Tuning the size and upconversion emission of NaYF4:Yb3+/Pr3+ nanoparticles through Yb3+ doping

Shuwei Haoa, Wei Shaoa, Hailong Qiua, Yunfei Shanga, Rongwei Fanc, Xuyun Guod, Lili Zhaoae, Guanying Chen*ab and Chunhui Yang*a
aSchool of Chemical Engineering and Technology, Harbin Institute of Technology, 150001 Harbin, People's Republic of China. E-mail: yangchh@hit.edu.cn
bInstitute for Lasers, Photonics and Biophotonics, Department of Chemistry, University at Buffalo, The State University of New York, Buffalo, NY 14260, USA. E-mail: guanying@buffalo.edu
cNational Key Laboratory of Tunable Lasers, Institute of Optical-Electronics, Harbin Institute of Technology, 150001 Harbin, People's Republic of China
dDepartment of Chemistry, the Hong Kong University of Scienece and Technology, Clear water Bay, Hong Kong SPA
eHarbin Huigong Technology Co., Ltd., People's Republic of China

Received 27th September 2014 , Accepted 23rd October 2014

First published on 23rd October 2014


Abstract

We introduce a simple method to tune the resulting size as well as the upconversion luminescence of NaYF4:Yb3+/Pr3+ nanoparticles through varying the sensitizer ytterbium concentration. Various amounts of ytterbium from 10–70% were doped into the fluoride nanoparticles, producing a tunable size from 29 to 153 nm. Meanwhile, the blue upconversion luminescence intensity of NaYF4:Yb3+/Pr3+ nanoparticles was monotonously enhanced, reaching a maximum ∼3.4 fold enhancement at an ytterbium concentration of 70% owing to an improved energy transfer from the ytterbium to the praseodymium ions. Moreover, the luminescence intensity ratio of the blue to the green upconversion was tailored by the composition-dependent cross relaxation process. The result here provides a paradigm for simultaneous control of the physical dimensions as well as the luminescence properties of lanthanide-doped upconversion nanoparticles.


Introduction

Lanthanide-doped upconversion nanoparticles (UCNPs) have attracted a great deal of consideration due to their intriguing optical properties that promise their potential applications in fields as diverse as solid-state lasers,1 optical data storage,2 solar energy conversion,3–6 biological imaging,7–10 and photodynamic therapy.11,12 UCNPs are able to convert two or more long wavelength light photons into short wavelength emissions through the use of energy level of trivalent lanthanide ion that is embedded into an inorganic host lattice. In particular, lanthanide doped fluoride UCNPs generally exhibit the highest upconversion (UC) or downconversion (DC) efficiency,3 as the fluoride materials have high physicochemical stabilities, as well as intrinsic low phonon energies (<350 cm−1) that are able to minimize energy losses at the intermediate states of lanthanide ions. Typically, UCNPs of NaYF4 doped with a sensitizer Yb3+ and an activator of Er3+, Ho3+ or Tm3+ has been of extensive study, and considered to be one of the most efficient UC systems.13 This is not only because the sensitizer Yb3+ has a unique one excited energy 2F5/2 (∼10[thin space (1/6-em)]000 cm−1) which possesses a significantly higher extinction coefficient (typically ∼10 l mol−1 cm−1) and which matches the ladder-like energy gaps of the Er3+, Ho3+ or Tm3+ ions to empower efficient resonant energy transfers to produce efficient UC emissions.14–18 However, limited success has been met in NaYF4-based UCNPs doped with other lanthanide ions like Pr3+ ion that has a unique pattern of energy levels.

