Next Article in Journal
Design of Multifunctional Janus Metasurface Based on Subwavelength Grating
Previous Article in Journal
Simultaneous Removal of Arsenic and Manganese from Synthetic Aqueous Solutions Using Polymer Gel Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nano-Crystallization of Ln-Fluoride Crystals in Glass-Ceramics via Inducing of Yb3+ for Efficient Near-Infrared Upconversion Luminescence of Tm3+

Guangdong Provincial Key Laboratory of Optical Fiber Sensing and Communications, Institute of Photonics Technology, Jinan University, Guangzhou 511443, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(4), 1033; https://doi.org/10.3390/nano11041033
Submission received: 31 March 2021 / Revised: 12 April 2021 / Accepted: 16 April 2021 / Published: 18 April 2021

Abstract

:
Transparent glass-ceramic composites embedded with Ln-fluoride nanocrystals are prepared in this work to enhance the upconversion luminescence of Tm3+. The crystalline phases, microstructures, and photoluminescence properties of samples are carefully investigated. KYb3F10 nanocrystals are proved to controllably precipitate in the glass-ceramics via the inducing of Yb3+ when the doping concentration varies from 0.5 to 1.5 mol%. Pure near-infrared upconversion emissions are observed and the emission intensities are enhanced in the glass-ceramics as compared to in the precursor glass due to the incorporation of Tm3+ into the KYb3F10 crystal structures via substitutions for Yb3+. Furthermore, KYb2F7 crystals are also nano-crystallized in the glass-ceramics when the Yb3+ concentration exceeds 2.0 mol%. The upconversion emission intensity of Tm3+ is further enhanced by seven times as Tm3+ enters the lattice sites of pure KYb2F7 nanocrystals. The designed glass ceramics provide efficient gain materials for optical applications in the biological transmission window. Moreover, the controllable nano-crystallization strategy induced by Yb3+ opens a new way for engineering a wide range of functional nanomaterials with effective incorporation of Ln3+ ions into fluoride crystal structures.

1. Introduction

Trivalent lanthanide (Ln3+) ions doped upconversion (UC) luminescent materials have been extensively investigated for applications in lighting, solar cells, solid-state laser, biological imaging, and temperature sensing due to the multiple luminescence wavelength from ultraviolet to near-infrared (NIR) regions [1,2,3,4,5,6,7,8,9]. Among these, Tm3+ doped materials possess UC luminescence in the NIR region at around 800 nm (3H43H6), which is located in the biological transmission window, and the light in that region can easily penetrate biological tissues [10,11,12,13]. Moreover, the UC emissions of Tm3+ are dramatically enhanced via the energy transfer (ET) from Yb3+, which possesses a large absorption section for the commercial 980 nm lasers. Thus, Yb3+-Tm3+ co-doped materials were more efficient candidates for NIR-to-NIR UC luminescence and were significant light sources for biologically non-destructive detection. In the past decades, a large number of investigations about Yb3+-Tm3+ co-doped UC materials have been widely reported for achieving high-efficiency NIR UC luminescence [14,15,16,17,18,19,20].
Glass exhibits high transmittance, easy fabrication, and excellent processability due to its amorphous and softening characteristics. These properties make glass an appealing matrix for the design and fabrication of special optics devices, particularly of optical fiber lasers [21,22], which can hardly be achieved by using crystals. Furthermore, nano-crystallized glass ceramic (GC) composites can be obtained via the precipitation of functional nanocrystals in a glass matrix via heat treatments. The nano-crystallized GCs still possess high optical transmittance and the UC luminescence efficiency in GCs can be greatly enhanced, as compared to glass, when Ln3+ ions are successfully incorporated into the crystal structures featuring lower probability of non-radiative transition ascribed to the lower phonon energy. So far, Yb3+-Tm3+ co-doped GCs have been considered as significantly optical gain materials for UC NIR fiber lasers, noncontact optical thermometers, and bio-imaging [23,24,25,26]. Generally, Ln3+ ions were expected to incorporate into the crystal structures via the ionic substitution for Y3+, La3+, Lu3+, Sc3+, and Sr2+ in the GCs embedded with Na(Y/La/Lu)F4, LaF3, YF3, KSc2F7, and SrF2 crystals, etc [27,28,29,30,31]. However, the ionic substitution process was uncontrollable and the amount of Ln3+ incorporated into the crystal structures was usually small due to the large mismatch between Ln3+ and the replaceable sites. The enhancements of luminescence in GCs were limited. Moreover, a lot of remaining crystals not occupied by Ln3+ ions made no contribution to the enhancement of luminescence but will trigger severe issues of optical scattering and low optical transmittance in glass. These made the traditional GCs difficult to be used in the practical applications. Accordingly, it is highly desirable for design and fabrication of novel GCs for controllably incorporating a large number of Ln3+ into crystal structures to achieve high-efficiency NIR UC luminescence of Tm3+.
Ln-fluoride crystal, intrinsically containing Ln3+ (for example, Yb3+ or Er3+) in the crystal lattices, is a significant host for the incorporation of other active Ln3+ ions because the mismatches in ionic radius between different Ln3+ ions are smaller and active Ln3+ ions enter Ln-fluoride crystals more easily [32]. In this work, GCs containing Ln-fluoride nanocrystals were designed to greatly enhance the UC luminescence of Tm3+ and achieve pure, high-efficiency NIR-to-NIR emissions. Two kinds of Ln-fluoride nanocrystals (KYb3F10 and KYb2F7) were controllably precipitated from a glass matrix via the heat treatments dependent on the doping of Yb3+. Tm3+ ions entered the fluoride crystal lattices easily by replacing Yb3+ sites. As a result, a large number of Ln3+ were incorporated into the fluoride crystals featuring extremely low phonon energy and the UC emission of Tm3+ in the GCs were dramatically enhanced. Moreover, the pure NIR UC emissions were also obtained by adjusting the doping concentration of Tm3+. Thus, the designed GCs provide an efficient material for pure NIR UC luminescence and offer highly promising developments for UC fiber lasers, non-contact optical thermometers, and bio-imaging.

