Open Access
16 April 2015 Labeling of HeLa cells using ZrO2:Yb3+-Er3+ nanoparticles with upconversion emission
Author Affiliations +
Abstract
This work reports the synthesis, structural characterization, and optical properties of ZrO2:Yb3+-Er3+ (2–1 mol%) nanocrystals. The nanoparticles were coated with 3-aminopropyl triethoxysilane (APTES) and further modified with biomolecules, such as Biotin-Anti-rabbit (mouse IgG) and rabbit antibody-AntiKi-67, through a conjugation method. The conjugation was successfully confirmed by Fourier transform infrared, zeta potential, and dynamic light scattering. The internalization of the conjugated nanoparticles in human cervical cancer (HeLa) cells was followed by two-photon confocal microscopy. The ZrO2:Yb3+-Er3+ nanocrystals exhibited strong red emission under 970-nm excitation. Moreover, the luminescence change due to the addition of APTES molecules and biomolecules on the nanocrystals was also studied. These results demonstrate that ZrO2:Yb3+-Er3+ nanocrystals can be successfully functionalized with biomolecules to develop platforms for biolabeling and bioimaging.

1.

Introduction

Lanthanide-doped nanomaterials are promising platforms for bioapplications due to their ability to convert low-energy near-infrared (NIR) radiation into higher-energy visible luminescence through a process called upconversion (UPC).1,2 There are several potential benefits for the use of nanocrystals with UPC emission in biological applications, such as no damage of tissues; anti-Stokes emission; long lifetimes; photostability; increased contrast in biological specimens due to the absence of autofluorescence upon excitation with IR light; and simultaneous detection of multiple targeted analytes.36 Other advantages of the UPC emission are the reduction of photobleaching and scattering in tissues, which avoid the use of complicated and high-cost femtosecond lasers and photomultiplier tubes.710

For biomedical applications, such as cancer detection, biolabeling, and bioimaging, luminescent nanoparticles preferably have to form a stable colloidal solution under physiological conditions. However, common nanomaterials with strong UPC emission, such as Yb3+-Er3+ co-doped Y2O2S, Yb3+-Ho3+ co-doped Y2O3, and Yb3+-Er3+-Tm3+ doped NaYF4, are hydrophobic.1113 Some efforts have been made to convert hydrophobic UPC nanoparticles into hydrophilic ones using techniques such as polymer capping, surface silanization, and surface ligand oxidation.1318 Recent methods also include ligand exchange in NaGdF4:Ho3+-Yb3+ and NaYF4:Yb3+-Er3+ phosphors.19

Cancer detection in early stages is a priority for many medical groups around the world. In 2012, according to World Health Organization, cervical cancer was one of the most prevalent cancer types in the world. To detect and diagnose cancer, there are several biomarkers;2023 for example, the Ki-67 protein is expressed in all phases of the cell division cycle, but its expression level is strongly downregulated in the resting G0 phase. This characteristic makes the Ki-67 protein an excellent biomarker for cell proliferation.2426 This biomolecule can be used as a prognostic marker in many types of cancers.2731 Moreover, it has been demonstrated that cervical human cancer (HeLa) cells can be labeled using doped or undoped nanomaterials, such as NaYF4:Yb3+-Er3+, NaYF4:Yb3+-Er3+@CaF2 core@shell, NaGdF4:Yb3+-Er3+/Silica/Au, CaF2 and carbon nanoparticles. These nanomaterials were internalized in HeLa cells observing visible light from the nanoparticles under IR excitation.3236 Though these platforms are efficient for labeling HeLa cells, they still show several problems related to the complexity of their fabrication. For example, the synthesis of NaGdF4:Yb3+, Er3+ nanoparticles have some drawbacks for biomedical applications; therefore, gold or silica need to be used to render them with biocompatible properties.3739 In addition, one of the problems with carbon nanoparticles is the fact that they need to be excited with near-UV light, which can damage tissues around the cancer cells.34

Rare earth doped zirconia (ZrO2) nanophosphors present efficient emission in the visible region when they are under IR excitation.4042 The ZrO2 low phonon energy (470cm1) increases the number and the probability of radiative transitions in rare earth doped ZrO2.43 Strong UPC emission has been obtained by doping ZrO2 with different pairs of rare earths, such as Yb3+-Tm3+, Yb-Ho3+, Er3+, and Yb3+-Er3+.44 Furthermore, ZrO2 nanophosphors can be synthesized by low-cost methods, such as sol-gel,45,46 sol-emulsion-gel,47,48 spray pyrolysis,49,50 and precipitation.51 Interestingly, ZrO2 is a nontoxic material; it has been used as a biocompatible dental material to make pigments.5255 Due to all of those reasons, ZrO2 is an excellent candidate for developing novel biolabeling and bioimaging platforms. In this work, ZrO2:Yb3+-Er3+ nanocrystals were chemically conjugated with an antiKi-67 protein by a novel method using (3-aminopropyl)triethoxysilane (APTES) and conjugated Biotin molecules as ligands. To the best of our knowledge, there are no reports about the use of luminescent ZrO2:Yb3+, Er3+ nanocrystals to label HeLa cells. Furthermore, the effect of the ligands on the luminescent properties of these nanoparticles was studied. In addition, the internalization of the conjugated nanoparticles in HeLa cells was followed by looking at their strong red luminescence using two-photon confocal microscopy. The results show the successful uptake of conjugated ZrO2:Yb3+-Er3+ nanoparticles in HeLa cells. We envision that this is a promising method for labeling different types of cancer cells for biosensing and bioimaging purposes.

2.

Experimental

2.1.

Preparation of ZrO2:Yb3+-Er3+ Nanoparticles

ZrO2:Yb3+-Er3+ nanoparticles were prepared following a precipitation process previously reported with some modifications.56 ZrOCl2·8H2O and YbCl3·6H2O (99.9%) were purchased from Aldrich, and ErCl3·6H2O (99.99%) was acquired from RE Acton. Ammonium hydroxide (NH4OH) at 30 vol% was supplied by Karal. In a typical experiment, Yb3+-Er3+ co-doped ZrO2 with a molar ratio of 21 for Yb3+:Er3+ was prepared by dissolving 2.6333 g of ZrOCl2, 0.2362 g of YbCl3·6H2O (2 mol % of Yb2O3), and 0.1351 g of ErCl3·6H2O (1 mol % of Er2O3) in 50 ml of a mixture of H2O/EtOH (11 wt%). After 15 min under stirring, the nonionic surfactant Pluronic F127 was introduced in the mixture at a molar ratio of F127/ZrO2=0.0082. Afterward, 30 ml of NH4OH was added to precipitate the salts. The resulted suspensions were transferred into a sealed autoclave and a hydrothermal treatment was carried out at 80°C for 12 h. After this, the autoclave was allowed to cool down for 30 min and the solutions were washed twice with absolute ethanol and water in a centrifuge at 4000 rpm for 10 min. Subsequently, the powders were put in a ceramic crucible and dried at 80°C for 12 h. Finally, all samples were annealed at 1000°C using a heating rate of 5°C/min.

