4+ ion. And the absorption peaks of Cr6+ are weakened. This suggests that the post-reduction strategy is highly effective at converting Cr6+ to Cr4+. Meanwhile, the optical band gap Eg calculated according to the Kubelka–Munk function formula42–44 (eqn (S3) and (S4), (ESI†)) is shown in Fig. S11 (ESI†). After reduction, Eg increases from 4.95 (air) to 5.1 eV (air + H2). This alteration is attributed to the post-reduction strategy, which reduces Cr6+ to Cr4+, prompting electron rearrangement within the host material and inducing distortion in the local coordination environment. These factors collectively contribute to the widening of the band gap.45,46 After the post-reduction strategy, the excitation peak of the sample exhibits alterations consistent with the DR spectra. The excitation peak attributed to Cr6+ at 375 nm vanishes, while the intensity and broadness of the excitation peaks (at 465 nm and 645 nm) corresponding to Cr4+ notably increases (Fig. 3c). This occurs because a significant portion of Cr6+ ions occupying the [SiO4] tetrahedra is reduced to Cr4+, with a larger ionic radius through the post-reduction strategy. This reduction induces increased crystal field splitting and lattice distortion, thereby broadening the absorption peaks at 465 nm and 645 nm. Simultaneously, the increase in Cr4+ ion concentration enhances the energy transfer and electron–phonon coupling effects between Cr4+ ions, further contributing to the broadening of the absorption peaks. In addition, the excitation peak of Cr6+ also vanishes in the normalized wavelength-dependent PLE spectra (Fig. S12, (ESI†)).
At the same time, the decay curves (Fig. 3d) can still be fitted by a single exponential function, and the lifetime increases from 8.52 μs (air) to 38.31 μs (air + H2). And the reasons for the increased fluorescence lifetime may include the following aspects: firstly, the post-reduction strategy reduces Cr6+ to Cr4+, which reduces the non-radiative transition and causes increased fluorescence lifetime. Secondly, an increase in Cr4+ concentration improves luminous efficiency and fluorescence lifetime. Finally, changes in the local environment of the central Cr4+ ion promote energy level splitting and increase the radiative transition probability, thereby further extending the fluorescence lifetime. In general, the increase in fluorescence lifetime is the result of a combination of factors. The post-reduction strategy can effectively extend the fluorescence lifetime of LSCSH:Cr4+ by reducing non-radiative transitions, optimizing the doping concentration and local environment.
Through the post-reduction strategy, the IQE exhibits a notable enhancement, from 2.2% to 27% (Fig. S13, (ESI†)). Additionally, the radius matching principle is exceptionally important in doped luminescent materials, even playing a decisive role, according to the distortion index Dr:47
(5) |
Fig. 3e and Fig. S14 (ESI†) illustrate the XRD Rietveld refinement results of LSCSH:yCr4+ (y = 0.1% and 3%). After post-reduction, LSCSH:yCr4+ still maintains the physical phase of Li4SrCa(SiO4)2. The related results are presented in Table S3 (ESI†). Notably, the calculated unit cell volumes are recorded as V = 693.99 Å3 (y = 0.1%) and V = 694.02 Å3 (y = 3%), both of which are larger than the unit cell volume before post-reduction (air, V = 693.13 Å3, LSCS:0.1%Cr4+ and V = 693.22 Å3, LSCS:3%Cr4+). This means that the post-reduction strategy reduces Cr6+ with a smaller ionic radius to Cr4+ with a larger ionic radius, which is consistent with the spectral analysis results.
To analyze the effect of the post-reduction strategy on the local environment of Cr ions, the Raman spectra of typical samples LSCS:3%Cr4+ and LSCSH:3%Cr4+ are shown in Fig. 3f. Six distinct bands at 383, 430, 812, 844, 863, and 897 cm−1 are observed in the Raman spectra. The Raman peak at 383 cm−1 is attributed to the vibrational bending mode of the Si–O–Si group, while the intense Raman peaks in the range of 810–865 cm−1 correspond to the stretching mode of the Si–O bonds. Additionally, the Raman peaks in the 865–1000 cm−1 range are associated with the stretching mode of the Ca–Si–O and Sr–Si–O groups.23,27 Compared with LSCS:3%Cr4+, the Raman spectra of LSCSH:3%Cr4+ reveal a noticeable decrease in intensity and symmetry of the vibrational bending mode of the Si–O–Si group and the stretching mode of the Si–O bond. This phenomenon supports that the post-reduction strategy facilitates the reduction of Cr6+ ions located in [SiO4] sites to Cr4+ ions with larger ionic radii, resulting in an increase in lattice distortion. This distortion not only reduces lattice symmetry but also alters the local environment (such as charge rearrangement). Additionally, variations in crystal field strength and symmetry impact the transition probability between the excited state and ground state of the doped ions, thereby affecting the luminescence intensity. In summary, the post-reduction strategy significantly enhances the luminescence intensity, lifetime and quantum efficiency of LSCSH:Cr4+ luminescent materials.