2. Materials and Methods

2.1. Materials Preparation

Samples with a molar composition of 70SiO2-15KF-15ZnF2-xYbF3-yTm2O3 (x = 0~3.5, y = 0~0.3) were prepared using a melt quenching method. First, 30 g reagent grade stoichiometric mixtures were mixed thoroughly in an agate mortar and melted in covered quartz crucibles in an electric furnace at 1550 °C for 30 min; the melts were poured onto a brass plate and then pressed using another brass plate to obtain precursor glasses (PGs). PG samples were heated at 540 °C for 5 and 10 h to obtain GC samples according to the differential scanning calorimetry (DSC) results in ref. [33]. PG and GC samples were cut and polished to 2 mm thick for measurements.

2.2. Characterizations

To identify the crystalline phase in GCs, X-ray diffraction (XRD) patterns were performed on a X-ray diffractometer (Bruker, Fällanden, Switzerland) with Cu/Ka (λ = 0.1541 nm) radiation. The morphology and size distribution of the nanocrystals in GCs were measured via high-resolution transmission electron microscopy (HRTEM) (FEI, Hillsboro, OR, USA). UC emission spectra of samples were recorded using an Edinburgh FLS980 fluorescence spectrometer (Edinburgh Instruments, Edinburgh, UK). A 980 nm laser diode (LD) was used as the exciting source for the measurement of UC emission spectra. The emission decay curves were measured using the same spectrometer with a microsecond lamp as the excitation source. All measurements were performed at room temperature.