The Pr3+ ion has been of great use as multi-wavelength laser activator or colored emitting ions, as it offers the possibility to achieve a simultaneous blue, green or red emission lasing or other purposes.19 Visible UC lasers at multiple wavelengths have been realized in Pr3+ doped ZBLAN (Zr–Ba–La–Al–Na) bulk glass under an infrared laser pump at ∼840 nm.20 Laser action of the blue UC from the transition of the excited 3P0 state to the ground 3H4 state has been demonstrated in various bulk crystals, involving the frequency conversion mechanisms of either excited state absorption or photon avalanche.1,21–23 Moreover, yellow-to-blue frequency UC processes in Pr3+ ions have also been extensively studied in a range of low phonon crystals and glass,24,25 while intense white light emitting were developed in Pr3+/Er3+ co-doped tellurite glass sensitized by Yb3+ ion.27 Despite successes on Pr3+-doped bulk materials, till this point, limited investigation has been devoted to Pr3+-doped nanoparticles. The relative work in literature is on visible UC in Yb3+/Pr3+ co-doped Y2O3 nanoparticles.26 However, the employed low Yb3+ concentration of 1–2 mol% produces a low UC efficiency owing to the large Yb3+–Pr3+ ion–ion distance that limits the efficiency of energy transfer from the Yb3+ to the Pr3+. It is noted that the sensitizer Yb3+ ion has an exclusive excited state; increase of its concentration, therefore, exclude the possibility producing cross relaxation process induced quenching effect. Varying Yb3+ concentration to shorten the sensitizer–activator distance is an important approach to enhance UC luminescence; this conclusion has been verified by Duan et al., Han et al. as well as by our group in Yb3+/Tm3+-codoped NaYF4 and YF3 UCNPs.28–34 On the other hand, doping of trivalent lanthanide ions into nanomaterials can influence its growth dynamics due to the dopant-induced transient electric dipole on the surface of the growing nanoparticles.35–37 For example, our group showed that varying dopant concentration of Yb3+ or Gd3+ can tune the size and phase of the resulting multifunctional CeO2 oxide nanoparticles.37 Here, we show that varying the sensitizer Yb3+ concentration in Yb3+/Pr3+-codoped NaYF4 nanoparticles not only can produce a size-tunable UNCPs, but also can produce enhanced and tailored upconversion emissions from the Pr3+ ion under ∼980 nm NIR light excitation.

Results and discussion

The ionic radii of Yb3+ (0.868 Å) and Y3+ (0.893 Å) are different, which may result in the variation of size and morphology in the formation of NaREF4 UCNPs. We prepared NaYF4:Yb3+/Pr3+ UCNPs doped with varied Yb3+ ion concentrations of 10–70 mol%. During the synthesis, all synthetic parameters were kept exactly the same, except for varying the doping concentration of Yb3+ ions. Fig. 1(a–d) displays the transmission electron microscopy (TEM) images of NaYF4 UCNPs doped with 10, 30, 50, 70 mol% Yb3+, respectively. High resolution TEM imaging and selected area electron diffraction patterns confirmed the high crystallinity and hexagonal phase of all resulting nanoparticles (Fig. S1, ESI). The energy dispersive X-ray spectra (EDAX) indicate that the elemental content of the resulting NaYF4:Yb3+/Pr3+ UCNPs is in general agreement with the mixture of cationic precursor, suggesting stoichiometric doping of Yb3+ ions at a precisely defined concentration (Fig. S2, ESI). The results demonstrate that all the rare earth ions can be effectively incorporated into the final UCNPs with our synthetic conditions. Therefore, the Yb3+ and Pr3+ doping concentration can be readily controlled by varying the initial reactant ratio.
image file: c4ra11357h-f1.tif
Fig. 1 Transmission electron images for NaYF4 nanocrystals codoped with 0.5% Pr3+ and various concentration of (a) 10% Yb3+, (b) 30% Yb3+, (c) 50% Yb3+, and (d) 70% Yb3+. (e) Schematic representation of the hexagonal structure of NaYF4 or NaYbF4 nanoparticles (a = b = 6.245 Å, c = 4.392 Å for NaYF4 and a = b = 5.912 Å, c = 4.653 Å for NaYbF4).

As one can see in Fig. 1(a), very small ∼29 nm sphere-like UCNPs were formed at Yb3+ concentration of 10%. However, it evolves into larger ∼62 nm hexagonal-shape UCNPs when Yb3+ concentration of 30% were doped [Fig. 1(b)]. For higher Yb3+ concentration of 50 and 70%, the hexagonal shape remains unchanged, yet the size increased further to 103 and 153 nm, respectively [Fig. 1(c and d)]. These results agree with the corresponding histogram of the average size distribution (ESI, Fig. S3). The size evolution can be attributed to the Yb3+-varied crystal growth rate via the modification of electron charge density on the nanoparticle surface. Liu et al. calculated, based on density functional theory (DFT), that the electron charge density of the crystal surface was increased when a Y3+ (0.893 Å) ion in NaYF4 was replaced by a larger Gd3+ (0.938 Å) ion, which thus repelled the anion F to produce a smaller nanoparticle size.35 Therefore, when Y3+ (0.893 Å) is replaced by a smaller Yb3+ (0.868 Å) ion replacing, the electron charge density on the surface of growing UCNPs will be decreased, thus allowing more attraction of F ion to the particle surface to form a larger size UCNPs.

