Abstract
Upconverting luminescent nanoparticles are photostable, nonblinking, and low chemically toxic fluorophores that are emerging as promising fluorescent probes at the single molecule level. High luminescence intensity upconversion nanoparticles (UCNPs) have previously been achieved by doping with high amounts of rare-earth ions using high excitation power (>2.5 MW/cm2). However, such particles are inadequate for in vitro live-cell imaging and single-particle tracking, as high excitation power can cause photodamage. Here, we compared UCNP luminescence intensities with different dopant concentrations and presented more efficient (about seven times) UCNPs at low excitation power by increasing the concentrations of Yb3+ and Tm3+ dopants (NaYF4: 60% Yb3+, 8% Tm3+) and adding a core-shell structure.
1 Introduction
Upconversion nanoparticles (UCNPs) possess the ability to convert low-energy infrared (IR) photons into high-energy visible or ultraviolet (UV) photons. This unique anti-Stokes feature has enabled various applications over the past decade, spanning from solar energy harvesting, superresolution microscopy, and anticounterfeit luminescent imprinting to agricultural crop care [1], [2], [3], [4], [5], [6]. Among the various nanoparticle systems developed thus far, semiconductor nanocrystals, 2D materials, and quantum dots (QDs) are widely used [7], [8], [9], [10]. UCNPs typically consist of a NaYF4 crystalline host matrix doped with rare-earth-based lanthanide sensitizer and activator. Such particles can be designed and synthesized based on desirable physicochemical characteristics such as sizes, shapes, optical properties, and magnetic properties [11], [12], [13], [14].
The development of near-IR (NIR)-responsive photonic materials is of great scientific interest, as life science, materials science, and aerospace industries can all take advantage of the penetrative ability of NIR light for invisible, high-precision remote activation of materials [1]. UCNPs can efficiently convert NIR light to UV or visible light and have attracted recent attention in biomedical applications for this ability. Nanoparticles feature efficient cellular uptake while retaining low cytotoxicity and display high optical penetrating power in deep tissue with minimal background noise [15], [16]. In addition, UCNPs exhibit neither photoblinking on millisecond and second-time scales nor photobleaching even with hours of continuous excitation. As a result, UCNPs are becoming one of the most promising nanoparticle systems for in vivo biological imaging, biosensing, and nanomedicine, with continuing efforts to improve their properties including the design of new synthetic strategies [17].
Single nanoparticle imaging is essential for cell tracking. Studies include long-term tracking of UCNPs in live cells [18], deep subdiffraction fluorescence microscopy [19], single-molecule tracking under lower irradiance [20], using agarose-gel electrophoresis to make background-free labels [21], and improving upconverting fluorescence with natural biomicrolens [22]. Single nanoparticle imaging is essential for the detection of single-cell imaging in techniques involving cellular tracking and providing accurate cellular movement.
For bioimaging applications in vitro, high excitation power is not recommended. Imaging and longitudinal real-time tracking of live cells prefer low power excitation. For example, power-dependent analysis is often used in various studies, including investigating energy transference relations between dopants [23], economically feasible UCNPs in anticounterfeiting ink printing [24], thin-film synthesis using electrodeposition [25], and upconverting luminescence depletion phenomena via simultaneous wavelength excitation [26]. Therefore, in practical usage of UCNPs in bioinformatics, maximizing luminescence efficiencies is of great significance. In this article, emphasis is placed on such augmentation of upconversion intensity in response to power irradiance. In this study, Tm3+ ions are chosen as dopants due to their emission wavelength of ~800 nm, which is within the “NIR biological window”. This allows for low autofluorescence and reduced photon-damage effects but also offers deep light penetration and low light scattering [27], [28].
Many studies have aimed to improve upconverting luminescence by the advancement of synthesis methodology and instrumental characterization at nanoscale over the last decades. Enhancement strategies include optimizing dopant concentrations and crystal hosts, coating inert/active shells around UCNPs, tuning the sizes of UCNPs, and coupling UCNPs with plasmonic noble metal nanostructures. One of the most commonly used of these methods is the passivation of the particle’s surface with an inert shell such as NaYF4, NaGdF4, NaLuF4, or CaF2 to form a dopant free layer. This can block the pathway of energy migration to surface quenchers [29], [30] (Figure S3). In the synthesis of UCNPs, a higher dopant concentration of sensitizer ions will enhance luminescence. However, an oversaturation of sensitizer concentration will decrease luminescence [31]. This concentration quenching consequence is an important factor to take into consideration when addressing dopant concentrations in such particles [32], [33]. The addition of an inactive shell surrounding the particle provides a protective layer that prevents luminescence quenching associated with surface defects on the core particles but at a cost of increasing nanoparticle size [34]. It is found that typical lanthanide-based UCNPs containing smaller particles below 10 nm in size endure more difficulty in displaying efficient upconverting luminescence due to their cubic crystal lattice structure [35]. However, nanoparticle size for efficient cellular uptake favors smaller sizes, which can decrease the toxicity of the particles within subcellular regions and increase biocompatibility [36], [37].
There are several reports on acquiring high-contrast cellular imaging through testing with a customized confocal microscope with a moderate resolution in response to the saturation effect of upconverting luminescence [38]. The saturation effect is easily achieved with high numerical aperture (NA) confocal microscopy with a power density in the range from 10 kW/cm2 to 1 MW/cm2 [39]. For live cell imaging, a probe with high quantum efficiency and low saturated excitation threshold power will benefit from long-term live cell imaging or single-particle tracking.
In this article, multifarious methods to improve luminescence intensity in UCNPs are investigated. More power efficient UCNPs are synthesized by scrutinizing dopant concentration, particle size, and core-shell influence. The variation of luminescence efficiency with respect to each of these attributes is also quantified and compared. Emphasis is given to the parameter values at which higher-performing nanoparticles undergo increased luminescence at lower power density than their weaker-performing counterparts. We explored UCNPs with highly doped Yb3+ sensitizer ions and Tm3+ activator dopants in a NaYF4 host and their core-shell structures. This combination of dopants is typical lanthanide dopant used in UCNP research and has been used in bioinformatic investigations [40], [41]. With a dopant concentration of 60% Yb3+, 8% Tm3+ in a NaYF4 host, we present about seven times more efficient UCNP that will be well suited for applications in live cell imaging.
2 Results and discussion
There are various methods to increase upconverting luminescence with concentrated samples. We also found multiple methods to increase upconverting luminescence for a single nanoparticle [33], [42], as shown in Figure 1. Nanoparticle size is an important aspect that can influence upconverting luminescence, as different nanoparticle sizes adjust the surface/volume ratio and directly influence the defects/unit. The size of nanoparticles is controlled below 50 nm in consideration toward use for bioapplications. We use optimized synthesis methods to synthesize individual nanoparticles within similar sizes, as shown in Figures S1 and S2. Similar sizes will maintain the same surface/volume ratio, which will greatly eliminate the influence of surface defects.

Multiple methods to increase upconverting luminescence for a single nanoparticle.
(A) Schematic illustration of the system setup for customized LSM. The excitation source is a 980 nm CW laser (power density of 30 MW/cm2) that is focused on the sample through an oil immersion objective lens (100×, NA 1.4, Olympus). The dichroic mirror into the detector reflects the emission from the sample. A large field PMT (Multialkali Amplified PMT, PMT1001, Thorlabs) detector is employed to collect emission fluorescence photons. Scanning is achieved by two conjugated placed galvanometer mirrors. A motorized wave plate and a polarizing beam splitter (PBS) are employed to control the power, and a cover glass and photon diode (PD) is used to measure the power on the sample after objective remotely. (B) Spatial confinement of generated fluorescence generation with upconverting nonlinear excitation. In one photon fluorescence process, the whole illumination cone will generate a fluorescence signal, but the upconverting nonlinear generated signal only localized the focal spot. Upconverting nonlinear process generated a fluorescence photon by absorbing more than two photons, whereas one photon fluorescence process case absorbs one photon. TEM images and microscopy quantitative measurement of the whole upconverting spectrum luminescence emission of single UCNPs of (C) NaYF4: 4% Tm3+, 20% Yb3+ UCNPs; (D) NaYF4: 4% Tm3+, 45% Yb3+ UCNPs; (E) NaYF4: 8% Tm3+, 20% Yb3+ UCNPs; (F) NaYF4: 4% Tm3+, 20% Yb3+ @ NaYF4 UCNPs; and (G) NaYF4: 8% Tm3+, 60% Yb3+ @ NaYF4 UCNPs. Scale bar in TEM and fluorescence images, 100 nm and 1 μm, respectively. The pixel dwell time in fluorescence image acquisition is 200 μs. Comparison of integrated single-particle upconverting luminescence emissions of (H) core (4% Tm3+, 20% Yb3+) and core-shell and (i) NaYF4: 4% Tm3+, 20% Yb3+ vs. NaYF4: 8% Tm3+, 60% Yb3+ @ NaYF4 UCNPs under different excitation power.
As Yb3+ acts as a sensitizer that can transfer 980 nm phonon energy to rare-earth ions, it is assumed that high sensitizer and rare-earth ion dopant concentrations will make upconverting luminescence much brighter under high excitation power [2]. As shown in Figure 1D, 45% Yb3+-doped UCNP is 1.33 times higher than 20% Yb3+ (Figure 1C)-doped particles under 1 MW/cm2 power density. Increasing Yb3+ concentration will decrease the atomic distance from sensitizers to activators; thus, energy transfer will be more efficient when high excitation power continually transfers energy from Yb3+ to activators. The luminescence intensity of single nanoparticles with different dopant concentrations has been measured using a customized laser scanning microscope (LSM; Figure 1A). The luminescence intensity of a single nanoparticle is represented as the maximum pixel value for each Gaussian spot.
Based on a previous report, a brighter upconverting luminescence can be obtained under high excitation power in high activator concentration (4% Tm3+) codoped with 20% Yb3+ UCNPs [39]. We tried to use a higher activator concentration to further increase luminescence, which is 8% Tm3+-doped UCNPs. As shown in Figure 1E, the same trend occurs such that luminescence is enhanced 1.64 times higher.
Concentration quenching also can be overcome at high excitation powers by introducing a thin layer of an inert shell onto the surface of the nanoparticle. The high-resolution transmission electron microscopy (HRTEM) core and core-shell characterization in Figure S4 and the inductively coupled plasma-optical emission spectrometry composition data for core-shell versus core-only nanoparticles are shown in Table S2 to confirm the synthesis of this structure. In Figure 1F, the core-shell particle increases luminescence by 1.48 times compared to the particles, which only contain the core structure. Yb3+-Yb3+ energy migration to surface defects, which lead to quenching, is primarily responsible for the depopulation of core UCNPs. This result demonstrates that, at single-particle laser powers, quenching typically associated with high Yb3+ content can be suppressed by external inert shells.
At the same time, Tm3+ ions mainly produce 3H4→3H6 transitions at low excitation power, but a different situation appears when excited under high power. The transitions shown in Figure 1H, blue wavelength (1D2→3F4, 1G4→→ 3H6), red wavelength (1G4→→ 3F4), and NIR (1D2→3F2, 1D2→→ 3F3, 3H4 →3H6), are all verified in Figure 1B. From the different wavelength emission changes under different excitation power, we find that the electron population on 1D2 and 1G4 states increased faster than 3H4 state with increasing excitation power. The 3H4 state releases the 800 nm emission; however, when additional emission energy is transferred from Yb3+ under high excitation power, 3H4 state eventually achieves full capacity and is excited to higher 1D2 and 1G4 states.
It is worth noting that the luminescence intensity ratio difference between core-shell and core-only UCNPs is higher under 10 kW/cm2 (3.65 times) than 1 MW/cm2 (1.48 times), as shown in Figure 1H. Low power density is required for in vivo imaging; therefore, importance is placed on attaining the same luminescence intensity with lower excitation power density. Based on all listed luminescence enhancement methods, such intensity with low power density is produced using higher sensitizer and activator concentration combinations along with a core-shell structure. As Figure 1I shows, the luminescence of the core-shell structure with high dopant concentrations (8% Tm3+, 60% Yb3+, as shown in Figure 1G) is 6.83 times higher than normal UCNP particles under 0.05 MW/cm2.
To systematically compare the brightness of all the improved methods, we measured the power-dependent luminescence curves for single particles from 5 kW/cm2 to 10 MW/cm2, as shown in Figure 2C, D, G, and H. The quantitative intensity images are shown in Figure 2A, B, E, and F and every image is taken within the same area for each sample. From previous recordings of data obtained from the sensitizer concentration and power dependence testing, it is foreseen that luminescence would be enhanced with the increase in excitation power. Figure 2A and D (4% Tm3+-doped UCNPs) shows that higher sensitizer concentrations will not only increase luminescence intensity under each excitation power but can also shift power dependence curve to lower power. In other words, we can use lower power excitation with higher sensitizer concentration UCNPs to get the same intensity as that using high power excitation with lower sensitizer concentration. Figure 2B and F (8% Tm3+-doped UCNPs) shows the same trends as 4% Tm3+ particles. However, higher Tm3+ dopant concentrations will increase luminescence intensity under each excitation power.

Microscopy quantitative measurement of the whole upconverting spectrum luminescence emission.
(A) single 4% Tm3+ UCNPs vs. Yb3+ concentration, (B) single 8% Tm3+ UCNPs vs. Yb3+ concentration, (E) single 4% Tm3+ core-shell UCNPs vs. Yb3+ concentration, and (F) single 8% Tm3+ core-shell UCNPs vs. Yb3+ concentration. The power-dependent saturation curves of (C) single 4% Tm3+ UCNPs vs. Yb3+ concentration, (D) single 8% Tm3+ UCNPs vs. Yb3+ concentration, (G) single 4% Tm3+ core-shell UCNPs vs. Yb3+ concentration, and (H) single 8% Tm3+ core-shell UCNPs vs. Yb3+ concentration. The saturation curves under different power densities from 5 kW/cm2 to 10 MW/cm2 obtained with large field PMT detector-based LSM. A log scale is used for the y-axis in the left panels of (C and D) and (G and H), and a linear y-scale is employed in the right panels of (C and D) and (G and H). Scale bar, 1 μm; pixel dwell time, 200 μs. Error bars represent standard deviation from the mean.
Another remarkable phenomenon is the power dependence slope increase by introducing a core-shell structure, in which surface defects are decreased. We can use 0.1 MW/cm2 for 4% Tm3+/45% Yb3+ core-shell particles to obtain the same highest luminescence intensity of the related core particle under 0.01 MW/cm2. Comparing 4% Tm3+ UCNPs to 8% Tm3+-doped particles, the power dependence slope for 8% Tm3+ is higher than 4% Tm3+ UCNPs. This indicates that 8% Tm3+-doped UCNP can reach its saturation point at a much lower power than 4% Tm3+ UCNP. Also, 8% Tm3+-doped UCNP is able to reach the same luminescence intensity using much lower excitation power. For instance, the highest luminescence intensity of NaYF4: 4% Tm3+/20% Yb3+ @ NaYF4 UCNP is under 1 MW/cm2, and the same intensity can be reached under 0.02 MW/cm2 by NaYF4: 8% Tm3+/60% Yb3+ @ NaYF4 UCNP. Based on these results, excitation power can be decreased by more than one order of magnitude using high Tm3+ with high Yb3+ dopants. It is indicative that short distances between sensitizer and activator ions as well as high power excitations that increase the electron populations of excited states play an important role in protecting the surface quenching of same-sized UCNPs.
Spatial resolution and collection efficiency are great advantages of large field photomultiplier tube (PMT) detector-based LSM for high nonlinear upconverting fluorescence nanoparticles. Several groups have reported their results of single upconverting nanoparticle characterization employing confocal microscopy [20], [43], [44], [45], [46] as well as confocal based laser scanning subdiffraction microscopy [2], [19], [47]. The merits of nonlinear contrast mechanisms are that the signal (S) depends supralinearly (SμIn) to the excitation light intensity (I). As a result, when focusing on the laser beam through a microscope objective, high-order nonlinearity absorption is spatially confined to the perifocal region. The high-order nonlinear property of upconverting luminescence shown in Figure 2 can benefit not only the spatial resolution of LSM but also the out-of-focus background, which could be suppressed by high-order nonlinear generated fluorescence. Thus, a pinhole is not needed to reject out-of-focus background. Of equal importance to the maximization of signal generation is the optimization of collection and detection efficiencies. The pinhole will also degrade the collection efficiency of the generated fluorescence once it is not well coupled. High-order nonlinear microscopes (Figure 1A) use PMTs, available with large sensitive areas and reasonable quantum efficiencies rather than the pinhole-based detector and come with a high spatial resolution of nonlinear upconverting fluorescence nanoparticles.
The resolution of confocal images degrades as laser power increases because excitation power has exceeded the nonlinear threshold. For microscopy, the image of any object is acquired by convolving the point spread function (PSF) of the optical system [48]. For confocal microscopy, PSFconfocal=PSFillumination×PSFdetection. For Nth order nonlinearity LSM with whole area detection, the effective PSF, PSFN-order=(PSFillumination)N=PSFN (u, v) [48]. As shown in Figure 2, log-scale power dependence fluorescence saturated curves have a higher slope under low power density. For example, 4% Tm3+, 20% Yb3+ core UCNPs have a slope of ~1.6 below 1 MW/cm2 (saturated intensity), and 8% Tm3+, 60% Yb3+ core-shell UCNPs have a slope of ~2 with below the saturated intensity. In Figure 3, the image under lower power density of 5 kW/cm2 has a much smaller full-width at half-maximum (FWHM) of 279.0 nm compared to the case with 100 kW/cm2 (saturated absorbed). Note that 4% Tm3+, 20% Yb3+ core UCNPs have a much narrow PSF but really low efficiency of generating fluorescence under low density excitation power. However, 8% Tm3+, 60% Yb3+ core-shell UCNPs have also an enhanced resolution (294.4 nm) due to the high-order nonlinearity of upconverting fluorescence with decent fluorescence generating efficiency under low power excitation density (5 kW/cm2).

Laser scanning upconverting nonlinear image and saturated image of single 4% Tm3+, 20% Yb3+ core UCNPs and single 8% Tm3+, 60% Yb3+ core-shell UCNPs.
(A) 4% Tm3+, 20% Yb3+ core single UCNPs under nonlinear excitation (5×103 W/cm2) and saturated excitation (100×103 W/cm2), respectively. (B) 8% Tm3+, 60% Yb3+ core-shell single UCNPs under nonlinear excitation (5×103 W/cm2) and saturated excitation (100×103 W/cm2), respectively. Scale bar, 1 µm. The cross-section on the green dashed line in (A and B) are shown in (C and D). The fitted curves of nonlinear excitation (green) have a smaller FWHM than the saturated excitation ones (gray) due to different order of nonlinearity. (E and F) FWHM under different power density excitation of 4% Tm3+, 20% Yb3+ core UCNPs and 8% Tm3+, 60% Yb3+ core-shell UCNPs, respectively.
Using UCNPs with continuous wave (CW) IR excitation has similar merits as a multiphoton microscopy for deep tissue brain imaging, longer scattering length, and high-order nonlinearity. As the microscopy focus goes deep, the benefits of confocal detection diminish, and attention has to be paid to the detection of scattered light. Scattered light will spread to a volume that is much larger than when the excitation focus is deep below the scattering tissue. Large field detector-based LSM using nonlinear upconverting nanoprobe has the ability to realize deep tissue imaging, which is similar to femtosecond ultrafast laser-based multiphoton microscopy. Using upconverting microscopy with CW laser will make nonlinear deep tissue microscopy more cost-efficient and portable.
3 Conclusion
This work emphasizes the importance of controlling activator and sensitizer dopant concentrations to decrease excitation power for emitting the same upconverting luminescence intensity. We found the optimal activator and sensitizer concentration in single UCNPs that can increase upconverting luminescence and increase power dependence slopes. A shell coated on the surface to passivate the quenchers will shift the power dependence curve to lower excitation power. In addition, higher sensitizer concentration will help shift the power dependence curve to a lower excitation power. The core-shell structure will help improve luminescence intensity. This means that saturation intensity is more easily achieved under lower power excitation when activator concentration is increased.
Acknowledgments
The authors thank Jing Tai (Analysis and Testing Center, Institute of Chemistry Chinese Academy of Sciences) for help with TEM element mapping. Dr. Chunyan Shan thanks the support by the National Center for Protein Sciences at Peking University (Beijing, China).
References
[1] Chen EY, Milleville C, Zide JMO, Doty MF, Zhang J. Upconversion of low-energy photons in semiconductor nanostructures for solar energy harvesting. MRS Energy Sustain 2018;5:1–17.10.1557/mre.2018.15Search in Google Scholar
[2] Liu Y, Lu Y, Yang X, et al. Amplified stimulated emission in upconversion nanoparticles for super-resolution nanoscopy. Nature 2017;543:229–33.10.1038/nature21366Search in Google Scholar PubMed
[3] Wilhelm S. Perspectives for upconverting nanoparticles. ACS Nano 2017;11:10644–53.10.1021/acsnano.7b07120Search in Google Scholar PubMed
[4] Giust D, Lucío MI, El-Sagheer AH, et al. Graphene oxide-upconversion nanoparticle based portable sensors for assessing nutritional deficiencies in crops. ACS Nano 2018;12:6273–9.10.1021/acsnano.8b03261Search in Google Scholar PubMed
[5] Li Y, Wang T, Ren W, et al. BiOCl: Er3+ nanosheets with tunable thickness for photon avalanche phosphors. ACS Appl Nano Mater 2019;13:7652–60.10.1021/acsanm.9b01735Search in Google Scholar
[6] Li Y, Cheng Z, Yao L, Yang S, Zhang Y. Boosting NIR-driven photocatalytic activity of BiOBr:Yb3+/Er3+/Ho3+ nanosheets by enhanced green upconversion emissions via energy transfer from Er3+ to Ho3+ ions. ACS Sustain Chem Eng 2019;7:18185–96.10.1021/acssuschemeng.9b05878Search in Google Scholar
[7] Tao W, Kong N, Ji X, et al. Emerging two-dimensional monoelemental materials (Xenes) for biomedical applications. Chem Soc Rev 2019;48:2891–912.10.1039/C8CS00823JSearch in Google Scholar
[8] Zhang Y, Lim C-K, Dai Z, et al. Photonics and optoelectronics using nano-structured hybrid perovskite media and their optical cavities. Phys Rep 2019;795:1–51.10.1016/j.physrep.2019.01.005Search in Google Scholar
[9] Guo C, Suo H. Design of single-phased multicolor-emission phosphor for LED. Phosphors Up Conversion Nano Particles Quant Dots Their Appl 2016;1:459–508.10.1007/978-3-662-52771-9_15Search in Google Scholar
[10] Luo W, Liu Y, Chen X. Lanthanide-doped semiconductor nanocrystals: electronic structures and optical properties. Sci China Mater 2015;58:819–50.10.1007/s40843-015-0091-9Search in Google Scholar
[11] Lemyre JL, Ritcey AM. Synthesis of lanthanide fluoride nanoparticles of varying shape and size. Chem Mater 2005;17:3040–3.10.1021/cm0502065Search in Google Scholar
[12] Li X, Zhang F, Zhao D. Lab on upconversion nanoparticles: optical properties and applications engineering via designed nanostructure. Chem Soc Rev 2015;44:1346–78.10.1039/C4CS00163JSearch in Google Scholar
[13] Wang F, Han Y, Lim CS, et al. Simultaneous phase and size control of upconversion nanocrystals through lanthanide doping. Nature 2010;463:1061–5.10.1038/nature08777Search in Google Scholar PubMed
[14] Wang G, Peng Q, Li Y. Lanthanide-doped nanocrystals: synthesis, optical-magnetic properties, and applications. Acc Chem Res 2011;44:322–32.10.1021/ar100129pSearch in Google Scholar PubMed
[15] Liu X, Deng R, Zhang Y, et al. Probing the nature of upconversion nanocrystals: instrumentation matters. Chem Soc Rev 2015;44:1479–508.10.1039/C4CS00356JSearch in Google Scholar PubMed
[16] Loo JF-C, Chien Y-H, Yin F, Kong S-K, Ho H-P, Yong K-T. Upconversion and downconversion nanoparticles for biophotonics and nanomedicine. Coord Chem Rev 2019;400:213042.10.1016/j.ccr.2019.213042Search in Google Scholar
[17] Wu X, Chen G, Shen J, Li Z, Zhang Y, Han G. Upconversion nanoparticles: a versatile solution to multiscale biological imaging. Bioconjug Chem 2015;26:166–75.10.1021/bc5003967Search in Google Scholar PubMed PubMed Central
[18] Wang F, Wen S, He H, et al. Microscopic inspection and tracking of single upconversion nanoparticles in living cells. Light Sci Appl 2018;7:18006–7.10.1038/lsa.2018.7Search in Google Scholar PubMed PubMed Central
[19] Chen C, Wang F, Wen S, et al. Multi-photon near-infrared emission saturation nanoscopy using upconversion nanoparticles. Nat Commun 2018;9:4–9.10.1038/s41467-018-05842-wSearch in Google Scholar PubMed PubMed Central
[20] Liu Q, Zhang Y, Peng CS, Yang T, Joubert L-M, Chu S. Single upconversion nanoparticle imaging at sub-10 W cm−2 irradiance. Nat Photonics 2018;12:548–53.10.1038/s41566-018-0217-1Search in Google Scholar PubMed PubMed Central
[21] Hlaváček A, Mickert MJ, Soukka T, et al. Large-scale purification of photon-upconversion nanoparticles by gel electrophoresis for analogue and digital bioassays. Anal Chem 2019;91:1241–46.10.1021/acs.analchem.8b04488Search in Google Scholar PubMed
[22] Li Y, Liu X, Yang X, Lei H, Zhang Y, Li B. Enhancing upconversion fluorescence with a natural bio-microlens. ACS Nano 2017;11:10672–80.10.1021/acsnano.7b04420Search in Google Scholar PubMed
[23] Xue X, Thitsa M, Cheng T, et al. Laser power density dependent energy transfer between Tm3+ and Tb3+: tunable upconversion emissions in NaYF4:Tm3+, Tb3+, Yb3+ microcrystals. Opt Express 2016;24:26307.10.1364/OE.24.026307Search in Google Scholar PubMed
[24] Yao W, Tian Q, Liu J, et al. Large-scale synthesis and screen printing of upconversion hexagonal-phase NaYF4:Yb3+, Tm3+/Er3+/Eu3+ plates for security applications. J Mater Chem C 2016;4:6327–35.10.1039/C6TC01513ASearch in Google Scholar
[25] Jia H, Xu C, Wang J, Chen P, Liu X, Qiu J. Synthesis of NaYF4:Yb-Tm thin film with strong NIR photon up-conversion photoluminescence using electro-deposition method. CrystEngComm 2014;16:4023–8.10.1039/C4CE00078ASearch in Google Scholar
[26] Zhang H, Jia T, Chen L, et al. Depleted upconversion luminescence in NaYF4:Yb3+, Tm3+ nanoparticles: via simultaneous two-wavelength excitation. Phys Chem Chem Phys 2017;19:17756–64.10.1039/C7CP00099ESearch in Google Scholar PubMed
[27] Li X, Wang R, Zhang F, et al. Nd3+ sensitized up/down converting dual-mode nanomaterials for efficient in-vitro and in-vivo bioimaging excited at 800 nm. Sci Rep 2013; 3:1–7.10.1038/srep03536Search in Google Scholar PubMed PubMed Central
[28] Becker A, Hessenius C, Licha K, et al. Receptor-targeted optical imaging of tumors with near-infrared fluorescent ligands. Nat Biotechnol 2001;19:327–31.10.1038/86707Search in Google Scholar PubMed
[29] Wang F, Wang J, Liu X. Direct evidence of a surface quenching effect on size-dependent luminescence of upconversion nanoparticles. Angew Chem Int Ed 2010;49:7456–60.10.1002/anie.201003959Search in Google Scholar PubMed
[30] Johnson NJJ, He S, Diao S, Chan EM, Dai H, Almutairi A. Direct evidence for coupled surface and concentration quenching dynamics in lanthanide-doped nanocrystals. J Am Chem Soc 2017;139:3275–82.10.1021/jacs.7b00223Search in Google Scholar PubMed
[31] Zheng X, Shikha S, Zhang Y. Elimination of concentration dependent luminescence quenching in surface protected upconversion nanoparticles. Nanoscale 2018;10: 16447–54.10.1039/C8NR03121ESearch in Google Scholar
[32] Wu X, Zhang Y, Takle K, et al. Dye-sensitized core/active shell upconversion nanoparticles for optogenetics and bioimaging applications. ACS Nano 2016;10:1060–6.10.1021/acsnano.5b06383Search in Google Scholar PubMed PubMed Central
[33] Ma C, Xu X, Wang F, et al. Optimal sensitizer concentration in single upconversion nanocrystals. Nano Lett 2017;17:2858–64.10.1021/acs.nanolett.6b05331Search in Google Scholar PubMed
[34] Fan Y, Liu L, Zhang F. Exploiting lanthanide-doped upconversion nanoparticles with core/shell structures. Nano Today 2019;25:68–84.10.1016/j.nantod.2019.02.009Search in Google Scholar
[35] Chen G, Ohulchanskyy TY, Kumar R, Ågren H, Prasad PN. Ultrasmall monodisperse NaYF4:Yb3+/Tm3+ nanocrystals with enhanced near-infrared to near-infrared upconversion photoluminescence. ACS Nano 2010;4:3163–8.10.1021/nn100457jSearch in Google Scholar PubMed PubMed Central
[36] Gnach A, Lipinski T, Bednarkiewicz A, Rybka J, Capobianco JA. Upconverting nanoparticles: assessing the toxicity. Chem Soc Rev 2015;44:1561–84.10.1039/C4CS00177JSearch in Google Scholar PubMed
[37] Chithrani BD, Ghazani AA, Chan WCW. Determining the size and shape dependence of gold nanoparticle uptake into mammalian cells. Nano Lett 2006;6:662–8.10.1021/nl052396oSearch in Google Scholar PubMed
[38] Peng X, Huang B, Pu R, et al. Fast upconversion super-resolution microscopy with 10 μs per pixel dwell times. Nanoscale 2019;11:1563–9.10.1039/C8NR08986HSearch in Google Scholar
[39] Zhao J, Jin D, Schartner EP, et al. Single-nanocrystal sensitivity achieved by enhanced upconversion luminescence. Nat Nanotechnol 2013;8:729–34.10.1038/nnano.2013.171Search in Google Scholar PubMed
[40] Zhang X, Guo Z, Zhang X, et al. Mass production of poly(ethylene glycol) monooleate-modified core-shell structured upconversion nanoparticles for bio-imaging and photodynamic therapy. Sci Rep 2019;9:1–10.10.1038/s41598-019-41482-wSearch in Google Scholar PubMed PubMed Central
[41] Zhang S, Wang J, Xu W, et al. Fluorescence resonance energy transfer between NaYF4:Yb, Tm upconversion nanoparticles and gold nanorods: Near-infrared responsive biosensor for streptavidin. J Lumin 2014;147:278–83.10.1016/j.jlumin.2013.11.052Search in Google Scholar
[42] Ma C, Xu X, Wang F, et al. Probing the interior crystal quality in the development of more efficient and smaller upconversion nanoparticles. J Phys Chem Lett 2016;7:3252–8.10.1021/acs.jpclett.6b01434Search in Google Scholar PubMed
[43] Bhawalkar JD, Swiatkiewicz J, Pan SJ, et al. Three-dimensional laser scanning two-photon fluorescence confocal microscopy of polymer materials using a new, efficient upconverting fluorophore. Scanning 1996;18:562–6.10.1002/sca.4950180805Search in Google Scholar PubMed
[44] Yin Z, Zhu Y, Xu W, et al. Remarkable enhancement of upconversion fluorescence and confocal imaging of PMMA opal/NaYF4:Yb3+, Tm3+/Er3+ nanocrystals. Chem Commun 2013;49:3781–3.10.1039/c3cc40829aSearch in Google Scholar PubMed
[45] Liu Q, Yang T, Feng W, Li F. Blue-emissive upconversion nanoparticles for low-power-excited bioimaging in vivo. J Am Chem Soc 2012;134:5390–7.10.1021/ja3003638Search in Google Scholar PubMed
[46] Lu Y, Zhao J, Zhang R, et al. Tunable lifetime multiplexing using luminescent nanocrystals. Nat Photonics 2014;8:32–6.10.1038/nphoton.2013.322Search in Google Scholar
[47] Zhan Q, Liu H, Wang B, et al. Achieving high-efficiency emission depletion nanoscopy by employing cross relaxation in upconversion nanoparticles. Nat Commun 2017;8:1–11.10.1038/s41467-017-01141-ySearch in Google Scholar PubMed PubMed Central
[48] Helmchen F, Denk W. Deep tissue two-photon microscopy. Nat Methods 2005;2:932–40.10.1038/nmeth818Search in Google Scholar PubMed
Supplementary Material
The online version of this article offers supplementary material (https://doi.org/10.1515/nanoph-2019-0526).
© 2020 Chenshuo Ma and Xusan Yang et al., published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Editorial
- 2D Xenes: from fundamentals to applications
- Reviews
- Monolayer MoS2 for nanoscale photonics
- 2D photonic memristor beyond graphene: progress and prospects
- MXenes: focus on optical and electronic properties and corresponding applications
- Advances in photonics of recently developed Xenes
- Nonlinear optical properties of anisotropic two-dimensional layered materials for ultrafast photonics
- Tunable electronic structure of two-dimensional transition metal chalcogenides for optoelectronic applications
- Recent advances in graphene and black phosphorus nonlinear plasmonics
- Fabrication, optical properties, and applications of twisted two-dimensional materials
- Novel layered 2D materials for ultrafast photonics
- 2D organic-inorganic hybrid perovskite materials for nonlinear optics
- Fine structures of valley-polarized excitonic states in monolayer transitional metal dichalcogenides
- MXenes for future nanophotonic device applications
- Two-dimensional nanomaterials for Förster resonance energy transfer–based sensing applications
- 2D materials integrated with metallic nanostructures: fundamentals and optoelectronic applications
- Graphene plasmonic devices for terahertz optoelectronics
- Research Articles
- Real-time dynamics of soliton collision in a bound-state soliton fiber laser
- Ultra-strong anisotropic photo-responsivity of bilayer tellurene: a quantum transport and time-domain first principle study
- Topological insulator overlayer to enhance the sensitivity and detection limit of surface plasmon resonance sensor
- Magnons scattering induced photonic chaos in the optomagnonic resonators
- Quantum confinement-induced enhanced nonlinearity and carrier lifetime modulation in two-dimensional tin sulfide
- Phosphorene-assisted silicon photonic modulator with fast response time
- High-performance monolayer MoS2 photodetector enabled by oxide stress liner using scalable chemical vapor growth method
- Enhancing the generating and collecting efficiency of single particle upconverting luminescence at low power excitation
- Biexcitons in 2D (iso-BA)2PbI4 perovskite crystals
- Broadband nonlinear optical response in GeSe nanoplates and its applications in all-optical diode
- Plasmonic nanocavity enhanced vibration of graphene by a radially polarized optical field
- Facile synthesis of sulfur@titanium carbide Mxene as high performance cathode for lithium-sulfur batteries
- The pump fluence and wavelength-dependent ultrafast carrier dynamics and optical nonlinear absorption in black phosphorus nanosheets
- Indium selenide film: a promising saturable absorber in 3- to 4-μm band for mid-infrared pulsed laser
- Temperature-stable black phosphorus field-effect transistors through effective phonon scattering suppression on atomic layer deposited aluminum nitride
- Real-time and noninvasive tracking of injectable hydrogel degradation using functionalized AIE nanoparticles
- MXene-Ti3C2 assisted one-step synthesis of carbon-supported TiO2/Bi4NbO8Cl heterostructures for enhanced photocatalytic water decontamination
- Nanofocusing of acoustic graphene plasmon polaritons for enhancing mid-infrared molecular fingerprints
- Effects of gap thickness and emitter location on the photoluminescence enhancement of monolayer MoS2 in a plasmonic nanoparticle-film coupled system
Articles in the same Issue
- Editorial
- 2D Xenes: from fundamentals to applications
- Reviews
- Monolayer MoS2 for nanoscale photonics
- 2D photonic memristor beyond graphene: progress and prospects
- MXenes: focus on optical and electronic properties and corresponding applications
- Advances in photonics of recently developed Xenes
- Nonlinear optical properties of anisotropic two-dimensional layered materials for ultrafast photonics
- Tunable electronic structure of two-dimensional transition metal chalcogenides for optoelectronic applications
- Recent advances in graphene and black phosphorus nonlinear plasmonics
- Fabrication, optical properties, and applications of twisted two-dimensional materials
- Novel layered 2D materials for ultrafast photonics
- 2D organic-inorganic hybrid perovskite materials for nonlinear optics
- Fine structures of valley-polarized excitonic states in monolayer transitional metal dichalcogenides
- MXenes for future nanophotonic device applications
- Two-dimensional nanomaterials for Förster resonance energy transfer–based sensing applications
- 2D materials integrated with metallic nanostructures: fundamentals and optoelectronic applications
- Graphene plasmonic devices for terahertz optoelectronics
- Research Articles
- Real-time dynamics of soliton collision in a bound-state soliton fiber laser
- Ultra-strong anisotropic photo-responsivity of bilayer tellurene: a quantum transport and time-domain first principle study
- Topological insulator overlayer to enhance the sensitivity and detection limit of surface plasmon resonance sensor
- Magnons scattering induced photonic chaos in the optomagnonic resonators
- Quantum confinement-induced enhanced nonlinearity and carrier lifetime modulation in two-dimensional tin sulfide
- Phosphorene-assisted silicon photonic modulator with fast response time
- High-performance monolayer MoS2 photodetector enabled by oxide stress liner using scalable chemical vapor growth method
- Enhancing the generating and collecting efficiency of single particle upconverting luminescence at low power excitation
- Biexcitons in 2D (iso-BA)2PbI4 perovskite crystals
- Broadband nonlinear optical response in GeSe nanoplates and its applications in all-optical diode
- Plasmonic nanocavity enhanced vibration of graphene by a radially polarized optical field
- Facile synthesis of sulfur@titanium carbide Mxene as high performance cathode for lithium-sulfur batteries
- The pump fluence and wavelength-dependent ultrafast carrier dynamics and optical nonlinear absorption in black phosphorus nanosheets
- Indium selenide film: a promising saturable absorber in 3- to 4-μm band for mid-infrared pulsed laser
- Temperature-stable black phosphorus field-effect transistors through effective phonon scattering suppression on atomic layer deposited aluminum nitride
- Real-time and noninvasive tracking of injectable hydrogel degradation using functionalized AIE nanoparticles
- MXene-Ti3C2 assisted one-step synthesis of carbon-supported TiO2/Bi4NbO8Cl heterostructures for enhanced photocatalytic water decontamination
- Nanofocusing of acoustic graphene plasmon polaritons for enhancing mid-infrared molecular fingerprints
- Effects of gap thickness and emitter location on the photoluminescence enhancement of monolayer MoS2 in a plasmonic nanoparticle-film coupled system