3. Results and Discussion

Figure 1a shows the XRD patterns of Yb3+-Tm3+ co-doped PG and GCs. A broad band is observed in the XRD pattern of PG due to the amorphous characteristic of glass. Sharp peaks are observed in the XRD patterns of the two GCs. The main peaks at 26.99°, 31.27°, 44.81°, and 53.1° are ascribed to the (222), (400), (440), and (622) crystal facets of KYb3F10 (No: 74-2204), respectively, indicating that KYb3F10 crystals have been precipitated in the GCs. Moreover, weak peaks at 18.86°, 31.04°, and 38.28° attributed to the diffraction peaks of K2SiF6 (No: 85-1382) crystals are also observed in the XRD patterns of GC heated at 540 °C for 5 h. It is also found that the intensities of diffraction peaks for KYb3F10 crystal are all increased when the heating time increases from 5 to 10 h. However, the diffraction peaks of K2SiF6 crystal all weaken due to the further heat treatment to 10 h. These results indicate that KYb3F10 crystals prefer to precipitate in the 1.0Yb3+-0.005Tm3+ co-doped GCs via the further heat treatment. Additionally, the average size of crystals can be calculated by the Scherrer’s equation. The diffraction peak, around 2θ = 27.01° and 18.42°, was selected for the calculation, and the average size of the KYb3F10 and K2SiF6 nanocrystals in the GC heated at 540 °C for 5 h was calculated to be approximately 13.35 and 31.34 nm, respectively.
The HAADF-TEM image shown in Figure 1b reveals that the nanocrystals are in-situ, precipitated among the glass matrix. The measured size of these nanoparticles is from 10 to 30 nm. The HR-TEM image is shown in Figure 1c. The crystal lattice fringes are obvious, which is different from that of the amorphous glass matrix. The interval of the crystal lattice fringes d can be measured directly, and its value is about 0.286 nm, which corresponds to the (400) crystal facet of cubic KYb3F10. These also prove that KYb3F10 nanocrystals are precipitated in the GCs.
The optical transmission spectra of the 1.0Yb3+-0.1Tm3+ co-doped samples are shown in Figure 2a. The PG sample possesses high transmittance (~90%) from 300 to 800 nm with a thickness of 2 mm. Though nanocrystals are precipitated in the GCs, the optical transmittances of GCs are still as high as 85%. Interestingly, the transmittance of the GC heated for 10 h is higher than that heated for 5 h. As calculated from the XRD patterns, the average size of KYb3F10 crystals is smaller than that of K2SiF6 crystals. When the heat treatment time increases from 5 to 10 h, more KYb3F10 crystals and less K2SiF6 crystals are precipitated in the GC, resulting in the increase of transmittance.
Figure 2b shows the emission spectra of 1.0Yb3+-0.005Tm3+ co-doped PG and GCs recorded at room temperature upon the excitation of a 980 nm laser diode (LD). The spectra all include a NIR emission peak at 802 nm attributed to the 3H43H6 transition of Tm3+. Meanwhile, the blue (450 and 480 nm), red (650 nm), and deep red emission peaks (680 and 700 nm) attributed to 1D23F4, 1G43H6, 1G43F4, 3F23H4, and 3F33H4 transition of Tm3+ [34,35], respectively, are all observed in the emission spectra of PG and GC samples. Since Tm3+ exhibits no absorption for 980 nm light, these emissions are all obtained via the ET processes from Yb3+ to Tm3+. In addition, the emission peaks of GCs are narrower than that of PG and the splitting of spectrum by crystal fields are observed in the spectra of GCs, indicating the incorporation of Tm3+ into crystal coordinated sites. It is also found from the spectra that the emission intensity at 450 and 802 nm increases 15 and 6 times after the heat treatment for 10 h, respectively. Figure 2c,d show the emission decay curves monitored at 450 and 802 nm, respectively. The emission lifetime at 450 nm increases from 441 to 601 and 719 μs, and that at 802 nm increases from 393 to 658 and 887 μs when the sample is heated at 540 °C for 5 and 10 h, respectively. These results are attributed to the fact that the low phonon energy environment reduces the possibility of non-radiative transition and enhances the emission intensity and lifetime in the sample, which in turn prove the incorporation of Tm3+ into the fluoride nanocrystals in GCs. Actually, it is difficult for Tm3+ to enter K2SiF6 crystal structures due to the lack of appropriate sites for the substitution of Tm3+. However, Tm3+ can easily incorporate into KYb3F10 crystal structures by occupying the sites of Yb3+ because the ionic radiuses of Tm3+ (R = 0.086 nm) and Yb3+ (R = 0.085 nm) are very similar. Accordingly, the incorporation of Tm3+ into KYb3F10 nanocrystals is responsible for the enhancements of emission intensities and lifetimes in GCs.
The double-logarithmic plots of the excitation power dependency on the emission intensities are presented in Figure 3a. The fitted slope (n) of the plot of Ln (IUC) versus Ln (Power) is used to determine the absorbed photon numbers per UC emitted [36,37]. The obtained value of n was 1.67, 1.73, 2.67, and 3.40 corresponding to the 802,700, 650, and 450 nm emission of 1.0Yb3+-0.005Tm3+ co-doped GC, respectively. These prove that the emissions are all attributed to UC emission of Tm3+ and obtained via the ET from Yb3+. Two, two, three, and four pump photons are needed to pump the electrons to 3H4, 3F2,3, 1G4, and 1D2 energy levels of Tm3+ to achieve the corresponding UC emissions, respectively.
Based on these, the energy levels of Tm3+ and Yb3+ and the relative transitions corresponding to the above UC emissions are illuminated in Figure 3b [38,39,40]. Under 980 nm LD excitation, the electrons of Yb3+ are excited from the ground state 2F7/2 to the 2F5/2 excited state, then the energy is transferred to Tm3+ through the ET1 process: [2F5/2 (Yb3+) + 3H6 (Tm3+)→2F7/2 (Yb3+) + 3H5 (Tm3+) + phonons]. The electrons of Tm3+ are excited from the ground state to the 3H5 excited state and then undergo a non-radiative relaxation transition to the 3F4 energy level. Via the ET2 process, [2F5/2 (Yb3+) + 3F4 (Tm3+)→2F7/2 (Yb3+) + 3F2,3 (Tm3+) + phonons], the electrons are populated to the 3F2, 3 energy level. A part of the electrons undergo a non-radiative relaxation transition to 3H4 energy level. In this process, 3H43H6 (802 nm) and 3F2,33H6 (700 and 680 nm) transitions are obtained. Next, part of the electrons at 3H4 energy levels are excited to 1G4 energy through the process of ET3: [2F5/2 (Yb3+) + 3H4 (Tm3+)→2F7/2 (Yb3+) + 1G4 (Tm3+) + phonons]. This produces 1G43H6 (480 nm) and 1G43F4 (650 nm) transitions. Finally, the electrons of the 1G4 energy level are populated to the 1D2 energy level through the process of ET4, [2F5/2 (Yb3+) + 1G4 (Tm3+)→2F7/2 (Yb3+) + 1D2 (Tm3+) + phonons], producing a 1D23F4 (450 nm) transition.