After the post-reduction strategy, the luminescence intensities of H-Li4Sr1+zCa1−z(SiO4)2:3%Cr4+ (−1 < z < −0.4) samples improved, and the PL spectra are illustrated in Fig. S16 (ESI†). The effect of the post-reduction strategy (Fig. 4a) on improving the luminescence intensity of Cr4+ is very significant. In addition, the PLE and PL spectra of H-Li2SrSiO4:3%Cr4+ and H-Li2CaSiO4:Cr4+ samples are shown in Fig. 4b. When excited by light at 465 nm, the emission peak of the H-Li2CaSiO4:3%Cr4+ sample is situated at 1150 nm, spanning the NIR-II region from 900 to 1600 nm. In contrast, the emission wavelength of H-Li2SrSiO4:3%Cr4+ undergoes a red shift to 1260 nm because the d orbitals of Cr4+ ions experience different energy level splits under varying crystal fields, which influences the absorption and emission. Although the crystal structures of Li2SrSiO4 and Li2CaSiO4 samples differ from that of Li4SrCa(SiO4)2, the luminescence intensities are also notably enhanced after the post-reduction strategy. As depicted in Fig. S17 (ESI†), the PL intensities increase 34 times (Li2SrSiO4:3%Cr4+) and 15 times (Li2CaSiO4:3%Cr4+) of those obtained under the air-sintering conditions. Hence, the post-reduction strategy proves to be universally effective in enhancing Cr4+ luminescence. In summary, the applicable conditions for obtaining efficient Cr4+-activated NIR phosphors by the post-reduction strategy are as follows: (i) the matrix should lack suitable sites for Cr3+ occupation to prevent facile reduction of Cr ions to trivalent (+3) states. (ii) In the [AO4] tetrahedron occupied by Cr4+ (CN = 4, r = 0.41 Å), the radius of the cation Ax+ should be smaller than that of Cr4+. The larger radius difference between Ax+ and Cr4+, the better the post-reduction strategy, such as Si4+ (CN = 4, r = 0.26 Å), Ge4+ (CN = 4, r = 0.39 Å) and Al3+ (CN = 4, r = 0.39 Å).
Fig. 4 (a) Luminescence peak intensity of Li4Sr1+zCa1−z(SiO4)2:3%Cr4+ (z = −0.4 to 0.5) under different sintering conditions. (b) The PLE and PL spectra of H-Li2CaSiO4:3%Cr4+ and H-Li2SrSiO4:3%Cr4+. |
Evidently, the LSCSH:3%Cr4+ sample, characterized by ultra-long wavelength emission, demonstrates a significant Stokes shift that renders it highly sensitive to temperature variations, thereby indicating poor thermal stability.51 The activation energy (ΔE) can further evaluate thermal stability and can be calculated by using the Arrhenius formula52–55 (eqn (S5) (ESI†)). The calculated ΔE = 0.19 eV for LSCSH:3%Cr4+ is shown in Fig. 5d. Electron–phonon coupling induces cross-relaxation, causing electrons to return from the excited state to the ground state via nonradiative transitions rather than radiative transitions. The energy barrier ΔE necessary for these nonradiative transitions is illustrated in Fig. S19 (ESI†). The extremely small ΔE = 0.19 eV of LSCSH:3%Cr4+ indicates that electrons readily surmount the energy barrier, leading to more significant dissipation of energy in the form of heat through non-radiative transitions.
In luminescence temperature measurement, emission intensity, lifetime and emission spectrum shape are the three most commonly used features. For NIR-II luminescent materials, the environmental background luminescence has minimal impact, providing distinct advantages in optical temperature measurement. The spectral shift (Δλ) and decay time (τ) are chosen as indicators for optical temperature measurement in LSCSH:3%Cr4+ luminescent materials, relying on its temperature-dependent luminescence properties. The absolute sensitivity (SA) and relative sensitivity (SR) are used to assess the optical temperature sensor's responsiveness to temperature changes. The fitting formula is as follows:22,56–59
(6) |
(7) |
Fig. 6 The transmission spectra for ethanol, water and peanut oil. The calculated transmission spectra are shown below. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4tc02781g |
This journal is © The Royal Society of Chemistry 2024 |