Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Facile synthesis of Gd2O3:Ho3+/Yb3+ nanoparticles: an efficient upconverting material for enhanced photovoltaic performance of dye-sensitized solar cells

Open Access Open Access

Abstract

A series of Ho3+/Yb3+-codoped Gd2O3 nanoparticles were prepared by a facile urea-based homogeneous precipitation method. Under the excitation of 980 nm light, all the samples exhibited strong green and red upconversion (UC) emissions corresponding to the (5F4,5S2) → 5I8 and 5F55I8 transitions of Ho3+ ions, respectively. A gradual enhancement in the UC emission intensity was observed with increasing the Yb3+ ion concentration, achieving its optimum value when the doping concentration was 3 mol%. In addition, with the introduction of Ho3+/Yb3+-codoped Gd2O3 nanoparticles into the TiO2 porous film of dye-sensitized solar cells (DSSCs), the power conversion efficiency of the cells (7.403%) was ~10.47% higher than that of the DSSCs with pure TiO2 porous film (6.701%), which is mainly caused by increased short-circuit current density due to their enhanced light-harvesting properties via an efficient UC process. The result suggested that the incorporation of upconverting nanoparticles into the TiO2 porous film is an effective approach to improve the photovoltaic characteristics of DSSCs.

© 2016 Optical Society of America

1. Introduction

Dye-sensitized solar cells (DSSCs), which were first reported by O’Regan and Grätzel in 1991, are thought to be a promising candidate to replace commercial silicon solar cells because of their low cost of fabrication, eco-friendly feature and simple structure [1–3]. Typically, DSSCs are made up of a dye-sensitized TiO2 porous membrane as the photoanode, an electrolyte solution containing the I-/I3- redox couples and the platinum-coated fluorine tin oxide (FTO) glass as the counter electrode. As is known, the photovoltaic properties of DSSCs can be greatly influenced by the light harvesting ability of the dye [4,5]. However, the commonly used dyes, such as N3, N-719 and N-749, can only absorb the light in the visible range (400-800 nm) of the total solar irradiation. This means that nearly 50% of the sun light energy in the ultraviolet (~7%) and infrared (~43%) regions has not been harvested by these dyes, resulting in a relatively low efficiency of DSSCs. Nowadays, the introduction of luminescent materials (especially, upconverting nanoparticles) into the TiO2 porous film is considered to be an alternative method to improve the light harvesting ability of DSSCs as well as the cell efficiency [6–8]. The upconversion (UC) emission is a non-linear optical process in which low-energy photons (infrared light) are absorbed, high-energy photons (visible light) that can be reabsorbed by the dyes are generated, and more electrons will be produced, leading to the increased photovoltaic performances of DSSCs. Diao et al. pointed out that the power conversion efficiency (PCE) of DSSCs was increased from 5.791 to 6.661% using the CeO2:Er3+/Yb3+ nanofiber as the upconverting layer [6]. Furthermore, it was also revealed that the Y2O3:Er3+ nanorods can be used to improve the photovoltaic performances of DSSCs [9]. Although some inspiring achievements have been obtained, they still cannot meet the requirement of practical applications and more efforts should be paid.

Over the last decades, the upconverting materials based on the rare-earth (RE) doped materials, such as fluorides, inorganic oxides and ceramics, have been intensively studied as a result of their superior luminescent properties [10–12]. Among these RE ions, holmium (Ho3+) ion is thought to be an excellent activator candidate for upconverting materials because of its unique luminescent properties. Generally, the Ho3+ ion exhibits two emissions in the green and red regions corresponding to the (5F4,5S2) → 5I8 and 5F55I8 transitions, respectively [13,14]. Furthermore, ytterbium (Yb3+) ion is usually codoped with other RE ions, such as Er3+, Tm3+, Pr3+ and Ho3+, as the sensitizer to improve the luminescent properties due to its large absorption in the infrared region and efficient energy transfer (ET) from Yb3+ to these RE ions [6,15–17]. Meanwhile, gadolinium oxide (Gd2O3), which has a cubic phase as an important part of the inorganic oxides, is regarded as a good luminescent host lattice because of its high physical durability, thermal stability and low phonon energy (as low as ~600 cm−1) [18,19]. It was reported that strong UC emissions were observed in RE ions (Er3+/Yb3+) doped Gd2O3 nanoparticles which have potential applications in optical temperature sensors and bioimaging [18,20]. However, to the best of our knowledge, the investigation on the UC emission properties of Ho3+/Yb3+-codoped Gd2O3 nanoparticles and especially its application in DSSCs is still insufficient. In this work, a series of sphere-like Ho3+/Yb3+-codoped Gd2O3 nanoparticles were prepared by a simple urea-based homogeneous precipitation method and their phase structure, morphology and luminescent properties were investigated in detail. Additionally, to explore their effect in DSSCs, the Ho3+/Yb3+-codoped Gd2O3 nanoparticles were introduced into the TiO2 porous membrane to form a hybrid photoelectrode.