2.2.

Conjugation and Functionalization of ZrO2:Yb3+-Er3+ Nanoparticles

The conjugation of ZrO2:Yb3+-Er3+ nanoparticles with the Ki-67 protein was carried out by following a previously reported method with some modifications.15 This process was performed as follows: 0.1 g of Yb3+-Er3+ doped ZrO2 nanoparticles were stirred with 490μl of APTES for 24 hours. This bifunctional compound has amine- and alkoxysilane groups. The alkoxysilane reacts with the OH moieties on the ZrO2:Yb3+-Er3+ nanoparticles surface, leaving the amino groups exposed for further functionalization. The samples were washed once with ethanol and water to eliminate the excess of residues and centrifuged at 6000 rpm for 10 min. The samples were dried at 40°C for 12 h. The Yb3+-Er3+ co-doped ZrO2 nanoparticles coated with APTES were dispersed in 670μl of phosphate buffered saline (PBS) 1× (pH=7.4) and then 200μl of 1500 Biotin-Anti-rabbit (mouse IgG) from BIOCARE was added to the suspension to bind the carboxylic acid groups of the IgG with the amino groups exposed in the nanoparticles; this suspension was kept at 4°C for 12 h. After that, the nanoparticles were washed with distilled water and centrifuged at 6000 rpm for 10 min to remove the supernatant. The conjugated material was kept at 37°C for 12 h. Subsequently, 300μl of PBS 1× and 10μl of antigen Ki-67-rabbit antibody from BIOCARE were added to the nanoparticles and stored for another 12 h at 4°C. Finally, the conjugated ZrO2:Yb3+-Er3+ nanoparticles were washed with distilled water and centrifuged at 6000 rpm for 10 min. The final material was dispersed and stored in distilled water.

2.3.

Structural Characterization (X-Ray Diffraction, Raman, HRTEM, SEM, Fourier Transform Infrared)

X-ray diffraction (XRD) patterns were obtained using a SIEMENS D-5005 equipment using a Cu tube with Kα radiation at 1.5405 Å, scanning in the 20 to 80 deg 2θ range with increments of 0.02 deg and a sweep time of 2 s. Raman patterns were obtained using a Renishaw Raman System (inVia Raman Microscope), which uses a 785-nm laser and a 50× objective. The nanoparticles were suspended in isopropyl alcohol at room temperature and dispersed with ultrasonication. Afterward, the solution of nanoparticles was dropped on 3-mm-diameter lacey carbon copper grids to obtain the HRTEM images in an FEI Titan 80-300 with an accelerating voltage set to 300 kV. In addition, the nanoparticle micrographs were obtained by an SEM Hitachi SU8010 at 30.0 kV. The Fourier transform infrared (FTIR) spectra were obtained using a Perkin-Elmer spectrophotometer with a deuterated triglycine sulfate detector and a spectral resolution of 4cm1. The samples were prepared using the KBr pellet method and the spectra were obtained in the range of 1000 to 4000cm1.

2.4.

Photoluminescence Characterization

Photoluminescence characterization was performed using a continuous wave semiconductor laser diode with an excitation power of 350 mW and centered at 970 nm. The luminescence emission was analyzed with a Spectrograph Spectra Pro 2300i and a R955 photomultiplier tube from Hamamatsu. The system was PC controlled with Spectra Sense software. The samples were supported in 1 mm capillary tubes in order to guarantee the same quantity of excited material. Special care was taken to maintain the alignment of the setup in order to compare the intensities between different characterized samples. All measurements were performed at room temperature.

2.5.

Incubation and Confocal Microscopy

HeLa cells were grown at a density of 5×104cells/mL in six-well culture plates with coverslips at the bottom and incubated in 3 mL of RPMI-1640 cell media for 24 h at 37°C under 5% CO2. After this, the cell media was replaced by 3 ml of ZrO2:Yb3+-Er3+ nanoparticles, ZrO2:Yb3+-Er3+-APTES, and ZrO2:Yb3+-Er3+-APTES-Biotin-Anti-rabbit/rabbit antibody-AntiKi-67 with a concentration of 100μg/mL and was incubated for 6 h. Finally, the cell-plated coverslips corresponding to each sample were washed twice with PBS buffer (1 mM, pH 7.4) and stained with nuclei-staining NucBlue® Live solution for 15 min. All the cell-plated coverslips were fixed with a solution of 4% formaldehyde. The fixed and stained coverslips were placed in microscope slides and analyzed under a two-photon Olympus FV1000 MPE SIM laser scanning confocal microscope.

2.6.

Zeta Potential and Dynamic Light Scattering Measurements

Dynamic light scattering (DLS) and zeta potential measurements were carried out using a Malvern Instrument Zetasizer Nano (red laser 633 nm). The samples were dispersed in PBS (1 mM, pH=7.4) with a concentration of 1mg/mL. The DLS and zeta potential were analyzed at 25°C.

3.

Results and Discussion

3.1.

Crystalline Structure and Morphology

The XRD pattern of the ZrO2:Yb3+-Er3+ nanopowder is shown in Fig. 1(a). This plot shows peaks corresponding to (1,0,1), (0,1,1), (2,1,1), and (1,1,2) planes, respectively. All the peaks are associated with the tetragonal phase of zirconia, according to the JCPDS 37-1413 card.57 The ZrO2:Yb3+-Er3+ nanopowder obtained by the precipitation method was analyzed by Raman spectroscopy, see Fig. 1(b). The peaks at 626, 552, 525, 445, 336, 260, 238, and 185cm1 represent the spectrum. The peaks located at 445 and 626cm1 as well as the shoulders located at 185 and 260cm1 are in agreement with the tetragonal phase of zirconia.58 The nanocrystal sizes were determined by TEM and a representative micrograph is presented in Fig. 2(a). The nanocrystals have an average size of 20 nm and spherical shape. Besides, Fig. 2(b) is an SEM image, which shows well-dispersed nanocrystals, and this was caused by the introduction of PF127 during the synthesis process.59 The size and dispersion of the co-doped ZrO2:Yb3+-Er3+ nanocrystals was controlled from the nucleation process due to the presence of ammonia, water/ethanol, and surfactant Pluronic PF127.56 To promote the efficient internalization in HeLa cells, it is important to have particles in the nanoscale size regime. In addition, the colloidal stability of the nanoparticles is also significant to avoid the formation of aggregates, which may prevent the effective interaction between the nanoparticles and the cell surface.