Furthermore, the transitions of Tm3+ are also modulated via the interaction between neighboring Tm3+ ions, which is directly determined by the doping concentration of Tm3+ ions [41,42]. The dependence of UC emission spectrum on the doping concentration of Tm3+ in the co-doped PG and GC is presented in Figure 4a,b. It can be observed from the spectra that the blue emission intensities at 450 nm and 480 nm both decrease monotonically with the increase of Tm3+ concentration, while the NIR emission intensity at 802 nm experiences first an enhancement and then a decrease due to the further increase of Tm3+ concentration. Actually, as a result of the increase of Tm3+ content, the distance between Tm3+ becomes shorter. The interactions between neighboring Tm3+ ions get larger, which benefits the cross-relaxation (CR) process, 1G4 + 3H63F2,3 + 3F4, as shown in Figure 3b. The CR process depopulates electrons in the 1G4 level, while the electrons in 3F2,3 and 3F4 levels are populated via the CR process. Thus, blue emission intensities decrease quickly with the increase of Tm3+ even though the doping concentration is low. The NIR emission increases firstly and then decreases when the doping concentration is further increased due to the concentration quenching effect. Compared to a glass matrix, crystal exhibits a more compact structure. The distance between Tm3+ in GC is shorter than that in PG at the same doping concentration, resulting in the lower luminescence-quenching concentration in GC as compared to PG. The emission intensity at 802 nm reaches a maximum with the doping of 0.025 and 0.01Tm3+ in PG and GC, respectively. As shown in Figure 4b and the inset, the blue emissions around 450 and 480 nm in PG and GC both disappear and the UC spectra almost exhibit pure NIR emissions when the contents of Tm3+ exceed 0.01 mol%. A blue UC emission only being obtained in GC at a low doping concentration indicates most Tm3+ ions are incorporated into the KYb3F10 crystal structure with short interionic distance. The pure NIR emissions that are obtained at a very low doping concentration of Tm3+ indicates that the GC is an excellent candidate for pure NIR UC emission and is a significant UC luminescent material for promising applications in NIR fiber lasers and high-resolution biological imaging.
As mentioned above, Yb3+ works as sensitizer ion for the UC emissions of Tm3+. More importantly, Yb3+ also participates in the construction of fluoride nanocrystals and provides appropriate crystal sites for the incorporation of Tm3+. The doping of Yb3+ plays a vital role in the precipitation of nanocrystals and enhancement of the UC emissions. The Yb3+-concentration-dependent XRD patterns of the Yb3+-Tm3+ co-doped and no-doped GCs are presented in Figure 5. It is found that only K2SiF6 crystals are precipitated in the no-doped GC and these make no contribution to the enhancement of UC emission. Via the doping of Yb3+, KYb3F10 are precipitated in the GCs apart from K2SiF6 crystals as the Yb3+ content increases from 0.5 to 1.0 mol%. When Yb3+ concentration is set to 1.5 mol%, only the diffraction peaks of KYb3F10 crystals are observed in the XRD pattern of GC. These results indicate that the precipitation of KYb3F10 nanocrystals in the GCs is governed by the doping concentration of Yb3+ induced by Yb3+. More interestingly, KYb2F7 crystals are also precipitated in the GCs when the Yb3+ concentration is increased from 2.0 to 3.5 mol%. In the XRD pattern of the 3.5Yb3+-0.1Tm3+ co-doped sample, only the diffraction peaks of KYb2F7 crystals are observed. Therefore, the nano-crystallization in GC is induced by the doped Yb3+ and the crystalline phase in the GC can be controllably regulated from K2SiF6 to KYb3F10 and further changes to KYb2F7 crystals by adjusting the concentration of Yb3+ from 0 to 3.5 mol%.
Via the nano-crystallizations of KYb3F10 and KYb2F7, Yb3+ ions are spontaneously confined within fluoride crystal structures. More importantly, Tm3+ can easily incorporate these two crystal structures by the substitution of Yb3+. Owing to the extremely low phonon energies in the fluoride crystals, the probabilities of non-radiative transitions are low, which is beneficial for the efficient ET from Yb3+ to Tm3+ and enhanced UC emission of Tm3+. The UC emission spectra of Yb3+-Tm3+ co-doped GC with various concentrations of Yb3+ are shown in Figure 6a. Excited by a 980 nm LD, pure NIR UC emissions around 802 nm of Tm3+ are observed in the spectra of GCs. As the Yb3+ concentration increases from 0.5 to 1.5 mol%, the quantity of KYb3F10 nanocrystals precipitated in the GCs is increased, and the amount of Tm3+ and Yb3+ ions located in the fluoride crystal structures are increased. The emission intensity enhances firstly, reaching a maximum at 1.0 mol% Yb3+, and then decreases when the Yb3+ concentration is increased to 1.5 mol% due to the concentration quenching effect caused by the severe interactions between Ln3+ ions. When the Yb3+ concentration is further increased to 2.0 mol%, KYb2F7 crystals start to precipitate in the GCs. A part of the Tm3+ enter KYb2F7 nanocrystals in the 2.0Yb3+-0.1Tm3+ co-doped GC, which disperses the distribution of Tm3+ and increases the NIR UC emission intensity again. When the Yb3+ concentration increases from 2.0 to 3.5 mol%, more KYb2F7 crystals are precipitated in the GCs, more Tm3+ ions are incorporated into KYb2F7 crystals, and the UC emission intensity at 800 is increased monotonically. It is also found that the profile of the emission spectrum for the 3.5 Yb3+-0.1Tm3+ co-doped GCs is different from that of the 1.0Yb3+0.1Tm3+ co-doped samples (inset of Figure 6a), which in turn proves that Tm3+ are incorporated in two different crystal environments in the two GCs containing different nanocrystals. Compared to KYb2F7 crystal, more Yb3+ ions are distributed in the crystal structure in KYb3F10. The luminescent-quenching concentration of the UC emission in KYb2F7 crystal is higher than that in KYb3F10. The emission intensity in 3.5Yb3+-0.1Tm3+ co-doped GC is 7 times higher than that in the 1.0Yb3+-0.1Tm3+ sample, which indicates that the nano-crystallization of KYb2F7 provides more excellent crystal environments for achieving more efficient NIR UC emission of Tm3+.
The UC emission decay curves of the Yb3+-Tm3+ co-doped GCs monitored at 802 nm are shown in Figure 6b. The lifetime of the NIR emission decreases as Yb3+ concentration increases from 0.5 to 1.5 mol%. Especially, the lifetime of the 1.5Yb3+-0.1Tm3+ co-doped GC decreases dramatically to 373 μs due to the severe interaction between the neighboring Tm3+ in KYb3F10 crystals. However, the lifetime increases from 16.5 to 39.8 μs when the Yb3+ concentration is increased from 2.5 to 3.5 mol%. These results also prove that the coordinated environments of Tm3+ in high-doping GCs (2.5 and 3.5 mol% Yb3+-0.1Tm3+) is distinct to that in low-doping GCs (0.5, 1.0, and 1.5 mol% Yb3+-0.1Tm3+). Therefore, the precipitation of KYb2F7 nanocrystals in GCs further enhances the NIR UC emission of Tm3+ and provides a significant matrix for applications in photonic devices, in particular of high-efficiency UC fiber lasers.

