Next Article in Journal
Continuous Patterning of Silver Nanowire-Polyvinylpyrrolidone Composite Transparent Conductive Film by a Roll-to-Roll Selective Calendering Process
Next Article in Special Issue
Solvent Influence on the Magnetization and Phase of Fe-Ni Alloy Nanoparticles Generated by Laser Ablation in Liquids
Previous Article in Journal
Remarkable Single Atom Catalyst of Transition Metal (Fe, Co & Ni) Doped on C2N Surface for Hydrogen Dissociation Reaction
Previous Article in Special Issue
Extraction of Metal Ions by Interfacially Active Janus Nanoparticles Supported by Wax Colloidosomes Obtained from Pickering Emulsions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrothermal Synthesis and Properties of Yb3+/Tm3+ Doped Sr2LaF7 Upconversion Nanoparticles

1
Centre of Excellence for Photoconversion, Vinča Insitute of Nuclear Sciences—National Institute of the Republic of Serbia, University of Belgrade, P.O. Box 522, 11001 Belgrade, Serbia
2
Institute of Solid State Physics, University of Latvia, Kengaraga Street 8, LV-1063 Riga, Latvia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(1), 30; https://doi.org/10.3390/nano13010030
Submission received: 1 December 2022 / Revised: 14 December 2022 / Accepted: 19 December 2022 / Published: 21 December 2022
(This article belongs to the Special Issue Morphological Design and Synthesis of Nanoparticles)

Abstract

:
We report the procedure for hydrothermal synthesis of ultrasmall Yb3+/Tm3+ co-doped Sr2LaF7 (SLF) upconversion phosphors. These phosphors were synthesized by varying the concentrations of Yb3+ (x = 10, 15, 20, and 25 mol%) and Tm3+ (y = 0.75, 1, 2, and 3 mol%) with the aim to analyze their emissions in the near IR spectral range. According to the detailed structural analysis, Yb3+ and Tm3+ occupy the La3+ sites in the SLF host. The addition of Yb3+/Tm3+ ions has a huge impact on the lattice constant, particle size, and PL emission properties of the synthesized SLF nanophosphor. The results show that the optimal dopant concentrations for upconversion luminescence of Yb3+/Tm3+ co-doped SLF are 20 mol% Yb3+ and 1 mol% Tm3+ with EDTA as the chelating agent. Under 980 nm light excitation, a strong upconversion emission of Tm3+ ions around 800 nm was achieved. In addition, the experimental photoluminescence lifetime of Tm3+ emission in the SLF host is reported. This study discovered that efficient near IR emission from ultrasmall Yb3+/Tm3+ co-doped SLF phosphors may have potential applications in the fields of fluorescent labels in bioimaging and security applications.

Graphical Abstract

1. Introduction

Upconversion (UC) phosphor materials doped or co-doped with trivalent lanthanide (Ln3+) ions have drawn considerable attention, since four-electron configurations of Ln3+ ions should split by electron–electron repulsion and spin–orbit coupling, resulting in a rich energy-level pattern that can be easily populated by the near infrared (NIR) laser source [1,2]. Under a NIR laser source, UC phosphors effectively convert energy into shorter wavelength emissions (NIR, visible, or ultraviolet) via multiphoton absorption or efficient energy transfer [3]. These unique features of Ln3+-doped/co-doped UC phosphors enable a wide range of applications, such as LED devices [4], solar energy conversion [5], temperature sensors [6,7,8], latent fingerprint detection [9,10], biomedical imaging [11,12], food safety detection [13], etc. The NIR-to-NIR UC luminescence mechanism is gaining popularity due to its efficient way of producing NIR emission at nearly 800 nm when excited by a commercially available laser source (~980 nm) [9,14,15,16,17,18,19,20,21]. It is well known that the Yb3+ ion, with its simple energy level structure (ground state 2F7/2 and excited state 2F5/2), strong absorption band in the wavelength range of 860–1060 nm, and relatively long luminescence lifetime (1–2 ms), is an excellent sensitizer for energy transfer to RE ions [21,22,23]. Ln3+ ions such as Pr3+, Er3+, Tm3+, and Ho3+ have been recognized as co-activators in Yb3+/Ln3+ UC phosphors due to the position of their energy levels and the possibility of efficient radiative transitions under NIR light sources [9,14,15,16,17,18,19,20,21,22,23,24]. Among them, Yb3+/Tm3+ UC phosphors have been intensively studied because NIR-to-NIR UC emission is crucial for biomedical imaging [17,18,19,20], security printing applications [15,16] and latent fingerprint detection [9].
The UC luminescence mechanism has been explored in a variety of host materials, including chlorides, fluorides, oxides, vanadates, and others. The selection of appropriate host materials with low phonon energy frequencies to prevent non-radiative relaxation processes and thus improve emission efficiency is essential for UC luminescence. Chlorides have low phonon frequencies (≤300 cm−1) and poor chemical stability, which limits their application possibilities, whereas oxide host materials have relatively high phonon frequencies (>500 cm−1) and excellent chemical stability [25]. Fluoride materials are thus ideal hosts for UC luminescence due to their low phonon frequency (from 300 to 500 cm−1), good chemical stability, and simplicity of dispersion in colloidal form with water or various nonpolar solvents [25,26].
Even though alkaline-Ln3+ tetrafluoride phosphors (ALnF4, A = Na, K, Li) are the most preferred for efficient UC luminescence, their agglomeration limits applications that require nanoparticles, such as in biomedical imaging [27,28,29,30]. According to several recent studies, alkali–earth–Ln3+ nanophosphors (M2LnF7, M = Ca, Sr, Ba; Ln3+ = Y, La, Gd, Lu) are small enough for biomedical imaging applications while exhibiting extremely high UC luminescence [31,32,33,34]. Under 980 nm laser stimulation, Xie et al. observed efficient visible emission of Sr2LaF7:Yb3+, Er3+ UC nanophosphors with average particles of around 25 nm [34]. Guo et al. reported that Sr2GdF7:Er3+, Yb3+ nanocrystals incorporated into electrospun fibers promote energy transfer processes from Yb3+ to Er3+, which is crucial for potential applications in the field of noncontact biomedical temperature sensors [8]. In this work, a set of Sr2La1-x-yF7 phosphors with different concentrations of Yb3+ (x = 10, 15, 20, and 25 mol%) and Tm3+ (y = 0.75, 1, 2, and 3 mol%) ions with respect to La3+ ions are prepared hydrothermally at 180 °C. The room temperature photoluminescence spectra of SLF:Yb, Tm under 980 nm excitation clearly show intense NIR emissions in the wavelength range from 750 to 850 nm, with the highest intensity around 800 nm.
Herein, we propose a procedure for the hydrothermal synthesis of small Yb/Tm activated SFL nanoparticles. Further, we documented their NIR-to-NIR UC. This UC process has been given much less attention in Yb3+ and Tm3+ co-doped phosphors than blue and deep-red UC emissions, although it can considerably expand the fields of application of UC nanophosphors, especially as a suitable fluorescent marker in the development of latent fingerprints.