Fig. 1

Structural characterization of ZrO2:Yb3+-Er3+ nanocrystals: (a) x-ray diffraction and (b) Raman spectroscopy, using a laser of 785 nm.

JBO_20_4_046006_f001.png

Fig. 2

(a) HRTEM image and (b) SEM image of ZrO2:Yb3+-Er3+ nanoparticles.

JBO_20_4_046006_f002.png

3.2.

FTIR, Zeta-Potential, and DLS

Figure 3 shows the FTIR spectra of ZrO2:Yb3+-Er3+ nanocrystals and ZrO2:Yb3+-Er3+ prepared with APTES, Biotin-Anti-rabbit (mouse IgG), and rabbit antibody-AntiKi-67, respectively. These spectra provide information regarding functional groups and impurities on the surface of nanoparticles. They also corroborated that the process of functionalization and conjugation was successfully achieved. Figure 3(a) shows the FTIR spectra of nonfunctionalized nanoparticles ZrO2:Yb3+-Er3+. It depicts small peaks associated with OH groups in the range of 3200 to 3600cm1. Moreover, a peak is also observed at 450cm1, which is related to the ZrO stretching vibrations.60 The spectrum in Fig. 3(b) shows a broadening of the bands centered at 3600 and 564cm1 due to the presence of SiOH and SiOSi bonds, respectively.61,62 Other peaks at 2923 and 2351cm1 are related to CH bonds and CO2 impurities, respectively. The band located in the range of 3000 to 3400cm1 is associated with amine groups.63,64 The CO2 impurities adsorbed in the surface of the nanoparticles can come from the synthesis and/or the environment during the measurement process, which was probably caused by the granular characteristic of the nanopowder. The OH groups were introduced during the hydrolysis and condensation process where the M-OH (M=Zr, Er, and Yb) bond was formed due to the excess of hydroxyls in solution. According to the FTIR spectra in Figs. 3(a) and 3(b), the contamination produced by those hydroxyl groups is very low. Biotin-Anti-rabbit (mouse IgG) protein is conjugated to the nanoparticles containing APTES by forming an amide bond between the free amino groups located at the surface of ZrO2:Yb3+-Er3+/APTES and the carboxylic acid groups exposed in the IgG protein. The ZrO2:Yb3+-Er3+/APTES/Biotin FTIR spectrum is shown in Fig. 3(c); the spectrum illustrates a new band associated with the amide bond at 1770cm1. Moreover, a peak centered at 658cm1 is also associated with Biotin according to literature.65,66 These data further confirm the functionalization of nanoparticles. The next step is to analyze the process of conjugation with the antigen Ki-67-rabbit antibody (ZrO2:Yb3+-Er3+/APTES/Biotin/AntiKi-67), see Fig. 3(d). The bands related to Biotin are still observed, and there is a general decrease of the peaks related to impurities such as CO2 and OH radicals. However, it is observed that there is a widening of the 658cm1 band when AntiKi-67 is added. Based on this information, it is expected that AntiKi-67 is readily available to interact with HeLa cells.

Fig. 3

Fourier transform infrared spectra of (a) ZrO2:Yb3+-Er3+, (b) ZrO2:Yb3+-Er3+/(3-aminopropyl)triethoxysilane (APTES), (c) ZrO2:Yb3+-Er3+/APTES/Biotin, and (d) ZrO2:Yb3+-Er3+/APTES/Biotin/AntiKi67.

JBO_20_4_046006_f003.png

The conjugation of the nanoparticles was also analyzed by zeta potential and DLS measurements, see Table 1. The zeta potential changed from negative to positive when the ZrO2:Yb3+-Er3+ nanoparticles surface is modified with APTES, which is an indication that the amino groups are covering the nanoparticles surface. Moreover, the zeta potential was shifted from positive to negative after Biotin and antigen Ki-67 proteins were chemically attached to the surface of the material, suggesting the presence of carboxylate groups.67,68 A value of 36mV obtained in ZrO2:Yb3+-Er3+/APTES/Biotin/AntiKi-67 also indicates that nanoparticles can be stable in PBS due to their high electrostatic repulsion, which is suitable for bioapplications. A high negative value also suggests a high adsorption of nanoparticles on the nucleus of HeLa cells.69 Moreover, DLS measurements showed that the hydrodynamic diameter of the nanoparticles increased when the different molecules were added; the average sizes for ZrO2:Yb3+-Er3+, ZrO2:Yb3+-Er3+/APTES, and ZrO2:Yb3+-Er3+/APTES/Biotin/AntiKi-67 were 748, 1232, and 4694 nm, see Table 1. The size of the nanoparticles in ZrO2:Yb3+-Er3+ does not coincide with the one measured with TEM, probably due to the agglomeration of nanoparticles that were dispersed in PBS.

Table 1

Nanoparticle characterization using dynamic light scattering (DLS) and zeta potential for ZrO2:Yb3+-Er3+, ZrO2:Yb3+-Er3+-(3-aminopropyl)triethoxysilane (APTES), and ZrO2:Yb3+-Er3+-APTES-Biotin-AntiKi67 nanocrystals.

ZrO2:Yb3+-Er3+ZrO2:Yb3+-Er3+-APTESZrO2:Yb3+-Er3+-APTES-Biotin-AntiKi67
DLS (d, nm)74812324694
Zeta potential (mV)36.8+10.836.0

Figure 4 shows a schematic representation for the functionalization and conjugation of the nanoparticles. The OH moieties produced after ZrO2:Yb3+-Er3+ synthesis react with the alkoxysilane groups of APTES to afford a silica shell on the nanoparticles leaving the amine groups exposed on the surface of the material. In the next step of the reaction, the Biotin-Anti-rabbit (mouse IgG) molecule is conjugated to the amino groups by using the COOH moieties of the IgG. At this point, the Anti-rabbit can interact with the biomolecule AntiKi-67-rabbit antibody.