4. Conclusions

In summary, transparent Yb3+-Tm3+ co-doped GCs were prepared in this work for greatly enhancing the UC luminescence of Tm3+. The GCs possessed high transmittance (>85.0%) though fluoride nanocrystals were precipitated among the glass matrices. K2SiF6 crystals are precipitated in the no-doped GC. However, KYb3F10 and KYb2F7 nanocrystals were controllably and successively precipitated in the GCs via the inducing of Yb3+ when the doping concentration of Yb3+ changed from 0.5 to 3.5 mol%. UC emissions of Tm3+ were dramatically enhanced when Ln3+ ions were incorporated into KYb3F10 crystal structures as the Yb3+ concentration changed from 0.5 to 1.5 mol%. Pure NIR UC emissions were obtained by adjusting the concentration of Tm3+. KYb2F7 nanocrystals were precipitated in GCs when the Yb3+ concentration exceeded 2.0 mol%. More efficient UC emissions were achieved via the precipitation of KYb2F7 than that of KYb3F10 in GCs. The designed GCs offer potential optical gain materials for highly-efficient NIR UC photoluminescence. More importantly, the Yb3+-induced nano-crystallization strategy paves a new way for the design and fabrication of emerging GCs to provide excellent crystal environments for Ln3+.

Author Contributions

Conceptualization, Z.F. and J.L.; methodology, J.L.; Y.L. and Q.Z.; formal analysis, J.L.; Y.L.; Q.Z. and S.Z.; investigation, J.L.; Y.L.; Q.Z. and Z.F.; resources, S.Z.; Z.F. and B.-O.G.; data curation, J.L.; Y.L.; Q.Z. and Z.F.; writing-original draft, J.L.; writing-review and editing, S.Z.; Z.F. and B.-O.G.; supervision, Z.F. and B.-O.G.; project administration, Z.F. and B.-O.G.; funding acquisition, Z.F.; S.Z. and B.-O.G. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support from the National Natural Science Foundation of China (No. 61905093, 61805105). This work was also supported by the Open Fund of the Guangdong Provincial Key Laboratory of Fiber Laser Materials and Applied Techniques (South China University of Technology) and the Fundamental Research Funds for the Central Universities (21619340).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, J.; Gu, F.; Liu, X.; Qiu, J. Enhanced multiphoton upconversion in single nanowires by waveguiding excitation. Adv. Opt. Mater. 2016, 4, 1174–1178. [Google Scholar] [CrossRef]
  2. Zhou, J.; Leano, J.L., Jr.; Liu, Z.; Jin, D.; Wong, K.L.; Liu, R.S.; Bunzli, J.G. Impact of lanthanide nanomaterials on photonic devices and smart applications. Small 2018, 14, e1801882. [Google Scholar] [CrossRef]
  3. Wen, S.; Zhou, J.; Zheng, K.; Bednarkiewicz, A.; Liu, X.; Jin, D. Advances in highly doped upconversion nanoparticles. Nat. Commun. 2018, 9, 2415. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, F.; Han, Y.; Lim, C.S.; Lu, Y.; Wang, J.; Xu, J.; Chen, H.; Zhang, C.; Hong, M.; Liu, X. Simultaneous phase and size control of upconversion nanocrystals through lanthanide doping. Nature 2010, 463, 1061–1065. [Google Scholar] [CrossRef]
  5. Wang, F.; Wang, J.; Liu, X. Direct evidence of a surface quenching effect on size-dependent luminescence of upconversion nanoparticles. Angew. Chem. Int. Ed. Engl. 2010, 49, 7456–7460. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, J.; Deng, R.; MacDonald, M.A.; Chen, B.; Yuan, J.; Wang, F.; Chi, D.; Hor, T.S.; Zhang, P.; Liu, G.; et al. Enhancing multiphoton upconversion through energy clustering at sublattice level. Nat. Mater. 2014, 13, 157–162. [Google Scholar] [CrossRef]
  7. Wang, J.; Wang, F.; Wang, C.; Liu, Z.; Liu, X. Single-band upconversion emission in lanthanide-doped KMnF3 nanocrystals. Angew. Chem. Int. Ed. Engl. 2011, 50, 10369–10372. [Google Scholar] [CrossRef]
  8. Zhou, B.; Shi, B.; Jin, D.; Liu, X. Controlling upconversion nanocrystals for emerging applications. Nat. Nanotechnol. 2015, 10, 924–936. [Google Scholar] [CrossRef]
  9. Auzel, F. Upconversion and Anti-Stokes Processes with f and d Ions in Solids. Chem. Rev. 2004, 104, 139–174. [Google Scholar] [CrossRef]