The crystalline phases of as-prepared samples were further determined by powder X-ray diffraction (XRD); the result is shown in Fig. 2. The XRD pattern for the sample doped with Yb3+ of 10 mol% is in good agreement with the standard hexagonal phase NaYF4 (JCPDS 16-0334), confirming the formation of a hexagonal crystal phase. A sequential shifting of the XRD peaks is observed towards high-angle when higher Yb3+ concentration were doped. This peak shifting indicates the decrease of a unit-cell volume due to the replacement of Y3+ by smaller ion radii Yb3+. In other words, the hexagonal phase NaYF4 gradually transforms to hexagonal phase NaYbF4 when elevated Yb3+ concentration was doped.


image file: c4ra11357h-f2.tif
Fig. 2 The XRD patterns of NaYF4 nanoparticles codoped with 0.5% Pr3+ and various concentration of Yb3+ ions of (a) 10% Yb3+, (b) 30% Yb3+, (c) 50% Yb3+, and (d) 70% Yb3+. The standard XRD patterns of hexagonal phase NaYF4 and NaYbF4 are also plotted as a reference.

It is interesting to note that doping of varied Yb3+ concentration also produced important influence on the UC luminescence. Fig. 3 shows the emission spectra of NaYF4 UCNPs codoped with 0.5% Pr3+ and varied Yb3+ ion concentrations of (a) 10%, (b) 30%, (c) 50%, (d) 70%, respectively, under ∼980 nm laser excitation. All UCNPs exhibit the typical blue emission band at ∼486 nm, the green emission band at ∼523 nm, the green emission band at ∼540 nm, the red emission band at ∼605 nm, as well as the red emission band at ∼639 nm, being ascribed to the 3P03H4, 1I63H5, 3P13H5, 3P03H6 and 3P03F2 transitions of Pr3+ ions, respectively. As one can see in Fig. 3, the blue emission at 486 nm is increased gradually when elevating the Yb3+ concentration, and enhances about ∼3.4 times for Yb3+ of 70%. We also compared the luminescence intensity of NaYF4:70%Yb3+/0.5%Pr3+ with that of well-investigated hexagonal NaYF4:20%Yb3+/0.5%Er3+ and NaYF4:20%Yb3+/0.5%Tm3+ nanocrystals (size ∼30 nm) (ESI, Fig. S4); it is lower than that of the typical NaYF4:20%Yb3+/0.5%Er3+ and NaYF4: 20%Yb3+/0.5%Tm3+ systems. This result indicates that a further design of this Yb3+/Pr3+ codoped system is in need to increase its upconversion efficiency to compete with currently well-known Yb3+/Er3+ and Yb3+/Tm3+ UC systems.


image file: c4ra11357h-f3.tif
Fig. 3 UC emission spectra of colloidal NaYF4 nanocrystals codoped with 0.5% Pr3+ and various concentrations of Yb3+ ions (10–70%) under diode laser excitation at 980 nm.

To shed light on the enhancement mechanism of blue UC emission as well as the population process of upper excited 3P0 levels of Pr3+ ions, the dependence of the intensities of various UC emission peaks on laser power for NaYF4:10%Yb3+/0.5%Pr3+ UCNPs were measured and displayed in Fig. 4. For an unsaturated UC process, the number of photons that are involved to populate the upper emitting state can be obtained by the relation.

 
IUCPn, (1)
where IUC is the UC intensity, P is the pump laser power, and n is the number of laser photons involved to populate the upper emitting levels. The value of n can be determined by the slope value of a linear fitting in a logarithmic–logarithmic plot of eqn (1). As shown in Fig. 4, slope values of 1.98, 1.91, 1.89, and 1.75 were observed for UC emissions peaked at 486, 523, 540, and 639 nm, respectively, for NaYF4:10%Yb3+/0.5% Pr3+ UCNPs. These slope values agree well with the previous results in Yb3+–Pr3+ codoped bulk materials under continuous wave laser excitation at ∼980 nm.38 This result illustrates that two-photon processes are involved to populate the 1I6, 3P1 and 3P0 states, respectively (consult Fig. 5). It is noted that the blue emission at 486 nm and the green emissions at 540 nm have the same slope value of ∼2 (see Fig. 5). The mechanisms for blue, green, and red UC generation are presented in Fig. 5.


image file: c4ra11357h-f4.tif
Fig. 4 log–log plots of the intensities of various UC emission bands in Fig. 3 on the excitation density in NaYF4 nanoparticles codoped with 10 mol% Yb3+ and 0.5 mol% Pr3+. The slope values of the linear fits (solid line) are presented in the inset together with the peak wavelength of each UC band.

image file: c4ra11357h-f5.tif
Fig. 5 Energy level diagram of the Pr3+ and Yb3+ ions as well as the proposed UC mechanism in NaYF4:Yb3+/Pr3+ nanocrystals via varying content of Yb3+ ions under a 980 nm laser excitation.

Fig. 5 depicts the pertinent energy levels of the Pr3+ and Yb3+ ions as well as the proposed UC mechanisms.27 Laser photons at ∼980 nm match the 2F7/22F5/2 transition of Yb3+ ion, resonantly exciting Yb3+ ion from the ground to its excited state. Energy transfer process from the Yb3+ ions to a neighboring Pr3+ ion can promote the Pr3+ ions in the 3H4 state to 1G4 state (ET1). The population in the 1G4 level can be promoted to the 3P0 levels either by energy transfer (ET2) from another excited Yb3+ ion or absorbing the energy of pumping photon. Once the 3P0 level is populated, the excited electron can release its energy by emitting visible emissions. The blue emission at 486 nm can be produced by radiative decay to the ground state from the 3P0 state. Generally, the pumping energy of 980 nm is very difficult to population in 3P1 and higher states of Pr3+ ions. But the cross relaxation between 3P03P2 and 3H53H4 in Pr3+ ions promotes the occupation in the 3P2 state, which through nonradiative relaxation process populates 1I6 and 3P1 states. Excited Pr3+ ions in 1I6 and 3P1 states decay to 3H5 state, leading to green emission centered at 523 and 540 nm. The red emissions at 605 and 639 nm originate from the 3P0 state to 3H6 and 3F2 transitions, respectively.

It is noted that the efficiency of cross relaxation processes strongly depends on the Pr3+ ion's concentration, while the efficiency of ET1 and ET2 processes are insensitive to Pr3+ ion concentration in the host lattice when varying in a small amount range.

As a result, we further prepared NaYF4 nanoparticles codoped with Yb3+ of 10%, and various Pr3+ concentrations of 0.1%, 0.25%, 0.5%, and 1% Pr3+. The UC spectra from them are displayed in Fig. 6(a). The intensity increases slightly when the Pr3+ concentration is increased from 0.1 to 0.5% molar fraction. This augment of emission intensity can be ascribed to the increased number of emitting Pr3+ centers. However, a further increase in the Pr3+ concentration results in a decrease in the emission intensity. This is partly because, for a given doping content of Yb3+ and overall excitation energy, the excitation received by individual Pr3+ ion decreased with the increase of Pr3+ ion concentration. More importantly, the high concentration of Pr3+ ions leads to a decrease in distance between Pr3+ ions, which enhances the probability of cross-relaxation, and thus suppresses the efficient energy transfer processes from Yb3+ ions to Pr3+ ions for population of 3P0 state, resulting in an increase in green/blue ratio with the increase of Pr3+ concentration [Fig. 6(b)]. It is noted that the intensity ratio of the band 540 nm to the band at 639 nm remain almost unchanged. This is because they arise from the same upper excited energy level of the 3P0 state.


image file: c4ra11357h-f6.tif
Fig. 6 (a) UC emission spectra of NaYF4:Yb3+/Pr3+ nanoparticles codoped with Yb3+ of 10%, and various concentrations of Pr3+ ions (0.1–1%) under diode laser excitation at 980 nm. (b) The intensity ratios of emissions at 486 nm (from the 3P0 state to the 3H4 state), at 540 nm (from the 3P1 state to the 3H5 state), to the emission at 639 nm (from the 3P0 state to the 3F2 state) versus Pr3+ concentration.

Fig. 7 shows the variation of UC emissions from NaYF4:10%Yb3+/0.5%Pr3+ UCNPs when irradiated with a varied pump power. For low pump power the intensity for emission at 486 nm is lower than that of the emission at 540 nm, while a reverse result is obtained under high pump power. This switching result illustrates that high laser irradiance also facilitates the increase of blue UC emission in NaYF4:10%Yb3+/0.5%Pr3+ nanoparticles. It is reasonable to assume that, for the given Yb3+ concentration in the samples, the excitation received by individual Pr3+ ions increased with the enhancement of laser irradiance.


image file: c4ra11357h-f7.tif
Fig. 7 (a) Pump-power-dependent UC 4f emission spectra of NaYF4:10%Yb3+/0.5%Pr3+ UCNPs. All spectra were recorded at room temperature under excitation of ∼980 nm diode laser at a power density of 5 W cm−2. (b) Emission intensities for the 3P03H4 transition and the 3P13H5 transition from NaYF4: 10%Yb3+/0.5%Pr3+ UCNPs as a function of laser pump power.

Conclusions

In summary, we have demonstrated that varying the sensitizer Yb3+ concentration can simultaneously tune the resulting size as well as the upconversion luminescence of NaYF4:Yb3+/Pr3+ UNCPs. The intensity of the blue UC emission was enhanced by ∼3.4 fold when elevating the sensitizer concentration from 10 to 70%, being ascribed to the improved energy transfer from the Yb3+ to Pr3+ ions due to the shorter distance between them. The intensity ratio of the blue to the green upconversion was also varied by the composition-dependent cross relaxation process. The reported result might have important implications in other lanthanide-doped UCNPs that involve the use of Yb3+ as sensitizer ions.

Acknowledgements

This work is supported by Natural Science Foundation of China (51102066 and 51402071), the Fundamental Research Funds for the Central Universities (Grant no. HIT. NSRIF.2015048 and AUGA5710052614), international scientific and technological cooperation projects (Grant no. 2014DFA50740), the Program for Basic Research Excellent Talents in Harbin Institute of Technology (BRETIII 2012018) and the National Science Fund for Distinguished Young Scholars (Grant no. 51325201).

Notes and references

  1. F. Auzel, Chem. Rev., 2004, 104, 139 CrossRef CAS PubMed.
  2. E. Downing, L. Hesselink, J. Ralston and R. Macfarlane, Science, 1996, 273, 1185 CAS.
  3. W. Q. Zou, C. Visser, J. A. Maduro, M. S. Pshenichnikov and J. C. Hummelen, Nat. Photonics, 2012, 6, 561 CrossRef.
  4. C. Z. Yuan, G. Y. Chen, P. N. Prasad, T. Y. Ohulchanskyy, Z. J. Ning, H. Tian, L. C. Sund and H. Agren, J. Mater. Chem., 2012, 22, 16709 RSC.
  5. G. Y. Chen, J. W. Seo, C. H. Yang and P. N. Prasad, Chem. Soc. Rev., 2013, 42, 8304 RSC.
  6. Y. N. Qian, R. Wang, B. Wang, B. F. Zhanga and S. P. Gao, RSC Adv., 2014, 4, 6652 RSC.
  7. J. Zhou, Z. Liu and F. Li, Chem. Soc. Rev., 2012, 41, 1323 RSC.
  8. T. S. Yang, Q. Liu, J. C. Li, S. Z. Pu, P. Y. Yang and F. Y. Li, RSC Adv., 2014, 4, 15613 RSC.
  9. G. Y. Chen, C. H. Yang and P. N. Prasad, Acc. Chem. Res., 2013, 46, 1474 CrossRef CAS PubMed.
  10. S. W. Hao, G. Y. Chen and C. H. Yang, Theranostics, 2013, 3, 331 CrossRef PubMed.
  11. N. Muhammad Idris, M. K. Gnanasammandhan, J. Zhang, P. Ho, R. Mahendran and Y. Zhang, Nat. Med., 2013, 18, 1580 CrossRef PubMed.
  12. G. Y. Chen, H. L. Qiu, P. N. Prasad and X. Y. Chen, Chem. Rev., 2014, 114, 5161 CrossRef CAS PubMed.
  13. F. Wang and X. G. Liu Chem, Chem. Soc. Rev., 2009, 38, 976 RSC.
  14. M. Haase and H. Schäfer, Angew. Chem., Int. Ed., 2012, 50, 5808 CrossRef PubMed.
  15. S. Heer, K. Kompe, H. U. Gudel and M. Haase, Adv. Mater., 2004, 16, 2102 CrossRef CAS.
  16. C. Li, J. Yang, Z. Quan, P. Yang, D. Kong and J. Lin, Chem. Mater., 2007, 19, 4933 CrossRef CAS.
  17. D. Q. Chen and P. Huang, Dalton Trans., 2014, 43, 11299 RSC.
  18. D. Q. Chen, Y. Chen, H. W. Lu and Z. G. Ji, Inorg. Chem., 2014, 53, 8638 CrossRef CAS PubMed.
  19. R. Piramidowicz, R. Mahiou, P. Boutinaud and M. Malinowski, Appl. Phys. B, 2011, 104, 873 CrossRef CAS.
  20. H. M. Pask and A. C. Hanna, Opt. Commun., 1997, 134, 139 CrossRef CAS.
  21. L. D. Merkle, B. Zandi, Y. Guyot, H. R. Verdun, B. McIntosh, B. H. T. Chai, J. B. Gruber, M. D. Seltzer, C. A. Morrison and R. Moncorge, OSA Proc. Adv. Solid-State Lasers, 1994, 20, 361 Search PubMed.
  22. L. Esterowitz, R. Allen, M. Krueger, F. Bartoli, L. S. Goldber, H. P. Jenssen, A. Linz and V. O. Nicolai, J. Appl. Phys., 1977, 48, 650 CrossRef CAS PubMed.
  23. T. Danger, T. Sandrock, E. Heumann, G. Huber and B. Chai, Appl. Phys. B, 1993, 57, 239 CrossRef.
  24. R. Balda, J. Fernandez, A. Mendioroz, M. Voda and M. Al-Saleh, Opt. Mater., 2003, 24, 91 CrossRef CAS.
  25. A. Remillieux, B. Jacquier, C. Linares, C. Lesergent, S. Artigaud, D. Bayard, L. Hamon and J. L. Beylat, J. Phys. D: Appl. Phys., 1996, 29, 963 CrossRef CAS.
  26. C. B. Zheng, Y. Q. Xia, F. Qin, Y. Yu, J. P. Miao, Z. G. Zhang and W. W. Cao, Chem. Phys. Lett., 2010, 496, 316 CrossRef CAS PubMed.
  27. Y. Dwivedi, R. Anita and S. B. Rai, J. Appl. Phys., 2008, 104, 043509 CrossRef PubMed.
  28. H. L. Qiu, G. Y. Chen, R. W. Fan, L. M. Yang, S. W. Hao, C. H. Yang and P. N. Prasad, Nanoscale, 2014, 6, 753–757 RSC.
  29. G. Y. Chen, T. Y. Ohulchanskyy, R. Kumar, H. Ågren and P. N. Prasad, ACS Nano, 2011, 4, 3163 CrossRef PubMed.
  30. H. Zhang, Y. J. Li, Y. C. Lin, Y. Huang and X. F. Duan, Nanoscale, 2011, 3, 963 RSC.
  31. G. Y. Chen, T. Y. Ohulchanskyy, W. C. Law, H. Ågren and P. N. Prasad, Nanoscale, 2011, 3, 2003 RSC.
  32. G. Y. Chen, J. Shen, T. Y. Ohulchanskyy, N. J. Patel, A. Kutikov, Z. P. Li, J. Song, R. K. Pandey, H. Ågren, P. N. Prasad and G. Han, ACS Nano, 2012, 6, 8280 CrossRef CAS PubMed.
  33. J. Shen, G. Y. Chen, T. Y. Ohulchanskyy, S. J. Kesseli, S. Buchholz, Z. P. Li, P. N. Prasad and G. Han, Small, 2013, 9, 3213 CAS.
  34. A. Punjabi, X. Wu, A. T. Apollon, M. E. Rifai, H. Lee, Y. W. Zhang, C. Wang, Z. Liu, E. M. Chan, C. Y. Duan and G. Han, ACS Nano, 2014 DOI:10.1021/nn505051d.
  35. F. Wang, Y. Han, C. S. Lim, Y. H. Lu, J. Wang, J. Xu, H. G. Chen, C. Zhang, M. H. Hong and X. G. Liu, Nature, 2010, 463, 1061 CrossRef CAS PubMed.
  36. D. Chen and Y. Wang, Nanoscale, 2013, 5, 4621–4637 RSC.
  37. H. Qiu, G. Chen, R. Fan, C. Cheng, S. Hao, D. Chen and C. Yang, Chem. Commun., 2011, 47, 9648–9650 RSC.
  38. R. Balda, J. Fernandez, A. Mendioroz, M. Voda and M. Al-Saleh, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 68, 165101 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra11357h

This journal is © The Royal Society of Chemistry 2014