Fig. 4

Schematic representation of the functionalization and conjugation of ZrO2:Yb3+-Er3+ nanoparticles.

JBO_20_4_046006_f004.png

3.3.

Luminescent Properties

The mechanism of UPC emission in Er3+-Yb3+ co-doped ZrO2 is well established in the literature.70 Figure 5 shows a strong red emission band with peaks at 653 and 657 nm as well as a weak green band after excitation at 970 nm. Green and red emission bands are assigned to H211/2+S43/2I415/2 and F9/24I415/2 transitions of the Er3+ ion, and they are caused by the successive absorption of two photons after energy transfer from Yb3+ ions.59 According to previous works, the emission is predominantly red because OH groups have a vibrational energy (3000 to 4000cm1) which produces nonradiative relaxations from the mixed level H211/2+S43/2 toward the F9/24 level.71 In our case, the presence of OH moieties in all samples is corroborated by the FTIR spectra in Fig. 3. The inset in Fig. 5 shows that the integrated emission corresponds to the red band. Moreover, it is observed that the red emission of the samples ZrO2:Yb3+-Er3+/APTES (Z-A) and ZrO2:Yb3+-Er3+/APTES/Biotin (Z-A-B) decreases progressively with respect to the sample of reference without conjugated (ZrO2:Yb3+-Er3+) (Z). Nevertheless, the emission was improved when the nanoparticles were conjugated with AntiKi-67 (APTES-Biotin-AntiKi-67) (Z-A-B-K). The integrated red emission diminished with the addition of APTES and Biotin molecules because other contaminants, such as CO2, CH, and amine groups, appeared and the presence of hydroxyls increased. These elements may act as quenching centers of luminescence and also create defects, which behave as traps for luminescence.71,72 It is important to point out that the sample with APTES-Biotin had the highest levels of impurities [see Fig. 3(c)]; therefore, it showed the lowest luminescence. In contrast, the sample with APTES-Biotin-AntiKi-67 had the lowest amount of contaminants (OH, CO2, and CH) [see Fig. 3(d)]; therefore, it presented the highest red emission, see Fig. 5.

Fig. 5

Photoluminescence spectra of ZrO2:Yb3+-Er3+, ZrO2:Yb3+-Er3+/APTES, ZrO2:Yb3+-Er3+/APTES/Biotin, and ZrO2:Yb3+-Er3+/APTES/Biotin/AntiKi67. Inset shows the integrated emission of the samples.

JBO_20_4_046006_f005.png

3.4.

Imaging of ZrO2:Yb-Er Nanocrystals Incubated in HeLa Cells

Figure 6 shows the images obtained by the two-photon confocal microscope after HeLa cells were incubated with the different materials synthesized in this work. Figure 6(a) shows that the ZrO2:Yb3+-Er3+ nanoparticles are situated out of the cell, probably due to the negative charge on the surface of the nanoparticles, which limits the internalization in HeLa cells. Figure 6(b) depicts ZrO2:Yb3+-Er3+/APTES nanoparticles located on the cytoplasm of HeLa cells; these nanoparticles have no AntiKi-67, but they have APTES on their surface; this indicates that the positive charge on the surface of the nanoparticles enhances the internalization in HeLa cells. Figure 6(c) is an image of HeLa cells with nanoparticles conjugated with AntiKi-67; it is observed that 6 h of incubation is sufficient to reach the cytoplasm of HeLa cells. It is observed that there are a greater number of particles within the cell and near the nucleus. The most accepted theory is that nanoparticles are internalized via endosome-mediated transport or through ribosome exchanges.33 In general, the red emission from nanoparticles is strong in all images, which demonstrates the efficient luminescence generated by the nanoparticles synthesized in this work. It is worth noting that there was no autofluorescence from the cells after exciting the UPC nanoparticles with 970 nm. Furthermore, these images denote different sizes of emission points; this is probably induced by the nanoparticle agglomeration.

Fig. 6

Confocal microscopy images of (a) ZrO2:Yb3+-Er3+, (b) ZrO2:Yb3+-Er3+/APTES, (c) ZrO2:Yb3+-Er3+/APTES/Biotin/AntiKi67 nanoparticles after 6 h of incubation in HeLa cells.

JBO_20_4_046006_f006.png

Compared to other methods for the conjugation of nanoparticles, our technique avoids the use of other elements, such as carbon and citrate, which are relatively toxic.32,34 Moreover, it uses biomolecules (antigen and antibody) to lead our nanoparticle toward a targeted organelle; to the best of our knowledge, this kind of molecule has not been used on luminescent nanoparticles. Finally, further research is needed not only to improve the distribution and internalization of nanoparticles, but also to label specific organelles inside the HeLa cells. Those studies are in progress, and they will be presented in a subsequent work.

4.

Conclusions

In summary, we conjugated Yb3+-Er3+ co-doped ZrO2 nanoparticles using Biotin-Anti-rabbit (mouse IgG) and rabbit antibody-AntiKi-67 biomolecules. The successful conjugation was confirmed by FTIR, zeta potential, and DLS. The nanoparticles internalized in HeLa cells demonstrated a strong red luminescence and were observed using a two-photon confocal microscope. The photoluminescence spectra indicated that the UPC red emission of Er3+ ions is affected by the molecules located on the nanocrystals’ surface. An enhancement of the red emission was obtained in the nanoparticles with the conjugation with AntiKi-67. This was mainly caused by an enormous reduction of impurities compared to the rest of the samples. Our results indicate that the method of conjugation depicted in this work can be a promising alternative to afford stable colloidal dispersions of nanoparticles in water and efficiently label cancer cells.

Acknowledgments

We acknowledge the financial support from CONACyT through research grant 134111 and a PhD scholarship for Andrea Ceja. Dr. J.V.-E. thanks the start-up support from UNC-Charlotte. We also acknowledge Dr. Richard Jew for critical reading of the manuscript and helpful suggestions.

References

1. 

X. M. Li, F. Zhang and D. Y. Zhao, “Highly efficient lanthanide upconverting nanomaterials: progresses and challenges,” Nano Today, 8 (6), 643 –676 (2013). http://dx.doi.org/10.1016/j.nantod.2013.11.003 1748-0132 Google Scholar

2. 

F. Wang and X. G. Liu, “Multicolor tuning of lanthanide-doped nanoparticles by single wavelength excitation,” Acc. Chem. Res., 47 (4), 1378 –1385 (2014). http://dx.doi.org/10.1021/ar5000067 ACHRE4 0001-4842 Google Scholar

3. 

L. C. Ong et al., “Bacterial imaging with photostable upconversion fluorescent nanoparticles,” Biomaterials, 35 (9), 2987 –2998 (2014). http://dx.doi.org/10.1016/j.biomaterials.2013.12.060 BIMADU 0142-9612 Google Scholar

4. 

F. Chen et al., “Functionalized upconversion nanoparticles: versatile nanoplatforms for translational research,” Curr. Mol. Med., 13 (10), 1613 –1632 (2013). http://dx.doi.org/10.2174/1566524013666131111122133 CMMUBP 1566-5240 Google Scholar

5. 

Y. H. Chien et al., “Near-infrared light photocontrolled targeting, bioimaging, and chemotherapy with caged upconversion nanoparticles in vitro and in vivo,” ACS Nano, 7 (10), 8516 –8528 (2013). http://dx.doi.org/10.1021/nn402399m 1936-0851 Google Scholar

6. 

B. G. Yust, F. J. Pedraza and D. K. Sardar, “Ultrasmall lanthanide-doped nanoparticles as multimodal platforms,” Proc. SPIE, 8956 895608 (2014). http://dx.doi.org/10.1117/12.2037389 PSISDG 0277-786X Google Scholar

7. 

D. Y. Li et al., Multifunctional Y2O3:Yb3+/Tm3+/Li+ Nanocrystals with Enhanced Near-Infrared to Near-Infrared Upconversion Photoluminescence, Trans Tech Publications Ltd, Stafa-Zurich (2011). Google Scholar

8. 

Z. Liu et al., “Near-infrared lanthanide luminescence for functional materials,” Prog. Chem., 23 (1), 153 –164 (2011). Google Scholar

9. 

T. Liu et al., “Rare-earth upconversion nanophosphors,” Prog. Chem., 24 (2–3), 304 –317 (2012). Google Scholar

10. 

C. F. Gainer and M. Romanowski, “A review of synthetic methods for the production of upconverting lanthanide nanoparticles,” J. Innov. Opt. Health Sci., 7 (2), 11 (2014). http://dx.doi.org/10.1142/S1793545813300073 JIOHAA 1793-7205 Google Scholar

11. 

B. C. Hwang et al., “Cooperative upconversion and energy transfer of new high Er3+- and Yb3+-Er3+-doped phosphate glasses,” J. Opt. Soc. Am. B Opt. Phys., 17 (5), 833 –839 (2000). http://dx.doi.org/10.1364/JOSAB.17.000833 JOBPDE 0740-3224 Google Scholar

12. 

R. S. Meltzer et al., “Effect of the matrix on the radiative lifetimes of rare earth doped nanoparticles embedded in matrices,” J. Lumin., 94 217 –220 (2001). http://dx.doi.org/10.1016/S0022-2313(01)00281-2 JLUMA8 0022-2313 Google Scholar

13. 

T. Zako et al., “Development of near infrared-fluorescent nanophosphors and applications for cancer diagnosis and therapy,” J. Nanomater., 2010 491471 (2010). http://dx.doi.org/10.1155/2010/491471 1932-7447 Google Scholar

14. 

J. Aldana, Y. A. Wang and X. G. Peng, “Photochemical instability of CdSe nanocrystals coated by hydrophilic thiols,” J. Am. Chem. Soc., 123 (36), 8844 –8850 (2001). http://dx.doi.org/10.1021/ja016424q JACSAT 0002-7863 Google Scholar

15. 

P. L. A. M. Corstjens et al., “Infrared up-converting phosphors for bioassays,” IEE Proc. Nanobiotechnol., 152 (2), 64 –72 (2005). http://dx.doi.org/10.1049/ip-nbt:20045014 IPNEAY 1478-1581 Google Scholar

16. 

H. Kwon et al., “Synthesis and characterization of poly(vinylpyrrolidone)-capped silicon-nanoparticles,” J. Nanosci. Nanotechnol., 14 (8), 5983 –5987 (2014). http://dx.doi.org/10.1166/jnn.2014.8434 JNNOAR 1533-4880 Google Scholar

17. 

C. R. Wade et al., “Postsynthetic tuning of hydrophilicity in pyrazolate MOFs to modulate water adsorption properties,” Energy Environ. Sci., 6 (7), 2172 –2177 (2013). http://dx.doi.org/10.1039/c3ee40876k EESNBY 1754-5706 Google Scholar

18. 

K. Wei, X. M. Peng and F. Zou, “Folate-decorated PEG-PLGA nanoparticles with silica shells for capecitabine controlled and targeted delivery,” Int. J. Pharm., 464 (1–2), 225 –233 (2014). http://dx.doi.org/10.1016/j.ijpharm.2013.12.047 IJPHDE 0378-5173 Google Scholar

19. 

Q. B. Zhang et al., “NaYF4:Yb3+, Er3+ upconverting nanoparticles surface ligand exchange in ternary mixture solvent and optical properties,” Chem. J. Chin. Univ., 35 (2), 224 –229 (2014). http://dx.doi.org/10.7503/cjcu20130779 CRCUED 1005-9040 Google Scholar

20. 

C. M. M. Amaral et al., “MDM2 polymorphism associated with the development of cervical lesions in women infected with human papillomavirus and using of oral contraceptives,” Infect. Agent Cancer, 9 24 (2014). http://dx.doi.org/10.1186/1750-9378-9-24 IACNB4 1750-9378 Google Scholar

21. 

G. B. Park et al., “Antibody ligation of CM1 on cisplatin-exposed Hela cells induces apoptosis through reactive oxygen species-dependent Fas ligand expression,” Int. J. Oncol., 44 (6), 2016 –2024 (2014). http://dx.doi.org/10.3892/ijo.2014.2361 IJONES 1019-6439 Google Scholar

22. 

F. Guerra et al., “Argentophilic nucleolus organizer region as a proliferation marker in cervical intraepithelial neoplasia grade 1 of the uterine cervix,” J. Obstet. Gynaecol. Res., 40 (6), 1717 –1724 (2014). http://dx.doi.org/10.1111/jog.12380 JOGRFD 1447-0756 Google Scholar

23. 

K. K. L. Samaga et al., “Synthetic racemates of abyssinone I and II induces apoptosis through mitochondrial pathway in human cervix carcinoma cells,” Bioorg. Chem., 56 54 –61 (2014). http://dx.doi.org/10.1016/j.bioorg.2014.06.004 BOCMBM 0045-2068 Google Scholar

24. 

J. Bullwinkel et al., “Ki-67 protein is associated with ribosomal RNA transcription in quiescent and proliferating cells,” J. Cell. Physiol., 206 (3), 624 –635 (2006). http://dx.doi.org/10.1002/(ISSN)1097-4652 JCLLAX 0021-9541 Google Scholar

25. 

J. Gerdes et al., “Cell-cycle analysis of a cell proliferation-associated human nuclear antigen defined by the monoclonal-antibody Ki-67,” J. Immunol., 133 (4), 1710 –1715 (1984). JOIMA3 0022-1767 Google Scholar

26. 

J. Gerdes et al., “Production of a mouse monoclonal-antibody reactive with a human nuclear antigen associated with cell-proliferation,” Int. J. Cancer, 31 (1), 13 –20 (1983). http://dx.doi.org/10.1002/(ISSN)1097-0215 IJCNAW 1097-0215 Google Scholar

27. 

J. Bulten et al., “MIB1, a promising marker for the classification of cervical intraepithelial neoplasia,” J. Pathol., 178 (3), 268 –273 (1996). http://dx.doi.org/10.1002/(SICI)1096-9896(199603)178:3<268::AID-PATH482>3.0.CO;2-4 Google Scholar

28. 

D. W. Chan et al., “Over-expression of FOXM1 transcription factor is associated with cervical cancer progression and pathogenesis,” J. Pathol., 215 (3), 245 –252 (2008). http://dx.doi.org/10.1002/path.2355 Google Scholar

29. 

A. J. Kruse et al., “Ki-67 immunoquantitation in cervical intraepithelial neoplasia (CIN): a sensitive marker for grading,” J. Pathol., 193 (1), 48 –54 (2001). http://dx.doi.org/10.1002/1096-9896(2000)9999:9999%3C::AID-PATH719%3E3.3.CO;2-5 Google Scholar

30. 

S. Lobato et al., “Minichromosome maintenance 7 protein is a reliable biological marker for human cervical progressive disease,” J. Gynecol. Oncol., 23 (1), 11 –15 (2012). http://dx.doi.org/10.3802/jgo.2012.23.1.11 Google Scholar

31. 

M. Takagi et al., “Ki67 antigen contributes to the timely accumulation of protein phosphatase 1 gamma on anaphase chromosomes,” J. Biol. Chem., 289 (33), 22877 –22887 (2014). http://dx.doi.org/10.1074/jbc.M114.556647 JBCHA3 0021-9258 Google Scholar

32. 

T. Y. Cao et al., “Water-soluble NaYF4:Yb/Er upconversion nanophosphors: synthesis, characteristics and application in bioimaging,” Inorg. Chem. Commun., 13 (3), 392 –394 (2010). http://dx.doi.org/10.1016/j.inoche.2009.12.031 ICCOFP 1387-7003 Google Scholar

33. 

B. Sikora et al., “Transport of NaYF4:Er3+, Yb3+ up-converting nanoparticles into HeLa cells,” Nanotechnology, 24 (23), 11 (2013). http://dx.doi.org/10.1088/0957-4484/24/23/235702 NNOTER 0957-4484 Google Scholar

34. 

S. K. Bhunia et al., “Carbon nanoparticle-based fluorescent bioimaging probes,” Sci. Rep., 3 1473 (2013). http://dx.doi.org/10.1038/srep01473 SRCEC3 2045-2322 Google Scholar

35. 

Y.-F. Wang et al., “Rare-earth nanoparticles with enhanced upconversion emission and suppressed rare-earth-ion leakage,” Chem., 18 (18), 5558 –5564 (2012). http://dx.doi.org/10.1002/chem.201103485 Google Scholar

36. 

W. Yin et al., “Design of multifunctional alkali ion doped CaF2 upconversion nanoparticles for simultaneous bioimaging and therapy,” Dalton Trans., 43 (10), 3870 (2014). http://dx.doi.org/10.1039/c3dt52815d DTARAF 1477-9226 Google Scholar

37. 

J. C. Zhou et al., “Bioimaging and toxicity assessments of near-infrared upconversion luminescent NaYF4:Yb, Tm nanocrystals,” Biomaterials, 32 (34), 9059 –9067 (2011). http://dx.doi.org/10.1016/j.biomaterials.2011.08.038 BIMADU 0142-9612 Google Scholar

38. 

X. Wang et al., “One-step solvothermal synthesis of targetable optomagnetic upconversion nanoparticles for in vivo bimodal imaging,” Anal. Chem., 85 (21), 10225 –10231 (2013). http://dx.doi.org/10.1021/ac401934p ANCHAM 0003-2700 Google Scholar

39. 

S. Shan et al., “Preparation and biological application of rare earth upconversion fluorescent nanomaterials,” Prog. Biochem. Biophys., 40 (10), 925 –934 (2013). PIBBEV 0898-6711 Google Scholar

40. 

A. Patra et al., “Upconversion in Er3+:ZrO2 nanocrystals,” J. Phys. Chem. B, 106 (8), 1909 –1912 (2002). http://dx.doi.org/10.1021/jp013576z JPCBFK 1520-6106 Google Scholar

41. 

A. Patra et al., “Blue upconversion emission of Tm3+-Yb3+ in Zro2 nanocrystals: role of Yb3+ ions,” Chem. Phys. Lett., 407 (4–6), 477 –481 (2005). http://dx.doi.org/10.1016/j.cplett.2005.03.135 CHPLBC 0009-2614 Google Scholar

42. 

A. Speghini et al., “Preparation, structural characterization, and luminescence properties of Eu3+-doped nanocrystalline ZrO2,” J. Mater. Res., 20 (10), 2780 –2791 (2005). http://dx.doi.org/10.1557/JMR.2005.0358 JMREEE 0884-2914 Google Scholar

43. 

F. Gu et al., “Effect of Dy3+ doping and calcination on the luminescence of ZrO2 nanoparticles,” Chem. Phys. Lett., 380 (1–2), 185 –189 (2003). http://dx.doi.org/10.1016/j.cplett.2003.09.011 CHPLBC 0009-2614 Google Scholar

44. 

G. Y. Chen et al., “Upconversion mechanism for two-color emission in rare-earth-ion-doped Zro2 nanocrystals,” Phys. Rev. B, 75 (19), 195204 (2007). http://dx.doi.org/10.1103/PhysRevB.75.195204 PRBMDO 1098-0121 Google Scholar

45. 

P. Salas et al., “High temperature thermoluminescence induced on UV-irradiated tetragonal ZrO2 prepared by sol-gel,” Mater. Lett., 45 (5), 241 –245 (2000). http://dx.doi.org/10.1016/S0167-577X(00)00112-9 MLETDJ 0167-577X Google Scholar

46. 

M. Stoia, P. Barvinschi and F. Barvinschi, “Structural and morphologic characterization of zirconia-silica nanocomposites prepared by a modified sol-gel method,” J. Cryst. Growth, 401 462 –468 (2014). http://dx.doi.org/10.1016/j.jcrysgro.2014.02.008 JCRGAE 0022-0248 Google Scholar

47. 

P. Ghosh and A. Patra, “Role of surface coating in ZrO2/Eu3+ nanocrystals,” Langmuir, 22 (14), 6321 –6327 (2006). http://dx.doi.org/10.1021/la0604883 LANGD5 0743-7463 Google Scholar

48. 

L. A. Gomez et al., “Upconversion in Er3+-doped ZrO2 nanocrystals pumped at 1.426μm,” J. Appl. Phys., 103 (5), 4 (2008). http://dx.doi.org/10.1063/1.2844468 JAPIAU 0021-8979 Google Scholar

49. 

A. Martinez-Hernandez et al., “Synthesis and cathodoluminescence characterization of ZrO2:Er3+ films,” J. Lumin., 153 140 –143 (2014). http://dx.doi.org/10.1016/j.jlumin.2014.03.013 JLUMA8 0022-2313 Google Scholar

50. 

H. Torabmostaedi et al., “Process control for the synthesis of ZrO2 nanoparticles using FSP at high production rate,” Powder Technol., 246 419 –433 (2013). http://dx.doi.org/10.1016/j.powtec.2013.05.006 POTEBX 0032-5910 Google Scholar

51. 

V. H. Romero et al., “Brilliant blue, green and orange-red emission band on Tm3+-, Tb3+- and Eu3+-doped ZrO2 nanocrystals,” J. Phys. D Appl. Phys., 43 (46), 465105 (2010). http://dx.doi.org/10.1088/0022-3727/43/46/465105 JPAPBE 0022-3727 Google Scholar

52. 

T. J. Brunner et al., “In vitro cytotoxicity of oxide nanoparticles: comparison to asbestos, silica, and the effect of particle solubility,” Environ. Sci. Technol., 40 (14), 4381 (2006). http://dx.doi.org/10.1021/es052069i ESTHAG 0013-936X Google Scholar

53. 

G. Karunakaran et al., “Impact of nano and bulk ZrO2, TiO2 particles on soil nutrient contents and PGPR,” J. Nanosci. Nanotechnol., 13 (1), 678 –685 (2013). http://dx.doi.org/10.1166/jnn.2013.6880 JNNOAR 1533-4880 Google Scholar

54. 

S. Kanchana and S. Hussain, “Zirconia a bio-inert implant material,” IOSR J. Dent. Med. Sci., 12 (6), 66 –67 (2013). Google Scholar

55. 

S. P. C. Sam, V. S. Prasad and K. S. Kumar, “Novel nontoxic nanopigments based on zirconia and rare earth mixed oxides,” Cryst. Res. Technol., 45 (11), 1203 (2010). http://dx.doi.org/10.1002/crat.201000469 CRTEDF 0232-1300 Google Scholar

56. 

T. Lopez-Luke et al., “Solvent and surfactant effect on the self-assembly and luminescence properties of ZrO2:Eu3+ nanoparticles,” Appl. Phys. B Lasers Opt., 102 (3), 641 –649 (2011). http://dx.doi.org/10.1007/s00340-010-4233-1 APBOEM 0946-2171 Google Scholar

57. 

W. Cordova-Martinez et al., “Nanocrystalline tetragonal zirconium oxide stabilization at low temperatures by using rare earth ions: Sm3+ and Tb3+,” Opt. Mater., 20 (4), 263 –271 (2002). http://dx.doi.org/10.1016/S0925-3467(02)00067-8 OMATET 0925-3467 Google Scholar

58. 

B.-K. Kim, J.-W. Hahn and K. Han, “Quantitative phase analysis in tetragonal-rich tetragonal/monoclinic two phase zirconia by Raman spectroscopy,” J. Mater. Sci. Lett., 16 (8), 669 –671 (1997). http://dx.doi.org/10.1023/A:1018587821260 Google Scholar

59. 

T. Lopez-Luke et al., “Enhancing the up-conversion emission of ZrO2:Er3+ nanocrystals prepared by a Micelle process,” J. Phys. Chem. C, 111 (45), 17110 –17117 (2007). http://dx.doi.org/10.1021/jp072957v 1932-7447 Google Scholar

60. 

J. H. Iko Hyppänen et al., “Upconversion properties of nanocrystalline ZrO2:Yb3+, Er3+ phosphors,” J. Nanomater., 2007 16391 (2007). http://dx.doi.org/10.1155/2007/16391 1932-7447 Google Scholar

61. 

L. C. D. Rajput et al., “Silica coating of polymer nanowires produced via nanoimprint lithography from femtosecond laser machined templates,” Nanotechnology, 23 (10), 105304 (2012). http://dx.doi.org/10.1088/0957-4484/23/10/105304 NNOTER 0957-4484 Google Scholar

62. 

Q. Zhang, J. Luo and Y. Wei, “A silica gel supported dual acidic ionic liquid: an efficient and recyclable heterogeneous catalyst for the one-pot synthesis of amidoalkyl naphthols,” Green Chem., 12 (12), 2254 (2010). http://dx.doi.org/10.1039/c0gc00472c 1463-9262 Google Scholar

63. 

J. Pu et al., “Fabrication of novel graphene-fullerene hybrid lubricating films based on self-assembly for MEMS applications,” Chem. Commun., 50 (4), 469 –471 (2014). http://dx.doi.org/10.1039/C3CC47486K ICCOFP 1387-7003 Google Scholar

64. 

Y. Xu et al., “A fluorescent sensor for zinc detection and removal based on core-shell functionalized Fe3O4@SiO2 nanoparticles,” J. Nanomater., 2013 178138 (2013). http://dx.doi.org/10.1155/2013/178138 1932-7447 Google Scholar

65. 

A. A. Bunaciu et al., “FT-IR spectrophotometric analysis of ascorbic acid and Biotin and their pharmaceutical formulations,” Anal. Lett., 42 (10), 1327 (2009). http://dx.doi.org/10.1080/00032710902954490 ANALBP 0003-2719 Google Scholar

66. 

G. A. Mahmud et al., “A comparative study of aminosilanes for the application of reproducible, ultralow detection of biomolecules,” (2015) http://ma.ecsdl.org/content/MA2012-01/45/1630.full.pdf March ). 2015). Google Scholar

67. 

Z. Li et al., “Modification of NaYF4:Yb, Er@SiO2 nanoparticles with gold nanocrystals for tunable green-to-red upconversion emissions,” J. Phys. Chem. C, 115 (8), 3296 (2011). http://dx.doi.org/10.1021/jp110603r 1932-7447 Google Scholar

68. 

B. F. Pan et al., “Effects of carbon nanotubes on photoluminescence properties of quantum dots,” J. Phys. Chem. C, 112 (4), 939 –944 (2008). http://dx.doi.org/10.1021/jp068920c 1932-7447 Google Scholar

69. 

W. Lina et al., “Surface passivation of carbon nanoparticles with branched macromolecules influences near infrared bioimaging,” Theranostics, 3 (9), 677 –686 (2013). http://dx.doi.org/10.7150/thno.6535 Google Scholar

70. 

D. Solis et al., “Surfactant effect on the upconversion emission and decay time of Zro2:Yb-Er nanocrystals,” J. Lumin., 129 (5), 449 –455 (2009). http://dx.doi.org/10.1016/j.jlumin.2008.11.015 JLUMA8 0022-2313 Google Scholar

71. 

T. Lopez-Luke et al., “Improving pure red upconversion emission of co-doped Y2O3:Yb3+-Er3+ nanocrystals with a combination of sodium sulfide and surfactant pluronic-F127,” J. Lumin., 145 292 –298 (2014). http://dx.doi.org/10.1016/j.jlumin.2013.07.012 JLUMA8 0022-2313 Google Scholar

72. 

C. Landes et al., “Photoluminescence of CdSe nanoparticles in the presence of a hole acceptor: N-butylamine,” J. Phys. Chem. B, 105 (15), 2986 (2001). http://dx.doi.org/10.1021/jp0041050 JPCBFK 1520-6106 Google Scholar

Biography

Andrea Ceja-Fdez is a PhD student at Optics Research Center (Centro de Investigaciones en Optica A.C.). Her research is focused on the synthesis and optical characterization of metal and nonmetal nanoparticles. She also works in developing SERS substrates for low-detection applications of different molecules, such as glucose, or overexpressed proteins in human cells, which can be used for biolabeling, biomaging, and cancer detection.

Tzarara López-Luke is a researcher in Optics Research Center (Centro de Investigaciones en Óptica, A.C). The research is based on synthesis of luminescent metallic oxides (Rare earth doped ZrO2, Y2O3, Y2O2S), semiconductors (CdSe, CdTe, TiO2, TiO2:N) and metal nanoparticles (Au and Ag) with special structures, morphologies and optical properties, for the study in cancer detection and therapy, low concentration of glucose detection, new sources of illumination, and solar energy conversion.

Jorge Oliva obtained his PhD in optics from the Centro de Investigaciones en Optica A.C. in 2014. His research covers the fabrication and design of hybrid light-emitting diodes and solar cells for industrial applications and lighting. He also does research to develop SERS substrates for ultralow detection of proteins. He is an expert in synthesizing rare earth doped nanophosphors.

Juan Vivero-Escoto was a Carolina Postdoctoral Program for Faculty Diversity fellow at the University of North Carolina–Chapel Hill, and he became an assistant professor of chemistry at the University of North Carolina–Charlotte in 2012. He received the Wells Fargo Faculty Excellence Award 2013 and Ralph E. Powe Junior Faculty Enhancements Award (2013–2014). He has published over 30 publications, including papers and book chapters. His research focuses on designing hybrid nanomaterials for biomedical applications.

Ana Lilia Gonzalez-Yebra received his bachelor in chemistry from the Faculty of Chemistry of the University of Guanajuato (1997), and masters (2002–2004) and doctor of medicine (2004–2006) oriented to toxicology from the Institute of Medical Sciences at University of Guanajuato. Since 2006, he is professor in the Division of Health Sciences at the University of Guanajuato. He develops research on toxicology studies focused on chemical identification of cell damage. He has five publications in international refereed journals.

Ruben A. Rodriguez Rojas received his BE degree in electronic engineering from Technological Institute of Celaya and ME degree from Guanajuato University in electronics in 1998 and 2000, respectively. He also received his PhD in optics from the Center for Research in Optics (CIO) in 2004. He is actually a researcher in the Department of Exact and Technology of the University of Guadalajara since 2004. His interests are in optical materials and nanotechnology for lighting, solar cells, and biomedical applications.

Andrea Martínez-Pérez is a graduate student in biochemical engineering at the Guadalajara University in collaboration with the Optics Research Center (Centro de Investigaciones en Optica A.C.). Her thesis work was based on synthetized luminescent oxides and their bioconjugation.

Elder de la Rosa is the director of Optics Research Center (Centro de Investigaciones en Optica A.C.). His research interests are in linear and nonlinear optical properties of advanced materials for photonic applications, synthesis and characterization of nanostructured luminescent materials (oxides, semiconductors, metals) for lighting, solar cells and biomedical applications, preparation and characterization of soft glass luminescence (P2O5, TeO2) rare earth doped lasers, and fiber amplifiers.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Andrea Ceja-Fdez, Tzarara López-Luke, Jorge Oliva, Juan Vivero-Escoto, Ana Lilia Gonzalez-Yebra, Ruben A. Rodríguez-Rojas, Andrea Martínez-Pérez, and Elder de la Rosa "Labeling of HeLa cells using ZrO2:Yb3+-Er3+ nanoparticles with upconversion emission," Journal of Biomedical Optics 20(4), 046006 (16 April 2015). https://doi.org/10.1117/1.JBO.20.4.046006
Published: 16 April 2015
Lens.org Logo
CITATIONS
Cited by 13 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Ytterbium

Nanoparticles

Nanocrystals

Luminescence

FT-IR spectroscopy

Erbium

Cancer

Back to Top