  10. Chen, G.Y.; Ohulchanskyy, T.Y.; Kumar, R.; Agren, H.; Prasad, P.N. Ultrasmall monodisperse NaYF4:Yb3+/Tm3+ nanocrystals with enhanced Near-Infrared to Near-Infrared upconversion photoluminescence. Acs Nano 2010, 4, 3163–3168. [Google Scholar] [CrossRef] [Green Version]
  11. Fernandez-Bravo, A.; Yao, K.; Barnard, E.S.; Borys, N.J.; Levy, E.S.; Tian, B.; Tajon, C.A.; Moretti, L.; Altoe, M.V.; Aloni, S.; et al. Continuous-wave upconverting nanoparticle microlasers. Nat. Nanotechnol. 2018, 13, 572–577. [Google Scholar] [CrossRef] [Green Version]
  12. Liu, Q.; Sun, Y.; Yang, T.; Feng, W.; Li, C.; Li, F. Sub-10 nm hexagonal lanthanide-doped NaLuF4 upconversion nanocrystals for sensitive bioimaging in vivo. J. Am. Chem. Soc. 2011, 133, 17122–17125. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, W.J.; Zhang, J.P.; Wang, Z.; Wang, W.C.; Zhang, Q.Y. Spectroscopic and structural characterization of transparent fluorogermanate glass ceramics with LaF3:Tm3+ nanocrystals for optical amplifications. J. Alloys Compd. 2015, 634, 122–129. [Google Scholar] [CrossRef]
  14. Chen, G.; Lei, R.; Huang, F.; Wang, H.; Zhao, S.; Xu, S. Effects of Tm3+ concentration on upconversion luminescence and temperature-sensing behavior in Tm3+/Yb3+:Y2O3 nanocrystals. J. Biolumin. Chemilumin. 2018, 33, 1262–1267. [Google Scholar]
  15. Chen, S.; Song, W.; Cao, J.; Hu, F.; Guo, H. Highly sensitive optical thermometer based on FIR technique of transparent NaY2F7:Tm3+/Yb3+ glass ceramic. J. Alloys Compd. 2020, 825, 154011. [Google Scholar] [CrossRef]
  16. Li, X.; Yang, C.; Yu, Y.; Li, Z.; Lin, J.; Guan, X.; Zheng, Z.; Chen, D. Dual-modal photon upconverting and downshifting emissions from ultra-stable CsPbBr3 perovskite nanocrystals triggered by co-growth of Tm:NaYbF4 nanocrystals in glass. ACS Appl Mater. Interfaces 2020, 12, 18705–18714. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, Y.; Lu, Y.; Yang, X.; Zheng, X.; Wen, S.; Wang, F.; Vidal, X.; Zhao, J.; Liu, D.; Zhou, Z.; et al. Amplified stimulated emission in upconversion nanoparticles for super-resolution nanoscopy. Nature 2017, 543, 229–233. [Google Scholar] [CrossRef]
  18. Xiang, S.; Zheng, H.; Zhang, Y.; Peng, T.; Zhang, X.; Chen, B. Effect of Tm3+ concentration and temperature on blue and NIR upconversion luminescence in Tm3+/Yb3+ co-doped NaY(WO4)2 microstructures. J. Nanosci. Nanotechnol. 2016, 16, 636–642. [Google Scholar] [CrossRef]
  19. Xu, W.; Gao, X.; Zheng, L.; Zhang, Z.; Cao, W. An optical temperature sensor based on the upconversion luminescence from Tm3+/Yb3+ codoped oxyfluoride glass ceramic. Sens. Actuators B 2012, 173, 250–253. [Google Scholar] [CrossRef]
  20. Zhou, J.; Chen, G.; Zhu, Y.; Huo, L.; Mao, W.; Zou, D.; Sun, X.; Wu, E.; Zeng, H.; Zhang, J.; et al. Intense multiphoton upconversion of Yb3+-Tm3+ doped β-NaYF4 individual nanocrystals by saturation excitation. J. Mater. Chem. C 2015, 3, 364–369. [Google Scholar] [CrossRef]
  21. Zhang, C.; Zhang, J.; Lin, C.; Dai, S.; Chen, F. Improvement of third-order nonlinear properties in GeS2-Sb2S3-CsCl chalcogenide glass ceramics embedded with CsCl nano-crystals. Ceram. Int. 2020, 46, 27990–27995. [Google Scholar] [CrossRef]
  22. Velázquez, J.; Gorni, G.; Balda, R.; Fernández, J.; Pascual, L.; Durán, A.; Pascual, M. Non-linear optical properties of Er3+-Yb3+-doped NaGdF4 nanostructured glass-ceramics. Nanomaterials 2020, 10, 1425. [Google Scholar] [CrossRef] [PubMed]
  23. Cai, M.; Wei, T.; Zhou, B.; Tian, Y.; Zhou, J.; Xu, S.; Zhang, J. Analysis of energy transfer process based emission spectra of erbium doped germanate glasses for mid-infrared laser materials. J. Alloys Compd. 2015, 626, 165–172. [Google Scholar] [CrossRef]
  24. Chen, Y.; Liu, J.; Zhao, N.; Zhang, W.; Zhu, M.; Zhou, G.; Hou, Z. Manufacture and spectroscopic analysis of Tm3+-doped silica glass and microstructure optical fiber through the laser sintering technique. Appl. Phys. Express 2019, 12, 122012. [Google Scholar] [CrossRef]
  25. Mi, C.; Zhou, J.; Wang, F.; Lin, G.; Jin, D. Ultrasensitive ratiometric nanothermometer with large dynamic range and photostability. Chem. Mater. 2019, 31, 9480–9487. [Google Scholar] [CrossRef]