2. Experimental details

The conventional urea-based homogeneous precipitation method was employed to fabricate the sphere-like Ho3+/Yb3+-codoped Gd2O3 nanoparticles (abbreviated as Gd2O3:0.01Ho3+/xYb3+, where the Ho3+ ion concentration was 1 mol% and x = 0.01, 0.02, 0.03 and 0.04). High-purity powders of Gd(NO3)3·6H2O (99.9%), Ho(NO3)3·5H2O (99.9%) and Yb(NO3)3·5H2O (99.9%) were used as the starting materials. On the basis of the stoichiometric proportion, the Gd(NO3)3·6H2O, Ho(NO3)3·5H2O and Yb(NO3)3·5H2O were weighted and dissolved in the de-ionized (DI) water (200 mL) to form a transparent solution under vigorously mechanical stirring. Subsequently, moderate urea (i.e., molar ratio of urea to RE ions: 30) was added. Then, the mixed solution was sealed in a beaker and heated at 80 °C for 4 h under magnetic stirring. After that, the white precipitate was separated by centrifugation and washed with ethanol and DI water for several times to remove the remained ions. Ultimately, the precipitate was transferred into the aluminum crucible and calcined at 800 °C for 2 h in a furnace to create Ho3+/Yb3+-codoped Gd2O3 nanoparticles.

The DSSCs were assembled via a conventional fabrication process. Briefly, the TiO2 colloids (PST-18NR) were firstly deposited on the FTO glass to form a transparent photoelectrode with the thickness of about 5 μm by the doctor-blade method and sintered at 500 °C for 2 h. Subsequently, the TiO2 colloides (PST-400C) containing 2 wt% Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles were screen-printed on the TiO2 film, forming a hybrid TiO2 porous film with a thickness of around 5 μm. Thus, the thickness of the total TiO2 film was approximately 10 μm. After that, the TiO2-coated FTO glass was soaked in the N-719 dye (3 × 10−4 M ethanol) for 24 h. For comparison, a pure TiO2 porous film without upconverting nanoparticles was also prepared. Finally, the sandwich-type DSSCs were fabricated with the platinum-coated FTO glass and an electrolyte (Dyesol, electrolyte HPD).

The X-ray diffractometer (Bruker D8 Advance) with Cu Kα radiation (λ = 0.15406 nm) was applied to check the phase structure of the samples and the MDI JADE 5.0 software was employed to deal with the data. The morphological properties of the obtained materials were characterized by using a field-emission scanning electron microscope (FE-SEM) (LEO SUPPA 55, Carl Zeiss) and a transmission electron microscope (TEM) (JEM-2100F, JEOL). The elemental mappings of the as-synthesized particles were measured by the TEM (JEM-2100F, JEOL). As for the UC measurement, the excitation source was a 980 nm light laser diode and a fluorescence spectrophotometer (Ocean Optics USB 4000) was used to obtain UC emission spectra. The photocurrent density-voltage (J-V) curves of the as-prepared DSSCs were recorded by using a photocurrent system which consists of a solar simulator with 1000 W xenon short arc lamp and a source meter (Keithley 2400). The incident photon to current conversion efficiency (IPCE) characteristics of the DSSCs were measured by utilizing a 300 W xenon arc lamp as the light source coupled to a monochromator (TLS-300x xenon light source, Newport) with an optical power meter (Newport 2935-C).