2. Materials and Methods

2.1. Chemicals

Strontium nitrate (Sr(NO3)2, Alfa Aesar Karlsruhe, Germany, 99%), lanthanum (III) nitrate hexahydrate (La(NO3)3⋅6H2O, Alfa Aesar, 99.99%), ytterbium (III) nitrate hexahydrate (Yb(NO3)3⋅5H2O, Alfa Aesar, 99.9%), thulium (III) nitrate hexahydrate (Tm(NO3)3⋅6H2O, Alfa Aesar, 99.9%), disodium ethylendiaminetetraacetate dihydrate (EDTA-2Na, C10H14N2O8Na2∙2H2O, Kemika, Zagreb, Croatia, 99%), ammonium fluoride (NH4F, Alfa Aesar, 98%), 25% ammonium solution (NH4OH, Fisher, Loughborough, Leicestershire, United Kingdom) and de-ionized water were used as starting materials without further purification.

2.2. Synthesis of SLF:Yb,Tm

Sr2La1-x-yF7:xYb,yTm were synthesized hydrothermally using Sr(NO3)2, Ln3+ nitrates (Ln = La,Yb,Tm), and NH4F as precursors and EDTA-2Na as a stabilizing agent (see Figure 1). Typically, for the synthesis of 1 g of the representative sample Sr2LaF7 co-doped with 20mol% Yb3+ and 1 mol% Tm3+, all nitrates were weighed according to the stoichiometric ratio (precisely, 0.4762g Sr3+-nitrate, 0.3849 g La3+-nitrate, 0.1010 g Yb3+-nitrate and 0.0050 g Tm3+-nitrate) and then dissolved in 12.5 mL deionized water while stirring at room temperature. The above solution was then mixed for 30 min with a transparent solution of 0.4188 g EDTA-2Na in 12.5 mL in water (molar ratio EDTA-2Na:La = 1:1). Following that, a 10 mL aqueous solution containing 0.5001 g of NH4F (molar ratio NH4F:La = 12:1) was added and vigorously stirred for 1 h, yielding a white complex. Using 400 µL of NH4OH, the pH of the mixture was adjusted to around 6. This mixture was placed in a 100-mL Teflon-lined autoclave and heated in the oven at 180 °C for 20 h. After natural cooling, the final products were centrifuged and washed twice with water, then once with an ethanol:water = 1:1 mixture to remove any possible remnants before drying in an air atmosphere at 80 °C for 4 h. Undoped SLF and SLF phosphors with varying concentrations of Yb3+ (x = 10, 15, 20, and 25 mol%) and Tm3+ (y = 0.75, 1, 2, and 3 mol%) ions with respect to La3+ ions were prepared using the described procedure.

2.3. Measurement

X-ray diffraction (XRD) measurements were performed on a Rigaku SmartLab system operating with Cu Kα radiation (30 mA, 40 kV) in the 2θ range from 10° to 90°. Diffraction data were recorded with a step size of 0.02° and a counting time of 1°/min over the investigated 2θ. Results of the structural analysis (unit cell parameters, crystal coherence size, microstrain values, and data fit parameters) were obtained using the built-in PDXL2 software. The microstructure of the samples was characterized by a transmission electron microscope (TEM) Tecnai GF20 operated at 200 kV. The average particle size was calculated using ImageJ software. Diffuse reflectance measurements were performed with the Shimadzu UV-2600 (Shimadzu Corporation, Tokyo, Japan) spectrophotometer equipped with an integrated sphere (ISR-2600), using BaSO4 as the standard reference. Luminescence characterization was done using a 980 nm high power (3W) solid state IR laser as an excitation source. Luminescence emissions were recorded using a FHR1000 monochromator (Horiba Jobin Yvon) and an ICCD camera (Horiba Jobin Yvon 3771). All the measurements were performed at room temperature.

3. Results and Discussion

3.1. XRD Analysis

The MxLnF2x+3 fluorides crystallize in a cubic structure with F m 3 ¯ m space group [31]. XRD patterns of SLF:xYb3+,1 mol%Tm3+ and SLF:20 mol%Yb3+,yTm3+ nanophosphors are shown in Figure 2a,b, respectively. Despite the addition of Yb3+ and Tm3+ ions, the main diffraction peaks observed around 2θ = 26.4, 30.7, 43.8, 51.9, 54.4, 63.7, 70.2, and 80.5°, correspond to the main reflections from 111, 200, 220, 311, 222, 400, 331, and 422 crystal planes, respectively, and are well-aligned with the standard data of ICDD No. 00–053–0774 for pure SLF. Diffraction peaks corresponding to other phases and/or impurities were not noticed. The sharp diffraction peaks indicate a good crystallinity of SLF nanophosphors. Table 1 and Table 2 show the results of the structural analysis using whole pattern-fitting (WPF) refinement: crystallite coherence size (CS), microstrain values, unit cell parameters, unit cell volume (CV), and data fit parameters (Rwp, Rp, Re and GOF) of SLF:xYb3+,1 mol%Tm3+ and SLF:20 mol%Yb3+,yTm3+ nanophosphors. The CS of pure SLF is estimated to be 27.1 nm, and the lattice constant a is 5.8451 Å (CV = 199.70 Å3). The influence of Yb3+ doping in SLF lattice causes the linear host lattice shrinkage up to a = 5.8045 Å, CV = 195.57 Å3 for the sample SLF:25 mol%Yb3+,1 mol%Tm3+. This shrinkage could be ascribed to the fact that dopants with smaller ionic radii Yb3+ (0.868 Å) and Tm3+ (0.880 Å) replace the La3+ with larger ionic radii (1.032 Å) in SLF [35]. Tm3+ doping in SLF lattice also produces host lattice shrinkage up to a = 5.8107 Å, CV = 196.2 Å3 for the sample SLF:20 mol%Yb3+, 2 mol%Tm3+. When the concentration of Tm3+ ions is increased further, the other doping strategy occurs due to the Tm3+ ability to occupy the interstitial sites, leading to crystal lattice expansion (a = 5.8288 Å, CV = 198.03 Å3) [36]. To identify the strategy of Tm3+ and Yb3+ doping in SLF, the magnified (111) diffraction peak of the samples are shown in Figure 2c,d. With the addition of Yb3+ ions, the position of the (111) diffraction peak consistently shifts to higher degree values, and the shift becomes more notable with increasing dopant concentration (Figure 2c). Tm3+ doping in SLF has the same behavior at concentrations equal to or less than 2 mol%, while further addition of Tm3+ shifts the diffraction peak to lower degree values (Figure 2d). These findings confirm that both Yb3+ and Tm3+ (equal or less than 2 mol%) were successfully inserted into SLF by occupying La3+ sites.

3.2. Morphology Analysis

An assisted EDTA hydrothermal method was used to create SLF:Yb, Tm nanophosphors. EDTA is an efficient complexing agent for Ln3+ ions, with chelation constants (logK1) of 19.51 and 15.50 for Yb3+ and La3+ ions, respectively [31]. Due to the ability to improve crystalline seed dispersibility by forming [Sr-EDTA]2+ and [La-EDTA]+ complexes after mixing all of the chemicals, EDTA prevented SLF particle aggregation during the subsequent hydrothermal treatment. On the other hand, [La-EDTA]+ cations could be absorbed on the surfaces of SLF particles, limiting their further growth into large particles, and also increasing their stability [36,37].
TEM images of undoped SLF with different magnifications together with particle size distribution histogram are shown in Figure 3I. Nanoparticles show a similar, quasi-spherical shape as well as a high degree of crystallinity. HRTEM image of SLF phosphors (Figure 3Ie) shows that the measured d-spacing is around 3.4 Å, that corresponds to the (111) lattice plane of SLF, which agrees to the previous XRD data. The half-displayed particles were not considered when calculating the average particle size, and the histogram was fitted with a log-normal distribution. The average crystalline size of nanoparticles, considering around 120 particles, was estimated to be 38 ± 4 nm (see Figure 3If). The influence of Yb3+ and Tm3+ co-doping on the morphology of SLF samples can be observed by comparing features in Figure 3I, 3II and 3III, respectively. The average particle size of SLF nanophosphor doped with 10 mol% Yb3+ (Figure 3IIa–f) and SLF doped with 25 mol% Yb3+ (Figure 3IIIa–f) ions and a fixed concentration level of 1 mol% Tm3+ was calculated to be 25 ± 3 nm and 26 ± 2 nm, respectively. Therefore, the average particle size of SLF was reduced by doping from 38 to around 25 nm, which is well-aligned with the previous XRD analysis. HRTEM images of both SLF:10Yb,1Tm (Figure 3IIe) and SLF:25Yb,1Tm phosphors (Figure 3IIIe) show that the measured d-spacing is around 3.5 Å, which also corresponds to the (111) lattice plane of SLF. As previously explained, the addition of Yb3+/Tm3+ ions has a slight impact on the lattice constant when compared to undoped SLF because dopants with smaller ionic radii Yb3+ and Tm3+ replace the La3+ with larger ionic radii. The average particle size of SLF nanophosphor, on the other hand, is strongly influenced by the concentrations of Yb3+ and Tm3+.

3.3. Spectroscopic Properties

Figure 4a shows the room temperature diffuse reflectance spectra of a representative SLF:20Yb,1Tm sample in the 400–1300 nm wavelength range with typical optical features of Yb3+ and Tm3+ ions [38]. The absorption peaks of Yb3+ ions appear in the 885–1060 nm wavelength range due to electronic transitions from 2F7/2   2F5/2, with the highest intensity around 980 nm. In the case of Tm3+ ions, three major electronic transitions are involved: 3H63F2,3, 3H63H4, and 3H63H5, which correspond to absorption peaks at 677 nm, 770 nm, and 1206 nm, respectively.
The room temperature emission spectra of a representative SLF:20Yb,1Tm nanophosphor in the 450–900 nm wavelength range are shown in Figure 4b. In a typical multiphoton UC process, Yb3+ absorbs NIR radiation at around 980 nm which causes electron excitation from 2F7/2 to 2F5/2 energy level. Then, the Tm3+ is excited via cross-relaxation and energy transfer from excited Yb3+. The deexcitation from multiple Tm3+ excited levels provide emissions that cover UV–VIS–NIR spectra. The observed emission peaks, which occur at wavelengths ranging from 455 to 500 nm, 625 to 720 nm, and 750 to 850 nm, are attributed to the transitions from the 1G43H6, 1G43F4, and 1G43H5 / 3H43H6 of excited Tm3+ ions, respectively. The PL decay curve at room temperature is shown in the inset of Figure 4b. The average emission time (τav), calculated based on the double exponential model, was used as a measurement of PL lifetime (τ). Through the fit of our experimental data to the double exponential model, the two values of τ1 and τ2 are obtained:
I ( t ) = A 1 e t τ 1 + A 2 e t τ 2 + b g .
where, A1 and A2 are arbitrary constants (magnitudes of short and long decay components), and bg is a background correction. Because the measured signal (𝐼 (𝑡)) at delayed time 𝑡𝑑 is proportional to the number of excited states at the moment 𝑡𝑑, the simple weighted average formula is used to calculate 𝜏av:
τ a v = A 1 τ 1 + A 2 τ 2 A 1 + A 2 .
The PL lifetime of Tm3+ (3H43H6 transition) in a representative SLF:20Yb,1Tm nanophosphor was estimated to be 1.05 ms. Table 3 summarizes the arbitrary constants, background correction, two values of PL lifetime (τ1 and τ2), and average PL lifetime of the representative SLF:20Yb,1Tm nanophosphor. The deviation of the 3H43H6 emission decay from the single exponentiality indicates that energy back-transfer to Yb3+ occurs from 3H4 level. This can be further investigated by measuring the variation in lifetimes of Yb3+ 2F5/2 emission for different Tm3+ and Yb3+ concentrations.
The UC emission intensity relates to both Yb3+ and Tm3+ concentrations. Figure 4c presents the dependence of the integrated UC emission intensity of SLF with different concentrations of Yb3+ (x = 10, 15, 20, and 25 mol%) and a fixed Tm3+ concentration (1 mol%). With increasing Yb3+ concentration, the NIR emission intensity band increases, reaching a maximum value at 20 mol% of Yb3+. Similarly, when Yb3+ concentration is fixed at 20 mol%, the NIR emission of SLF monitored at different concentrations of Tm3+ (x = 0.75, 1, 2, and 3 mol%) has the highest intensity for 1 mol% Tm3+, as shown in Figure 4d. When the Tm3+ doping concentration is equal to or greater than 2, the emission intensity decreases gradually due to the concentration quenching. In contrast to the emission intensity, the shape and characteristic peak position of the UC emission spectra have not changed. For the Yb3+ and Tm3+ co-doped phosphors, blue (1G43H6) and deep-red (1G43F4) emissions have been widely investigated [27,38,39,40], while the efficient NIR-to-NIR Yb3+/Tm3+ UC emission in ultrasmall SLF nanophosphor has received far less attention. Therefore, ultrasmall SLF nanoparticles with intense emission around 800 nm are promising candidates as fluorescent labels in bioimaging and security applications.

4. Conclusions

In conclusion, ultrasmall SLF:Yb3+/Tm3+ nanoparticles were produced using a straightforward hydrothermal process at a variety of doping doses. With Yb3+ and Tm3+ ions present, the lattice constant and average particle size of SLF are both decreased from 38 nm to roughly 25 nm. At room temperature, Yb3+ and Tm3+ concentrations have a significant impact on the PL emission properties. When excited with a 980 nm high power (3W) solid state IR laser, these ultrasmall nanoparticles show simultaneous three-color (blue-green, deep-red, and NIR) UC emissions in the 450–900 nm wavelength range. The blue-green and deep-red emission bands are weak, while the NIR emission band is strong, which is beneficial for imaging biological tissues. Furthermore, the PL lifetime of Tm3+ (3H43H6 transition) in a representative SLF:20Yb,1Tm nanophosphor was estimated to be 1.05 ms. These findings also suggest that SLF:Yb,Tm could be a useful fluorescent marker in the development of latent fingerprints. Our future work will concentrate on the dual-mode fluorescent development of latent fingerprints using both NIR-to-VIS and NIR-to-NIR processes to achieve double fluorescent images in dark and bright fields, as well as additional contrast and sensitivity analysis of fingerprints or fingerprint residuals deposited on a variety of substrates. Furthermore, the proposed NIR-to-NIR UC mechanism of ultrasmall Yb3+/Tm3+ co-doped SLF nanophosphors could be a useful tool in security applications.

Author Contributions

Conceptualization, A.S. and M.D.D.; methodology, Z.R., K.S., A.S. and M.D.D.; formal analysis, B.M., J.P., Z.R., K.S., M.K., K.V., A.S. and M.D.D.; investigation, B.M., J.P., K.M., M.K., Ž.A. and K.V.; data curation, B.M. and J.P.; writing—original draft preparation, B.M. and Z.R.; writing—review and editing, A.S. and M.D.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science, Technological Development and Innovation of the Republic of Serbia and by the European Union’s Horizon 2020 Framework Programme H2020-WIDESPREAD-01-2016-2017-TeamingPhase2 under grant agreement No. 739508, project CAMART2.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors from Serbia acknowledge funding from the Ministry of Science, Technological Development and Innovation of the Republic of Serbia. The authors from Latvia acknowledge funding from the European Union’s Horizon 2020 Framework Programme H2020-WIDESPREAD-01-2016-2017-TeamingPhase2 under grant agreement No. 739508, project CAMART2.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Peijzel, P.S.; Meijerink, A.; Wegh, R.T.; Reid, M.F.; Burdick, G.W. A complete 4fn energy level diagram for all trivalent lanthanide ions. J. Solid State Chem. 2005, 178, 448–453. [Google Scholar] [CrossRef]
  2. Zhou, B.; Shi, B.; Jin, D.; Liu, X. Controlling upconversion nanocrystals for emerging applications. Nat. Nanotechnol. 2015, 10, 924–936. [Google Scholar] [CrossRef] [PubMed]
  3. Zou, W.; Visser, C.; Maduro, J.A.; Pshenichnikov, M.S.; Hummelen, J.C. Broadband dye-sensitized upconversion of near-infrared light. Nat. Photonics 2012, 6, 560–564. [Google Scholar] [CrossRef] [Green Version]
  4. Erol, E.; Vahedigharehchopogh, N.; Kıbrıslı, O.; Ersundu, M.Ç.; Ersundu, A.E. Recent progress in lanthanide-doped luminescent glasses for solid-state lighting applications-a review. J. Phys. Condens. Matter 2021, 33, 483001. [Google Scholar] [CrossRef]
  5. Chen, E.Y.; Milleville, C.; Zide, J.M.O.; Doty, M.F.; Zhang, J. Upconversion of low-energy photons in semiconductor nanostructures for solar energy harvesting. MRS Energy Sustain. 2018, 5, e16. [Google Scholar] [CrossRef] [Green Version]
  6. Skwierczyńska, M.; Stopikowska, N.; Kulpiński, P.; Kłonowska, M.; Lis, S.; Runowski, M. Ratiometric Upconversion Temperature Sensor Based on Cellulose Fibers Modified with Yttrium Fluoride Nanoparticles. Nanomaterials 2022, 12, 1926. [Google Scholar] [CrossRef]
  7. Ryszczyńska, S.; Trejgis, K.; Marciniak, Ł.; Grzyb, T. Upconverting SrF2:Er3+ Nanoparticles for Optical Temperature Sensors. ACS Appl. Nano Mater. 2021, 4, 10438–10448. [Google Scholar] [CrossRef]
  8. Guo, M.Y.; Shen, L.F.; Pun, E.Y.B.; Lin, H. Sr2GdF7: Er3+/Yb3+ nanocrystal-inlaid pliable fibers for synergistic feedback temperature monitoring. J. Lumin. 2022, 252, 119394. [Google Scholar] [CrossRef]
  9. Baride, A.; Sigdel, G.; Cross, W.M.; Kellar, J.J.; May, P.S. Near Infrared-to-Near Infrared Upconversion Nanocrystals for Latent Fingerprint Development. ACS Appl. Nano Mater. 2019, 2, 4518–4527. [Google Scholar] [CrossRef]
  10. Wang, M.; Li, M.; Yang, M.; Zhang, X.; Yu, A.; Zhu, Y.; Qiu, P.; Mao, C. NIR-induced highly sensitive detection of latent fingermarks by NaYF4:Yb,Er upconversion nanoparticles in a dry powder state. Nano Res. 2015, 8, 1800–1810. [Google Scholar] [CrossRef]
  11. Kavand, A.; Serra, C.A.; Blanck, C.; Lenertz, M.; Anton, N.; Vandamme, T.F.; Yves Mély, Y.; Przybilla, F.; Chan-Seng, D. Controlled Synthesis of NaYF4:Yb,Er Upconversion Nanocrystals as Potential Probe for Bioimaging: A Focus on Heat Treatment. ACS Appl. Nano Mater. 2021, 4, 5319–5532. [Google Scholar] [CrossRef]
  12. Chen, G.; Qiu, H.; Prasad, P.N.; Chen, X. Upconversion Nanoparticles: Design, Nanochemistry, and Applications in Theranostics. Chem. Rev. 2014, 114, 5161–5214. [Google Scholar] [CrossRef] [PubMed]
  13. Ji, G.; Wang, Y.; Qin, Y.; Peng, Y.; Li, S.; Han, D.; Ren, S.; Qin, K.; Li, S.; Gao, Z.; et al. Latest developments in the upconversion nanotechnology for rapid detection of food safety: A review. Nanotechnol. Rev. 2022, 11, 2110–2122. [Google Scholar] [CrossRef]
  14. May, P.S.; Baride, A.; Hossan, M.Y.; Berry, M. Measuring the Internal Quantum Yield of Upconversion Luminescence for Ytterbium-Sensitized Upconversion Phosphors Using the Ytterbium-(III) Emission as an Internal Standard. Nanoscale 2018, 10, 17212–17226. [Google Scholar] [CrossRef] [PubMed]
  15. 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]
  16. Baride, A.; Meruga, J.M.; Douma, C.; Langerman, D.; Crawford, G.; Kellar, J.J.; Cross, W.M.; May, P.S. A NIR-to-NIR Upconversion Luminescence System for Security Printing Applications. RSC Adv. 2015, 5, 101338–101346. [Google Scholar] [CrossRef]
  17. Kim, J.; Kwon, J.H.; Jang, J.; Lee, H.; Kim, S.; Hahn, Y.K.; Kim, S.K.; Lee, K.H.; Lee, S.; Pyo, H.; et al. Rapid and Background-free Detection of Avian Influenza Virus in Opaque Sample using NIR-to-NIR Upconversion Nanoparticle-based Lateral Flow Immunoassay Platform. Biosens. Bioelectron. 2018, 112, 209–215. [Google Scholar] [CrossRef]
  18. Levy, E.S.; Tajon, C.A.; Bischof, T.S.; Iafrati, J.; Fernandez-Bravo, A.; Garfield, D.J.; Chamanzar, M.; Maharbiz, M.M.; Sohal, V.S.; Schuck, P.J. Energy-Looping Nanoparticles: Harnessing Excited-State Absorption for Deep-Tissue Imaging. ACS Nano 2016, 10, 8423–8433. [Google Scholar] [CrossRef]
  19. Xia, A.; Chen, M.; Gao, Y.; Wu, D.; Feng, W.; Li, F. Gd3+ complex-modified NaLuF4-based upconversion nanophosphors for trimodality imaging of NIR-to-NIR upconversion luminescence, X-Ray computed tomography and magnetic resonance. Biomaterials 2012, 33, 5394–5405. [Google Scholar] [CrossRef]
  20. Ortgies, D.H.; Tan, M.; Ximendes, E.C.; Del Rosal, B.; Hu, J.; Xu, L.; Wang, X.; Martín Rodríguez, E.; Jacinto, C.; Fernandez, N.; et al. Lifetime-encoded infrared-emitting nanoparticles for in vivo multiplexed imaging. ACS Nano 2018, 12, 4362–4368. [Google Scholar] [CrossRef]
  21. Ragin, T.; Baranowska, A.; Kochanowicz, M.; Zmojda, J.; Miluski, P.; Dorosz, D. Study of Mid-Infrared Emission and Structural Properties of Heavy Metal Oxide Glass and Optical Fibre Co-Doped with Ho3+/Yb3+ Ions. Materials 2019, 12, 1238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Cao, W.; Feifei Huang, F.; Ye, R.; Cai, M.; Lei, R.; Zhang, J.; Shiqing Xu, S.; Zhang, X.H. Structural and fluorescence properties of Ho3+/Yb3+ doped germanosilicate glasses tailored by Lu2O3. J. Alloys Compd. 2018, 746, 540–548. [Google Scholar] [CrossRef]
  23. Kowalska, K.; Kuwik, M.; Pisarska, J.; Pisarski, W.A. Near-IR Luminescence of Rare-Earth Ions (Er3+, Pr3+, Ho3+, Tm3+) in Titanate–Germanate Glasses under Excitation of Yb3+. Materials 2022, 15, 3660. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, T.; Qin, J.; Zhang, J.; Guo, L.; Yang, M.; Wu, X.; You, M.; Peng, H. Recent Progresses in NIR-II Luminescent Bio/Chemo Sensors Based on Lanthanide Nanocrystals. Chemosensors 2022, 10, 206. [Google Scholar] [CrossRef]
  25. Ledoux, G.; Joubert, M.F.; Mishra, S. Upconversion Phenomena in Nanofluorides. In Photonic & Electronic Properties of Fluoride Materials, 1st ed.; Tressaud, A., Poeppelmeir, K.R., Eds.; Elsevier: Paris, France, 2016; pp. 35–63. [Google Scholar]
  26. Tiwari, S.P.; Maurya, S.K.; Yadav, R.S.; Kumar, A.; Kumar, V.; Joubert, M.-F.; Swart, H.C. Future prospects of fluoride based upconversion nanoparticles for emerging applications in biomedical and energy harvesting. J. Vac. Sci. Technol. B 2018, 36, 060801. [Google Scholar] [CrossRef] [Green Version]
  27. Krämer, K.W.; Biner, D.; Frei, G.; Güdel, H.U.; Hehlen, M.P.; Lüthi, S.R. Hexagonal Sodium Yttrium Fluoride Based Green and Blue Emitting Upconversion Phosphors. Chem. Mater. 2004, 16, 1244–1251. [Google Scholar] [CrossRef]
  28. Runowski, M.; Bartkowiak, A.; Majewska, M.; Martín, I.R.; Lis, S. Upconverting Lanthanide Doped Fluoride NaLuF4:Yb3+-Er3+-Ho3+- Optical Sensor for Multi-Range Fluorescence Intensity Ratio (FIR) Thermometry in Visible and NIR Regions. J. Lumin. 2018, 201, 104–109. [Google Scholar] [CrossRef]
  29. Gonçalves, I.M.; Pessoa, A.R.; Hazra, C.; Correales, Y.S.; Ribeiro, S.J.L.; Menezes, L.d.S. Phonon-assisted NIR-To-Visible Upconversion in Single β-NaYF4 Microcrystals Codoped with Er3+ and Yb3+ for Microthermometry Applications: Experiment and Theory. J. Lumin. 2021, 231, 117801. [Google Scholar] [CrossRef]
  30. Li, J.; Wang, Y.; Zhang, X.; Li, L.; Hao, H. Up-Converting Luminescence and Temperature Sensing of Er3+/Tm3+/Yb3+ Co-Doped NaYF4 Phosphors Operating in Visible and the First Biological Window Range. Nanomaterials 2021, 11, 2660. [Google Scholar] [CrossRef]
  31. Grzyb, T.; Przybylska, D. Formation Mechanism, Structural, and Upconversion Properties of Alkaline Rare-Earth Fluoride Nanocrystals Doped with Yb3+/Er3+ Ions. Inorg. Chem. 2018, 57, 6410–6420. [Google Scholar] [CrossRef]
  32. Xia, Z.; Du, P.; Liao, L. Facile hydrothermal synthesis and upconversion luminescence of tetragonal Sr2LnF7:Yb3+/Er3+ (Ln=Y, Gd) nanocrystals. Phys. Status Solidi A 2013, 210, 1734–1737. [Google Scholar] [CrossRef]
  33. Mao, Y.; Ma, M.; Gong, L.; Xu, C.; Ren, G.; Yang, Q. Controllable synthesis and upconversion emission of ultrasmall near-monodisperse lanthanide-doped Sr2LaF7 nanocrystals. J. Alloys Compd. 2014, 609, 262–267. [Google Scholar] [CrossRef]
  34. Xie, J.; Bin, J.; Guan, M.; Liu, H.; Yang, D.; Xue, J.; Liao, L.; Mei, L. Hydrothermal Synthesis and Upconversion Luminescent Properties of Sr2LaF7 Doped with Yb3+ and Er3+ Nanophosphors. J. Lumin. 2018, 200, 133–140. [Google Scholar] [CrossRef]
  35. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomie Distances in Halides and Chaleogenides. Acta Crystallogr. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  36. Wang, Z.; Li, Y.; Jiang, Q.; Zeng, H.; Ci, Z.; Sun, L. Pure near-infrared to near-infrared upconversion of multifunctional Tm3+ and Yb3+ co-doped NaGd(WO4)2 nanoparticles. J. Mater. Chem. C 2014, 2, 4495–4501. [Google Scholar] [CrossRef]
  37. Alkahtani, M.; Alfahd, A.; Alsofyani, N.; Almuqhim, A.A.; Qassem, H.; Alshehri, A.A.; Almughem, F.A.; Hemmer, P. Photostable and Small YVO4:Yb,Er Upconversion Nanoparticles in Water. Nanomaterials 2021, 11, 1535. [Google Scholar] [CrossRef]
  38. Li, L.; Pan, Y.; Chang, W.; Feng, Z.; Chen, P.; Li, C.; Zeng, Z.; Zhou, X. Near-infrared downconversion luminescence of SrMoO4:Tm3+,Yb3+ phosphors. Mater. Res. Bull. 2017, 93, 144–149. [Google Scholar] [CrossRef]
  39. Simpson, D.A.; Gibbs, W.E.K.; Collins, S.F.; Blanc, W.; Dussardier, B.; Monnom, G.; Peterka, P.; Baxter, G.W. Visible and near infra-red up-conversion in Tm3+/Yb3+ co-doped silica fibers under 980 nm excitation. Opt. Express 2008, 16, 13781–13799. [Google Scholar] [CrossRef]
  40. Güell, F.; Solé, R.; Gavaldà, J.; Aguiló, M.; Galán, M.; Díaz, F.; Massons, J. Upconversion luminescence of Tm3+ sensitized by Yb3+ ions in monoclinic KGd(WO4)2 single crystals. Opt. Mater. 2007, 30, 222–226. [Google Scholar] [CrossRef]
Figure 1. Synthesis of SLF:Yb,Tm nanophosphors using a simple EDTA-assisted hydrothermal method.
Figure 1. Synthesis of SLF:Yb,Tm nanophosphors using a simple EDTA-assisted hydrothermal method.
Nanomaterials 13 00030 g001
Figure 2. XRD patterns of (a) SLF:xYb3+,1 mol%Tm3+ and (b) SLF:20 mol%Yb3+,yTm3+ nanophosphors. (c) The evolution of the (111) diffraction peak magnified from (a). (d) The evolution of the (111) diffraction peak magnified from (b). Arrows indicate changes of the peak position maximum.
Figure 2. XRD patterns of (a) SLF:xYb3+,1 mol%Tm3+ and (b) SLF:20 mol%Yb3+,yTm3+ nanophosphors. (c) The evolution of the (111) diffraction peak magnified from (a). (d) The evolution of the (111) diffraction peak magnified from (b). Arrows indicate changes of the peak position maximum.
Nanomaterials 13 00030 g002
Figure 3. I. Undoped SLF: (ad) TEM images of hydrothermally synthesized SLF, (e) HRTEM image of SLF, (f) particle size distribution of SLF. II. SLF:10Yb,1Tm: (ad) TEM images (e) HRTEM image, (f) particle size distribution. III. SLF:25Yb,1Tm: (ad) TEM images, (e) HRTEM image, (f) particle size distribution.
Figure 3. I. Undoped SLF: (ad) TEM images of hydrothermally synthesized SLF, (e) HRTEM image of SLF, (f) particle size distribution of SLF. II. SLF:10Yb,1Tm: (ad) TEM images (e) HRTEM image, (f) particle size distribution. III. SLF:25Yb,1Tm: (ad) TEM images, (e) HRTEM image, (f) particle size distribution.
Nanomaterials 13 00030 g003
Figure 4. (a) Diffuse reflectance spectra of representative SLF:20Yb,1Tm nanophosphor. (b) PL spectra of SLF: xYb, yTm nanophosphors measured upon 980 nm excitation (Inset: UC lifetime profiles of Tm3+ (3H43H6 transition) of representative SLF:20Yb,1Tm). (c) Yb3+ concentration dependence of integrated UC emission intensity of SLF nanophosphor. (d) Tm3+ concentration dependence of integrated UC emission intensity of SLF nanophosphor.
Figure 4. (a) Diffuse reflectance spectra of representative SLF:20Yb,1Tm nanophosphor. (b) PL spectra of SLF: xYb, yTm nanophosphors measured upon 980 nm excitation (Inset: UC lifetime profiles of Tm3+ (3H43H6 transition) of representative SLF:20Yb,1Tm). (c) Yb3+ concentration dependence of integrated UC emission intensity of SLF nanophosphor. (d) Tm3+ concentration dependence of integrated UC emission intensity of SLF nanophosphor.
Nanomaterials 13 00030 g004
Table 1. Results of the structural analysis by using WPF refinement of SLF:xYb3+, 1 mol%Tm3+ nanophosphors.
Table 1. Results of the structural analysis by using WPF refinement of SLF:xYb3+, 1 mol%Tm3+ nanophosphors.
SLFSLF:10Yb,1TmSLF:15Yb,1TmSLF:20Yb,1 TmSLF:25Yb,1Tm
CS (nm)27.123.52525.424.4
Strain0.1440.2540.2540.2720.258
* Rwp6.004.744.584.334.09
** Rp4.623.793.513.353.28
*** Re3.423.483.373.373.28
GOF1.75691.35971.35971.28431.2484
a = b = c (Å)5.84515.83725.81905.81155.8045
CV3)199.70198.89197.04196.27195.57
* Rwp—the weighted profile factor; ** Rp—the profile factor; *** Re—the expected weighted profile factor; GOF—the goodness of fit.
Table 2. Results of the structural analysis by using WPF refinement of SLF:20 mol%Yb3+,yTm3+ nanophosphors.
Table 2. Results of the structural analysis by using WPF refinement of SLF:20 mol%Yb3+,yTm3+ nanophosphors.
SLFSLF:20Yb,0.75TmSLF:20Yb,1TmSLF:20Yb,2TmSLF:20Yb,3Tm
CS (nm)27.122.625.421.917.8
Strain0.1440.2110.2720.1850.02
* Rwp6.004.124.334.915.97
** Rp4.623.143.353.904.41
*** Re3.423.563.373.333.39
GOF1.75691.15891.28431.47471.7607
a = b = c (Å)5.84515.81295.81155.81075.8288
CV3)199.70196.42196.27196.20198.03
* Rwp—the weighted profile factor; ** Rp—the profile factor; *** Re–the expected weighted profile factor; GOF—the goodness of fit.
Table 3. Summary of the different parameters used to calculate PL lifetime of the representative SLF:20Yb,1Tm nanophosphor.
Table 3. Summary of the different parameters used to calculate PL lifetime of the representative SLF:20Yb,1Tm nanophosphor.
A 1 τ 1   ( m s ) A 2 τ 2   ( m s ) b g τ   ( m s )
0.68870.45030.31632.36070.00301.0516
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Milićević, B.; Periša, J.; Ristić, Z.; Milenković, K.; Antić, Ž.; Smits, K.; Kemere, M.; Vitols, K.; Sarakovskis, A.; Dramićanin, M.D. Hydrothermal Synthesis and Properties of Yb3+/Tm3+ Doped Sr2LaF7 Upconversion Nanoparticles. Nanomaterials 2023, 13, 30. https://doi.org/10.3390/nano13010030

AMA Style

Milićević B, Periša J, Ristić Z, Milenković K, Antić Ž, Smits K, Kemere M, Vitols K, Sarakovskis A, Dramićanin MD. Hydrothermal Synthesis and Properties of Yb3+/Tm3+ Doped Sr2LaF7 Upconversion Nanoparticles. Nanomaterials. 2023; 13(1):30. https://doi.org/10.3390/nano13010030

Chicago/Turabian Style

Milićević, Bojana, Jovana Periša, Zoran Ristić, Katarina Milenković, Željka Antić, Krisjanis Smits, Meldra Kemere, Kaspars Vitols, Anatolijs Sarakovskis, and Miroslav D. Dramićanin. 2023. "Hydrothermal Synthesis and Properties of Yb3+/Tm3+ Doped Sr2LaF7 Upconversion Nanoparticles" Nanomaterials 13, no. 1: 30. https://doi.org/10.3390/nano13010030

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