  26. Zhan, Q.Q.; Qian, J.; Liang, H.J.; Somesfalean, G.; Wang, D.; He, S.L.; Zhang, Z.G.; Andersson-Engels, S. Using 915 nm laser excited Tm3+/Er3+/Ho3+-Doped NaYbF4 upconversion nanoparticles for in vitro and deeper in vivo bioimaging without overheating irradiation. ACS Nano 2011, 5, 3744–3757. [Google Scholar] [CrossRef] [PubMed]
  27. Li, X.; Chen, D.; Huang, F.; Chang, G.; Zhao, J.; Qiao, X.; Xu, X.; Du, J.; Yin, M. Phase-Selective nanocrystallization of NaLnF4 in aluminosilicate glass for random laser and 940 nm LED-excitable upconverted luminescence. Laser Photonics Rev. 2018, 12, 1800030. [Google Scholar] [CrossRef]
  28. Gorni, G.; Velázquez, J.; Mosa, J.; Mather, G.; Serrano, A.; Vila, M.; Castro, G.; Bravo, D.; Balda, R.; Fernández, J.; et al. Transparent sol-gel oxyfluoride glass-ceramics with high crystalline fraction and study of re incorporation. Nanomaterials 2019, 9, 530. [Google Scholar] [CrossRef] [Green Version]
  29. Yanes, A.C.; Santana-Alonso, A.; Méndez-Ramos, J.; del-Castillo, J.; Rodríguez, V.D. Novel sol-gel nano-glass-ceramics comprising Ln3+-doped YF3 nanocrystals: Structure and high efficient UV up-conversion. Adv. Funct. Mater. 2011, 21, 3136–3142. [Google Scholar] [CrossRef]
  30. Chen, D.; Peng, Y.; Li, X.; Zhong, J.; Huang, H.; Chen, J. Simultaneous tailoring of dual-phase fluoride precipitation and dopant distribution in glass to control upconverting luminescence. ACS Appl. Mater. Interfaces 2019, 11, 30053–30064. [Google Scholar] [CrossRef]
  31. Walas, M.; Lewandowski, T.; Synak, A.; Łapiński, M.; Sadowski, W.; Kościelska, B. Eu3+ doped tellurite glass ceramics containing SrF2 nanocrystals: Preparation, structure and luminescence properties. J. Alloys Compd. 2017, 696, 619–626. [Google Scholar] [CrossRef]
  32. Chen, D.; Peng, Y.; Li, X.; Zhong, J.; Huang, P. Competitive nanocrystallization of Na3ScF6 and NaYbF4 in aluminosilicate glass and optical spectroscopy of Ln3+ dopants. Ceram. Int. 2018, 44, 15666–15673. [Google Scholar] [CrossRef]
  33. Fang, Z.; Chen, Z.; Peng, W.; Shao, C.; Zheng, S.; Hu, L.; Qiu, J.; Guan, B.O. Phase-separation engineering of glass for drastic enhancement of upconversion luminescence. Adv. Opt. Mater. 2019, 7, 1801572. [Google Scholar] [CrossRef]
  34. Zhou, J.; Chen, G.; Wu, E.; Bi, G.; Wu, B.; Teng, Y.; Zhou, S.; Qiu, J. Ultrasensitive polarized up-conversion of Tm3+-Yb3+ doped beta-NaYF4 single nanorod. Nano Lett. 2013, 13, 2241–2246. [Google Scholar] [CrossRef] [PubMed]
  35. Zhou, J.; Wen, S.; Liao, J.; Clarke, C.; Tawfik, S.A.; Ren, W.; Mi, C.; Wang, F.; Jin, D. Activation of the surface dark-layer to enhance upconversion in a thermal field. Nat. Photonics 2018, 12, 154–158. [Google Scholar] [CrossRef] [Green Version]
  36. Gao, G.; Busko, D.; Joseph, R.; Howard, I.A.; Turshatov, A.; Richards, B.S. Highly efficient La2O3:Yb3+,Tm3+ single-band NIR-to-NIR upconverting microcrystals for anti-counterfeiting applications. ACS Appl. Mater. Interfaces 2018, 10, 39851–39859. [Google Scholar] [CrossRef]
  37. Su, J.-Y.; Zhang, X.-Y.; Li, X.; Qu, M.-L. Synthesis and luminescence properties of Yb3+, Tm3+ and Ho3+ co-doped SrGd2(WO4)2(MoO4)2 nano-crystal. Adv. Powder Technol. 2020, 31, 1051–1059. [Google Scholar] [CrossRef]
  38. Arppe, R.; Hyppanen, I.; Perala, N.; Peltomaa, R.; Kaiser, M.; Wurth, C.; Christ, S.; Resch-Genger, U.; Schaferling, M.; Soukka, T. Quenching of the upconversion luminescence of NaYF4:Yb3+, Er3+ and NaYF4:Yb3+, Tm3+ nanophosphors by water: The role of the sensitizer Yb3+ in non-radiative relaxation. Nanoscale 2015, 7, 11746–11757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Fu, Y.; Zhao, L.; Guo, Y.; Yu, H. Highly sensitive optical thermometers based on unconventional thermometric coupled levels of Tm3+ following a Boltzmann-type distribution in oxyfluoride glass ceramics. New J. Chem. 2019, 43, 16664–16669. [Google Scholar] [CrossRef]