3. Results and discussion

In order to analyze the crystal structure and phase purity of the obtained phosphors, the X-ray diffraction (XRD) patterns of the Gd2O3:0.01Ho3+/xYb3+ nanoparticles were measured, as shown in Fig. 1. From Fig. 1(a), it is evident that all the diffraction peaks of the as-prepared samples can be indexed to the Ia-3 (206) space group of cubic Gd2O3 (JCPDS#12-0797) and no any other additional peaks for impurity phases were detected within the resolution of the device. This means that all the samples had pure cubic phase and the Ho3+/Yb3+ ions were diffused into the Gd2O3 host lattices. Furthermore, as shown in Fig. 1(b), the positions of the diffraction peaks slightly shifted to the larger angle with the addition of Ho3+ and Yb3+ ions. This is because the large sized Gd3+ (1.05 Å, when coordinate number (CN) is 8) ions were substituted by the small sized Yb3+ (0.99 Å, when CN is 8) and Ho3+ (1.02 Å, when CN is 8) ions, leading to a shrinkage of the lattices. To better understand the lattice shrinkage phenomenon, the lattice parameters of the as-synthesized phosphors were calculated using MDI JADE 5.0 software and the corresponding results are illustrated in Table. 1. Notably, with raising the Yb3+ ion concentration from 1 to 4 mol%, the lattice parameters, such as a, b and c, were decreased from 10.7939 to 10.7784 Å, which further confirmed the shrinkage of the lattices.

 figure: Fig. 1

Fig. 1 (a) XRD patterns of Gd2O3:0.01Ho3+/xYb3+ (x = 0.01, 0.02, 0.03 and 0.04) nanoparticles sintered at 800 °C. (b) Magnified XRD patterns in the 2θ range between 44 and 52°.

Download Full Size | PDF

Tables Icon

Table 1. Structural parameters of Gd2O3 host lattice and Gd2O3:0.01Ho3+/xYb3+ (x = 0.01, 0.02, 0.03 and 0.04) nanoparticles

The structural and morphological properties of the Ho3+/Yb3+-codoped Gd2O3 phosphors were characterized by FE-SEM and TEM images. Figure 2 shows the FE-SEM images of the products. It is clear that the as-prepared samples were made up of uniformly spherical particles with the diameter ranging from about 200 to 300 nm. Note that, all the samples exhibited the same morphology and the particle size did not change with the increase of Yb3+ ion concentration, as presented in Fig. 2, suggesting that the doping of the Yb3+ ions had little effect on the shape and size of the resultant phosphors. Furthermore, from Fig. 3, one knows that the particle size of sintered samples was slightly decreased along with the relatively rougher surface compared to the as-prepared precursor and the dehydration process can be responsible for the shrinkage of the particle [21]. The TEM image further confirmed that the obtained particles were in the nanometer range with sphere-like morphologies, as described in Fig. 4(a). Meanwhile, the high-resolution TEM (HR-TEM) image showed the clear lattice fringes and the distance between the adjacent lattice fringes was found to be about 0.33 nm which is well consistent with the d-space of the (222) plane of cubic Gd2O3 (JCPDS# 12-0797). In addition, the elemental mappings, which were taken from the scanning transmission electron microscope (STEM) image of inset of Fig. 4(b), indicated that the Gd, Ho, Yb and O were uniformly distributed over the entire particles, as illustrated in Fig. 4(c).

 figure: Fig. 2

Fig. 2 FE-SEM images of Gd2O3:0.01Ho3+/xYb3+ nanoparticles: (a) x = 0.01, (b) x = 0.02, (c) x = 0.03 and (d) x = 0.04.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 FE-SEM images of (a) as-prepared precursor (Gd(OH)CO3:0.01Ho3+/0.03Yb3+) and (b) Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles sintered at 800 °C.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 (a) TEM image and (b) HR-TEM image of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles. (c) Elemental mappings of Gd, Ho, Yb and O. Inset of (b) shows the STEM image of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles.

Download Full Size | PDF

Figure 5(a) depicts the representative room-temperature UC emission spectrum of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles under 980 nm light excitation. As shown in Fig. 5(a), the UC emission spectrum consisted of an intensive green UC emission located at about 550 nm and a weak red UC emission centered at approximately 665 nm which are ascribed to the (5F4,5S2) → 5I8 and 5F55I8 transitions of Ho3+ ions, respectively [17,22]. It is worth noting that the intensity of the green UC emission was nearly six times as high as that of the red UC emission. Hence, under 980 nm light excitation, the emission color of the samples was green, as illustrated in the inset of Fig. 5(a). It is known that the N-719 dye has a strong light absorption in the visible wavelength (300-700 nm) with a maximum absorption band located at about 530 nm [23], which matched well with the UC emissions of Ho3+/Yb3+-codoped Gd2O3 nanoparticles. Therefore, with the introduction of Ho3+/Yb3+-codoped Gd2O3 upconverting nanoparticles into the TiO2 porous membrane, the absorption range of the N-719 dye can be extended, and thus the enhanced photovoltaic performance of DSSCs is expected to be achieved.

 figure: Fig. 5

Fig. 5 (a) UC emission spectrum of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles. (b) UC emission intensity as a function of Yb3+ ion concentration. (c) Simplified energy level diagram of Ho3+ and Yb3+ ions. Inset of (a) shows the luminescence image excited by 980 nm light.

Download Full Size | PDF

The dependence of Yb3+ ion concentration on the UC emission intensity is displayed in Fig. 5(b). It can be seen that both green and red UC emission intensities showed an upward trend with the increment of Yb3+ ion concentration and its optimum value was achieved when x = 0.03 due to efficient ET from Yb3+ to Ho3+ ions. Nevertheless, the UC emission intensity began to decease when the Yb3+ ion concentration was over 3 mol%. The main reason for the deceased UC emission properties should be attributed to the energy back transfer (EBT) from Yb3+ to Ho3+ ions, i.e., (5F4,5S2) (Ho3+) + 2F7/2 (Yb3+) → 5I6 (Ho3+) + 2F5/2 (Yb3+), as shown in Fig. 5(c) [17]. Generally, the higher the Yb3+ ion concentration becomes, the higher the probability of EBT from Ho3+ to Yb3+ ions is. To better comprehend the UC mechanism in the Ho3+/Yb3+-codoped Gd2O3 nanoparticles, the simplified energy level diagram of Ho3+ and Yb3+ ions along with possible UC processes is demonstrated in Fig. 5(c). Similar as previous reports on Ho3+/Yb3+-codoped materials, the ET process from Yb3+ to Ho3+ ions is dominant [14,17,24]. Briefly, under the excitation at 980 nm, the Yb3+ ions absorb the infrared photon energy and the electrons are excited from the ground state to the excited level, i.e., (2F7/2 (Yb3+) + infrared photon → 2F5/2 (Yb3+)). Meanwhile, the energy can be transferred from Yb3+ to the adjacent Ho3+ ions (2F5/2 (Yb3+) + 5I8 (Ho3+) → 2F7/2 (Yb3+) + 5I6 (Ho3+)), leading to the population of the 5I6 level. After that, the electrons located at the 5I6 intermediary level are excited to the (5F4,5S2) level either by the absorption of an infrared photon or a second ET from Yb3+ to Ho3+ ions (2F5/2 (Yb3+) + 5I6 (Ho3+) → 2F7/2 (Yb3+) + (5F4,5S2) (Ho3+)). Then, the bright green UC emission is observed due to the radiative transition of Ho3+ ions to the ground state ((5F4,5S2) → 5I8). In comparison, two different paths are involved to generate the red UC emission. The first one is the non-radiative (NR) transition from the (5F4,5S2) level to the 5F5 level, as described in Fig. 5(c). Subsequently, the red UC emission corresponding to the 5F55I8 transition is observed. The second one is the NR transition of 5I65I7 followed by an ET process from Yb3+ to Ho3+ ions to populate the 5F5 level (2F5/2 (Yb3+) + 5I7 (Ho3+) → 2F7/2 (Yb3+) + 5F5 (Ho3+)). Finally, the red UC emission centered at about 665 nm is generated through the 5F55I8 transition.

To identify the effect of as-prepared upconverting nanospheres on the photovoltaic performance of DSSCs, the Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles were dispersed in the TiO2 porous film to fabricate a novel DSSC device. The surface morphologies of pure and hybrid TiO2 porous films are shown in Fig. 6. From the FE-SEM image of the hybrid TiO2 film, it is evident that there were some Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles attached to the surface of the TiO2 film, which can improve the light scattering properties to achieve a higher photocurrent [25]. Nevertheless, compared with that of the pure TiO2 porous film, the surface morphology of the hybrid TiO2 porous film changed a little, suggesting that the UC emission of the Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles was the main factor to affect the cell performance.

 figure: Fig. 6

Fig. 6 FE-SEM images of the surface of (a) pure TiO2 porous film and (b) hybrid TiO2 porous film with Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles.

Download Full Size | PDF

Figure 7(a) shows the J-V curves of the DSSCs with and without addition of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles and the corresponding photovoltaic parameters, such as short-circuit current density (JSC), open-circuit voltage (VOC), fill factor (FF) and PCE, are summarized in Table 2. As presented in Fig. 7(a) and Table 1, the DSSCs with pure TiO2 porous film exhibited the photovoltaic characteristics of JSC = 15.662 mA/cm2, VOC = 0.715 V, FF = 59.77% and PCE = 6.701%. In comparison, with the introduction of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles into the TiO2 porous film, the photovoltaic properties of the DSSCs were enhanced, that is, JSC = 17.091 mA/cm2, VOC = 0.715 V, FF = 60.51% and PCE = 7.403%. It is obvious that the PCE of the DSSCs was greatly improved (by an increment percentage of approximately 10.47%) with the addition of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles and the increased JSC should be responsible for the modified PCE. As we know, the increase of the JSC can be significantly affected by the incidence light wavelength range [3,9]. Here, the Gd2O3:0.01Ho3+/0.03Yb3+ upconverting nanoparticles had a strong light absorption in the near-infrared region and the energy can be transferred to the visible region by the UC process, leading to the extension of the absorption range of the DSSCs. As a result, the JSC and PCE of the DSSCs were increased. On the other hand, the IPCE, which directly shows how efficiently the incident photons are converted to the electrons, is thought to be a key parameter for solar cells. As shown in Fig. 7(b), the IPCE of the DSSCs with the addition of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles was much higher than that of the referenced DSSCs which coincided well with the result of the J-V curves, demonstrating that photocurrent was increased with the introduction of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles into the TiO2 porous film.

 figure: Fig. 7

Fig. 7 (a) J-V curves and (b) IPCE spectra of the DSSCs with and without Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles.

Download Full Size | PDF

Tables Icon

Table 2. Photovoltaic parameters of DSSCs without and with addition of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles.

4. Conclusion

In summary, high-efficiency upconverting Ho3+/Yb3+-codoped Gd2O3 nanoparticles were synthesized and incorporated into the TiO2 porous film of DSSCs. Under 980 nm light excitation, strong UC emissions were observed and the UC emission intensity increased gradually with the Yb3+ ion concentration, achieving its optimum value when x = 0.03. Furthermore, with the introduction of the Ho3+/Yb3+-codoped Gd2O3 nanoparticles into the TiO2 porous film, the DSSCs exhibited an efficiency of 7.403%, which indicates a percentage improvement of about 10.47% compared to that of the DSSCs with pure TiO2 porous film (6.701%). From these results, the Ho3+/Yb3+-codoped Gd2O3 nanoparticles provided excellent UC emission properties and were capable of enhancing the light harvest ability of N-719 dye as well as the photovoltaic performances of DSSCs.

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) Grant funded by the Korea Government (MSIP) (No. 2015R1A5A1037656 and No. 2013R1A1A2010037).

References and links

1. B. O’Regan and M. Grätzel, “A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films,” Nature 353(6346), 737–740 (1991). [CrossRef]  

2. J. H. Lim, Y. H. Ko, J. W. Leem, and J. S. Yu, “Improvement in light harvesting of dye-sensitized solar cells with antireflective and hydrophobic textile PDMS coating by facile soft imprint lithography,” Opt. Express 23(3), A169–A179 (2015). [CrossRef]   [PubMed]  

3. Z. Zhou, J. Wang, F. Nan, C. Bu, Z. Yu, W. Liu, S. Guo, H. Hu, and X. Z. Zhao, “Upconversion induced enhancement of dye sensitized solar cells based on core-shell structured β-NaYF4:Er3+, Yb3+@SiO2 nanoparticles,” Nanoscale 6(4), 2052–2055 (2014). [CrossRef]   [PubMed]  

4. P. Ramasamy and J. Kim, “Combined plasmonic and upconversion rear reflectors for efficient dye-sensitized solar cells,” Chem. Commun. (Camb.) 50(7), 879–881 (2014). [CrossRef]   [PubMed]  

5. K. Lee, S. W. Park, M. J. Ko, K. Kim, and N. G. Park, “Selective positioning of organic dyes in a mesoporous inorganic oxide film,” Nat. Mater. 8(8), 665–671 (2009). [CrossRef]   [PubMed]  

6. J. Bai, B. Zhao, G. Han, Z. Li, and G. Diao, “Synthesis of 1D upconversion CeO2:Er,Yb nanofibers via electrospinning and their performance in dye-sensitized solar cell,” RSC Advances 5(54), 43328–43333 (2015). [CrossRef]  

7. P. Zhao, Y. Zhu, X. Yang, X. Jiang, J. Shen, and C. Li, “Plasmon-enhanced efficient dye-sensitized solar cells using core-shell-structured β-NaYF4:Yb,Er@SiO2@Au nanocomposites,” J. Mater. Chem. A Mater. Energy Sustain. 2(39), 16523–16530 (2014). [CrossRef]  

8. M. Liu, Y. Liu, Z. B. Xie, and G. M. Chow, “Enhancing near-infrared solar cell response using upconverting transparent ceramics,” Sol. Energy Mater. Sol. Cells 95(2), 800–803 (2011). [CrossRef]  

9. J. Wang, J. Wu, J. Lin, M. Huang, Y. Huang, Z. Lan, Y. Xiao, G. Yue, S. Yin, and T. Sato, “Application of Y2O3:Er3+ nanorods in Dye-Sensitized Solar Cells,” ChemSusChem 5(7), 1307–1312 (2012). [CrossRef]   [PubMed]  

10. Z. Yi, S. Zeng, W. Lu, H. Wang, L. Rao, H. Liu, and J. Hao, “Synergistic Dual-Modality in Vivo upconversion Luminescence/X-ray Imaging and Tracking of Amine-Functionalized NaYbF4:Er Nanoprobes,” ACS Appl. Mater. Interfaces 6(6), 3839–3846 (2014). [CrossRef]   [PubMed]  

11. E. Pavitra, G. S. R. Raju, J. Oh, and J. S. Yu, “Pump power induced tunable upconverison emission from Er3+/Tm3+/Yb3+ ions tri-doped SrY2O4 nanocrystalline phosphors,” New J. Chem. 38(8), 3413–3420 (2014). [CrossRef]  

12. Z. Wu, Y. Zhang, G. Bai, W. Tang, J. Gao, and J. Hao, “Effect of biaxial strain induced by piezoelectric PMN-PT on the upconversion photoluminescence of BaTiO₃:Yb/Er thin films,” Opt. Express 22(23), 29014–29019 (2014). [CrossRef]   [PubMed]  

13. W. Gao, H. Zheng, Q. Han, E. He, and R. Wang, “Unusual upconversion emission from single NaYF4:Yb3+/Ho3+ microrodes under NIR excitation,” CrystEngComm 16(29), 6697–6706 (2014). [CrossRef]  

14. M. Kochanowicz, J. Zmojda, P. Miluski, J. Pisarska, W. A. Pisarski, and D. Dorosz, “NIR to visible upconversion in double-clad optical fiber co-doped with Yb3+/Ho3+,” Opt. Mater. Express 5(7), 1505–1510 (2015). [CrossRef]  

15. F. Vetrone, V. Mahalingam, and J. A. Capobianco, “Near-Infrared-to-Blue Upcnversion in Colloidal BaYF5:Tm3+,Yb3+ Nanocrystals,” Chem. Mater. 21(9), 1847–1851 (2009). [CrossRef]  

16. Y. Huang, L. Luo, J. Wang, Q. Zuo, Y. Yao, and W. Li, “The down-conversion and up-conversion photoluminescence properties of Na0.5Bi0.5TiO3:Yb3+/Pr3+ ceramics,” J. Appl. Phys. 118(4), 044101 (2015). [CrossRef]  

17. G. Ding, F. Gao, G. Wu, and D. Bao, “Bright up-conversion green photoluminescence in Ho3+-Yb3+ co-doped Bi3Ti3O12 ferroelectric thin films,” J. Appl. Phys. 109(12), 123101 (2011). [CrossRef]  

18. Y. Tian, B. Tian, C. Cui, P. Huang, L. Wang, and B. Chen, “Size-dependent upconversion luminescence and temperature sensing behavior of spherical Gd2O3:Yb3+/Er3+ phosphor,” RSC. Adv. 5, 14123–14128 (2015).

19. C. L. Lyuer, A. García-Murillo, E. Bernstein, and J. Mugnier, “Waveguide Raman spectroscopy of sol-gel Gd2O3 thin films,” J. Raman Spectrosc. 34(3), 234–239 (2003). [CrossRef]  

20. G. Yang, R. Lv, S. Gai, Y. Dai, F. He, and P. Yang, “Multifunctional SiO2@Gd2O3:Yb/Tm Hollow Capsules: Controllable Synthesis and Drug Release Properties,” Inorg. Chem. 53(20), 10917–10927 (2014). [CrossRef]   [PubMed]  

21. S. Gai, P. Yang, D. Wang, C. Li, N. Niu, F. He, and X. Li, “Monodisperse Gd2O3:Ln (Ln = Eu3+, Tb3+, Dy3+, Sm3+, Yb3+/Er3+, Yb3+/Tm3+, and Yb3+/Ho3+) nanocrystals with tunable sized and multicolor luminescent properties,” CrystEngComm 13(17), 5480–5487 (2011). [CrossRef]  

22. R. Li, H. Ji, Z. Li, J. Wang, Q. Liu, and L. Liu, “Fabrication and markedly enhanced white upconversion emission of core-shell structured NaGdF4:Tm3+/Yb3+/Ho3+@SiO2,” New J. Chem. 38(2), 611–615 (2014). [CrossRef]  

23. G. S. Han, Y. H. Song, D. H. Kim, M. J. Lee, D. G. Lee, S. H. Han, Y. Kim, M. K. Jung, D. H. Yoon, and H. S. Jung, “Green-emitting Lu3Al5O12:Ce3+ Phosphor as a visible light amplifier for dye-sensitized solar cells,” RSC. Adv. 5, 24737–24741 (2015).

24. Y. Qu, R. Wang, Z. Qiu, Y. Tao, and J. Zhou, “White-light upconversion emissions and color tenability of 12CaO·7Al2O3:Ho3+/Yb3+/Tm3+ single crystal,” Opt. Mater. Express 5(8), 1881–1889 (2015). [CrossRef]  

25. G. Zhu, H. Wang, Q. Zhang, and L. Zhang, “Enhanced photovoltaic performance of dye-sensitized solar cells based on NaYF4:Yb3+, Er3+-incorporated nanocrystalline TiO2 electrodes,” J. Colloid Interface Sci. 451, 15–20 (2015). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 (a) XRD patterns of Gd2O3:0.01Ho3+/xYb3+ (x = 0.01, 0.02, 0.03 and 0.04) nanoparticles sintered at 800 °C. (b) Magnified XRD patterns in the 2θ range between 44 and 52°.
Fig. 2
Fig. 2 FE-SEM images of Gd2O3:0.01Ho3+/xYb3+ nanoparticles: (a) x = 0.01, (b) x = 0.02, (c) x = 0.03 and (d) x = 0.04.
Fig. 3
Fig. 3 FE-SEM images of (a) as-prepared precursor (Gd(OH)CO3:0.01Ho3+/0.03Yb3+) and (b) Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles sintered at 800 °C.
Fig. 4
Fig. 4 (a) TEM image and (b) HR-TEM image of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles. (c) Elemental mappings of Gd, Ho, Yb and O. Inset of (b) shows the STEM image of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles.
Fig. 5
Fig. 5 (a) UC emission spectrum of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles. (b) UC emission intensity as a function of Yb3+ ion concentration. (c) Simplified energy level diagram of Ho3+ and Yb3+ ions. Inset of (a) shows the luminescence image excited by 980 nm light.
Fig. 6
Fig. 6 FE-SEM images of the surface of (a) pure TiO2 porous film and (b) hybrid TiO2 porous film with Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles.
Fig. 7
Fig. 7 (a) J-V curves and (b) IPCE spectra of the DSSCs with and without Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles.

Tables (2)

Tables Icon

Table 1 Structural parameters of Gd2O3 host lattice and Gd2O3:0.01Ho3+/xYb3+ (x = 0.01, 0.02, 0.03 and 0.04) nanoparticles

Tables Icon

Table 2 Photovoltaic parameters of DSSCs without and with addition of Gd2O3:0.01Ho3+/0.03Yb3+ nanoparticles.

Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.