  40. Zhao, J.; Jin, D.; Schartner, E.P.; Lu, Y.; Liu, Y.; Zvyagin, A.V.; Zhang, L.; Dawes, J.M.; Xi, P.; Piper, J.A.; et al. Single-nanocrystal sensitivity achieved by enhanced upconversion luminescence. Nat. Nanotechnol. 2013, 8, 729–734. [Google Scholar] [CrossRef]
  41. Fu, Y.; Zhao, L.; Guo, Y.; Wu, B.; Dong, H.; Yu, H. Ultrapure NIR-to-NIR single band emission of β-PbF2: Yb3+/Tm3+ in glass ceramics. J. Lumin. 2019, 208, 33–38. [Google Scholar] [CrossRef]
  42. Fu, Y.; Zhao, L.; Guo, Y.; Yu, H. Up-conversion luminescence lifetime thermometry based on the 1G4 state of Tm3+ modulated by cross relaxation processes. Dalton Trans. 2019, 48, 16034–16040. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD patterns of 1.0Yb3+-0.005Tm3+ co-doped PG and GCs. (b) TEM and (c) HRTEM images of 1.5Yb3+-0.1Tm3+ co-doped GC heat treated at 540 °C for 10 h.
Figure 1. (a) XRD patterns of 1.0Yb3+-0.005Tm3+ co-doped PG and GCs. (b) TEM and (c) HRTEM images of 1.5Yb3+-0.1Tm3+ co-doped GC heat treated at 540 °C for 10 h.
Nanomaterials 11 01033 g001
Figure 2. (a) Transmission spectra of 1.0Yb3+-0.1Tm co-doped PG and GC. (b) UC emission spectra 1.0Yb3+-0.005Tm3+co-doped PG and GCs. (c) Emission decay curves of 1.0Yb3+-0.005Tm3+ co-doped PG and GCs monitored at 450 nm emission. (d) Emission decay curves of 1.0Yb3+-0.005Tm3+ co-doped PG and GCs monitored at 802 nm emission.
Figure 2. (a) Transmission spectra of 1.0Yb3+-0.1Tm co-doped PG and GC. (b) UC emission spectra 1.0Yb3+-0.005Tm3+co-doped PG and GCs. (c) Emission decay curves of 1.0Yb3+-0.005Tm3+ co-doped PG and GCs monitored at 450 nm emission. (d) Emission decay curves of 1.0Yb3+-0.005Tm3+ co-doped PG and GCs monitored at 802 nm emission.
Nanomaterials 11 01033 g002
Figure 3. (a) Ln-Ln plots of UC emission intensity versus excited power density of 980 nm laser. (b) Schematic diagrams of energy levels and relative transitions in KYb3F10:Tm3+ GCs.
Figure 3. (a) Ln-Ln plots of UC emission intensity versus excited power density of 980 nm laser. (b) Schematic diagrams of energy levels and relative transitions in KYb3F10:Tm3+ GCs.
Nanomaterials 11 01033 g003
Figure 4. UC emission spectra of 1.0Yb3+-yTm3+ co-doped (a) PG and (b) GC excited by a 980 nm LD. (y = 0.005~0.3).
Figure 4. UC emission spectra of 1.0Yb3+-yTm3+ co-doped (a) PG and (b) GC excited by a 980 nm LD. (y = 0.005~0.3).
Nanomaterials 11 01033 g004
Figure 5. XRD patterns of xYb3+-0.1Tm3+co-doped and no-doped GCs heat treated at 540 °C for 10 h. (x = 0~3.5%).
Figure 5. XRD patterns of xYb3+-0.1Tm3+co-doped and no-doped GCs heat treated at 540 °C for 10 h. (x = 0~3.5%).
Nanomaterials 11 01033 g005
Figure 6. (a) UC emission spectra xYb3+-0.1Tm3+co-doped GCs heat treated at 540 °C for 5 h. (x = 0.0~3.5). (b) Emission decay curves of xYb3+-0.1Tm3+ co-doped PG and GCs heat treated at 540 °C for 5 h monitored at 802 nm emission. (x = 0.5~3.5).
Figure 6. (a) UC emission spectra xYb3+-0.1Tm3+co-doped GCs heat treated at 540 °C for 5 h. (x = 0.0~3.5). (b) Emission decay curves of xYb3+-0.1Tm3+ co-doped PG and GCs heat treated at 540 °C for 5 h monitored at 802 nm emission. (x = 0.5~3.5).
Nanomaterials 11 01033 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, J.; Long, Y.; Zhao, Q.; Zheng, S.; Fang, Z.; Guan, B.-O. Nano-Crystallization of Ln-Fluoride Crystals in Glass-Ceramics via Inducing of Yb3+ for Efficient Near-Infrared Upconversion Luminescence of Tm3+. Nanomaterials 2021, 11, 1033. https://doi.org/10.3390/nano11041033

AMA Style

Li J, Long Y, Zhao Q, Zheng S, Fang Z, Guan B-O. Nano-Crystallization of Ln-Fluoride Crystals in Glass-Ceramics via Inducing of Yb3+ for Efficient Near-Infrared Upconversion Luminescence of Tm3+. Nanomaterials. 2021; 11(4):1033. https://doi.org/10.3390/nano11041033

Chicago/Turabian Style

Li, Jianfeng, Yi Long, Qichao Zhao, Shupei Zheng, Zaijin Fang, and Bai-Ou Guan. 2021. "Nano-Crystallization of Ln-Fluoride Crystals in Glass-Ceramics via Inducing of Yb3+ for Efficient Near-Infrared Upconversion Luminescence of Tm3+" Nanomaterials 11, no. 4: 1033. https://doi.org/10.3390/nano11041033

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop