Tunable band gaps of Co3−xCuxO4 nanorods with various Cu doping concentrations

Chun-Te Ho , Tsu-Heng Weng , Chiu-Yen Wang , Shiang-Jie Yen and Tri-Rung Yew *
Department of Materials Science and Engineering, National Tsing-Hua University, Hsinchu, Taiwan 30013. E-mail: tryew@mx.nthu.edu.tw; Fax: +886-3-5722366; Tel: +886-936347230

Received 19th February 2014 , Accepted 17th April 2014

First published on 18th April 2014


Abstract

Co3−xCuxO4 nanorods with a crystalline structure were synthesized via a solution process in this study. The Co3−xCuxO4 nanorods exhibit dual direct bandgaps and could be adjusted to 1.3–1.5 eV, and 2.55–3.5 eV by using different Cu-doping concentrations. The Co3−xCuxO4 nanorods show an absorption coefficient of ∼104 cm−1 which increases with the increase of the Cu doping concentration.


Introduction

Cobalt oxide (Co3O4) based materials have been widely studied because of their potential applications in many technological fields. In particular, their applications in gas sensors,1,2 electrochromics devices,3 heterogeneous catalysts,4 high-temperature solar selective absorbers,5 electrocatalytic oxygen evolution reaction (OER),6 supercapacitors,7 and lithium batteries2,8,9 have been extensively investigated. These applications are mainly attributed to the high surface area and better electrochemical reactivity of nanostructured Co3O4.

Besides, there has been an increasing interest in developing metal-doped Co3O4, since the Co3O4 properties could be adjusted by the substitution of cobalt ions with foreign divalent metal ions in spinel structure. For example, Li- and Au-doped Co3O4 nanowires have been synthesized and exhibited a high storage capacity for lithium batteries.10 There are a few studies focusing on Cu-doped Co3O4 recently.11 The enhancement of Cu-doped Co3O4 in catalytic activity, OER,11 and the sensitivity of methane gas sensor12 were reported in these studies.

In addition, to estimate the electronic structure and band gap value before and after doping metal, theoretical calculation has been made in previous literatures.13–15

In this study, efforts have been made to synthesize Co3−xCuxO4 nanorods by low temperature hydrothermal process. The morphology, structure and optical property with various Cu doping concentrations were characterized. In addition, the growth of Co3−xCuxO4 nanorods on conductive substrates have also been investigated for the future application on self-assembled electrical devices.

Experimental

Cu-doped Co3O4 nanostructure preparation9

Co(NO3)2·6H2O (0.1 M) and Cu(CH3COO)2·3H2O (0.02 M, 0.01 M, and 0 M) were loaded into a 500 ml reactor respectively, which was prefilled with 34 ml NH4OH and 66 ml distilled (DI) water. The reactor was put in a reflux system and remained at 85 °C for 12 h, and then cooled down to room temperature. The reaction temperature and time have been optimized, which were presented in following paragraph. The final products were centrifuged and rinsed with distilled water and ethanol for several times to remove chemical residuals. After rinsing, the product was annealed in the furnace at 400 °C for 2 h to ensure that other chemical residuals were completely eliminated.

Cu-doped Co3O4 nanorods optimization

The optimization of Co3O4 nanorods at various reaction temperatures and concentrations of reduction agent (ammonia), while all using 10% molar ratio of Cu(CH3COO)2 in Co(NO3)2, were also investigated in this work as shown in Fig. 1. It could be observed that different processing temperatures (60, 85, and 120 °C) result in different surface morphology (Fig. 1a–c). With processing temperature at 60 °C, the Cu-doped Co3O4 nanoparticles with ∼50 nm in diameter were observed (Fig. 1a). When the processing temperature increases to 85 °C, the Cu-doped Co3O4 nanorods were observed (Fig. 1b). For the synthesis temperature up to 120 °C, the Cu-doped Co3O4 particles with a diameter of 600–1000 nm were observed in Fig. 1c. The synthesis using various NH4OH concentrations for 12 h at 85 °C result in different morphologies as shown in Fig. 1d–f. Fig. 1d presents the Cu-doped Co3O4 nanoparticles with a diameter of 200–400 nm using H2O to NH4OH molar ratio 4[thin space (1/6-em)]:[thin space (1/6-em)]1. As the H2O to NH4OH molar ratio was decreased to 2[thin space (1/6-em)]:[thin space (1/6-em)]1, the Cu-doped Co3O4 nanorods were synthesized (Fig. 1e). When the H2O to NH4OH molar ratio decreases to 1[thin space (1/6-em)]:[thin space (1/6-em)]1, the Cu-doped Co3O4 nanorods with slightly rough surface were observed (Fig. 1f). As the nanorods morphology could provide larger surface area light collection, the optimized condition was determined to be at 85 °C and H2O to NH4OH molar ratio 2[thin space (1/6-em)]:[thin space (1/6-em)]1.
image file: c4ra01463d-f1.tif
Fig. 1 The SEM images of Cu-doped Co3O4 nanorods grown on Pt/glass substrates at the growth temperature of (a) 60 °C, (b) 85 °C, and (c) 120 °C with H2O–NH4OH = 2[thin space (1/6-em)]:[thin space (1/6-em)]1 for 12 h. The samples were prepared using (d) H2O–NH4OH = 4[thin space (1/6-em)]:[thin space (1/6-em)]1, (e) H2O–NH4OH = 2[thin space (1/6-em)]:[thin space (1/6-em)]1, and (f) H2O–NH4OH = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 at a process temperature of 85 °C for 12 h. All above Cu-doped Co3O4 samples were, synthesized using 10% molar ratio of Cu(CH3COO)2 in Co(NO3)2.

Materials characterization

The morphology and structure of the products were characterized by field emission scanning electron microscopy (FESEM, JEOL-JSM 6500F), high resolution transmission electron microscopy (HRTEM, JEOL-JSM 2010) and X-ray diffraction spectroscopy (Shimadzu XRD 6000). The compositions of the materials were analyzed by the inductively coupled plasma mass spectrometry (ICP-MS, PerkinElmer). Raman spectra (HORIBA HR 800, laser excitation λ = 633 nm) were acquired to characterize the structural properties. The chemical composition of Cu-doped CO3O4 nanorods was characterized by X-ray photoelectron spectroscopy (XPS, PHI Quantera SXM). The optical properties were measured by an ultraviolet-visible near-infrared spectrophotometer (UV-Vis-NIR, Hitachi U-4100) over the range of 400–2000 nm, and a crygenic cathodoluminescence system (CL, Gatan MonoCL) at room temperature (∼300 K) with an excitation beam with a 30 kV accelerating voltage and ∼18 nA beam current.

Results and discussion

Morphology and structure analysis

The Cu-doped Co3O4 nanorods were synthesized by ammonia-evaporation-induced growth.9 The glass substrates with and without a Pt layer were immersed in an aqueous solution with Co(NO3)2, and different amounts of Cu(CH3COO)2 were added into the solution to adjust the Cu-doping level in Co3O4 nanorods. The effects of process temperatures and NH4OH concentrations on the morphology of Cu-doped Co3O4 nanorods were investigated. The results in Fig. 1 show the optimum synthesis condition at H2O–NH4OH = 2[thin space (1/6-em)]:[thin space (1/6-em)]1 and a process temperature of 85 °C for 12 h, which will be used for further study. It could be observed that different concentrations of Cu(CH3COO)2 resulted in different Co3O4 surface morphologies as shown in the scanning electron microscopy (SEM) images of Fig. 2a–c. The pristine Co3O4 nanorods with about 200–500 nm in diameter, and 5–10 μm in length can be observed in SEM image (Fig. 2a). With the increase of the molar ratio of Cu(CH3COO)2 in Co(NO3)2 from 10% to 20%, the corresponding diameter of Cu-doped Co3O4 nanorods increases from 300–600 nm to 500–800 nm as observed in the SEM images of Fig. 2b and c, respectively.
image file: c4ra01463d-f2.tif
Fig. 2 The SEM images of Co3−xCuxO4 nanorods grown on Pt/glass substrates (a) pristine Co3O4, (b) Co2.98Cu0.02O4, and (c) Co2.94Cu0.06O4.

The content of Cu in Co3O4 nanorods, synthesized by using 10% molar ratio of Cu(CH3COO)2 in Co(NO3)2, were characterized by inductively coupled plasma mass spectrometry (ICP-MS spectrometer). The signal of Cu element (0.28 ± 0.01 at%) was measured in Co3O4 nanorods, which was calculated to be a compound of Co2.98Cu0.02O4. The content of Cu in Co3O4 nanorods, synthesized by using 20% molar ratio of Cu(CH3COO)2 in Co(NO3)2, were also characterized showing a Cu composition of 0.92 ± 0.02 at%, i.e., a compound of Co2.94Cu0.06O4.

The crystallographic properties of the Co2.98Cu0.02O4 nanorods were examined by transmission electron microscopy (TEM), high-resolution TEM (HRTEM), and selected-area electron diffraction (SAED) as shown in Fig. 3a–d. The bright-field TEM image in Fig. 3a reveals an individual Co2.98Cu0.02O4 nanorod with a diameter of 200 nm. Fig. 3b presents its dark field image contributed from the (111) diffraction of the Co2.98Cu0.02O4. Fig. 3c shows the corresponding diffraction pattern with the dotted ring-shaped patterns that indicate the polycrystalline structure of Co2.98Cu0.02O4 nanorods. The d-spacings of 0.477 and 0.245 nm as calculated from the HRTEM lattice image of the Co2.98Cu0.02O4 nanorods (Fig. 3d) correspond to the spacing of 0.465 nm (111) and 0.243 nm (311), respectively, for the cubic spinel Co3O4 (a = 8.084 nm) according to X-ray JCPDS file no. 653103. It is interesting to observe the d-spacing is slightly larger than that of the JCPDS file reference, which suggests the doping of Cu atoms into Co3O4 nanorods lattice which leads to the expansion of Co3O4 lattice due to the larger size of Cu atom than Co.


image file: c4ra01463d-f3.tif
Fig. 3 The transmission electron microscopy (TEM) analysis of Co2.98Cu0.02O4. The (a) bright field image, (b) dark field TEM image contributed from (111) diffraction, (c) the corresponding diffraction pattern taken from (a), and (d) the high resolution TEM lattice image taken from the indicated boxed area of (c).

The crystal structure of Co3O4 nanorods with different amounts of Cu doping was also confirmed by X-ray diffraction spectroscopy (XRD), as shown in Fig. 4a. It shows that for pristine Co3O4 nanorods all peaks can be indexed to cubic spinel phase Co3O4 based on the standard data file (JCPDS file no. 653103). It is interesting to observe that the peaks shift to smaller-angles for both Co2.98Cu0.02O4 and Co2.94Cu0.06O4 compared to pristine Co3O4. The peak shift indicates that the lattice is enlarged after Cu-doping in Co3O4, which is mostly due to the radius of Cu atom is larger than that of Co. The results are consistent with the lattice enlargement observed in the previous lattice image of HRTEM (Fig. 3d).


image file: c4ra01463d-f4.tif
Fig. 4 (a) XRD patterns of pristine Co3O4, Co2.98Cu0.02O4 and Co2.94Cu0.06O4 nanorods and (b) their Raman spectra.

The structure properties of pristine Co3O4, Co2.98Cu0.02O4 and Co2.94Cu0.06O4 were also analyzed by Raman spectra as shown in Fig. 4b. The Co3O4 is belonged to spinel structure, Co2+(Co3+)2(O2−)4, with space group O7h. Co3+ and Co2+ placed at octahedral and tetrahedral sites of Co3O4, respectively. The vibrational modes of Co3O4 at k = 0 in irreducible representations of the factor group O7h are

 
Γ = A1g + Eg + 3F2g + 5F1u + 2A2u + 2Eu + 2F2u(1)

The A1g, Eg and the three F2g modes are Raman active. The F1g, 2A2u, 2Eu and 2F2u modes are inactive. In the five F1u modes, one is an acoustic mode and the other four are infrared active.16,17 In Raman spectra, the blueshift increases with the increase of Cu doping concentration in Co3O4 nanorods. The blueshift is believed to originate from the stress attributed to the lattice expansion.18,19 It is again consistent with the observation of lattice image HRTEM. The corresponding Raman data and the comparison of their characteristics peaks are summarized in Table 1.

Table 1 The summary of Raman analyses
  F2g (cm−1) Eg (cm−1) F2g (cm−1) F2g′ (cm−1) A1g (cm−1)
Pristine Co3O4 (ref. 2) 194 482 521 618 691
Pristine Co3O4 194 477 520 616 686
Co2.98Cu0.02O4 191 470 510 608 672
Co2.94Cu0.06O4 189 466 509 606 670


The purity and composition of Cu-doped Co3O4 nanorods were also analyzed using X-ray photoelectron spectroscopy (XPS), and the results show that the Cu has been doped into Co3O4 successfully, including Cu+ and Cu2+, shown in Fig. 5 and Table 2.


image file: c4ra01463d-f5.tif
Fig. 5 The XPS analyses show (a) the wide scan energy spectra of pristine Co3O4, Co2.98Cu0.02O4 and Co2.94Cu0.06O4, and the XPS spectra, including (b) Cu 2p3/2 spectrum, (c) O 1s spectrum, and (d) Co 2p3/2 and Co 2p1/2 spectra, for the Co2.98Cu0.02O4 nanorods.
Table 2 The summary of XPS
  Cu 2p3/220–22 Co 2p23–25 O 1s20–25 C 1s20–25
CuO Cu2O Cu multi-complex Co(III) Co(II) Satellite peaks Co3O4 CuO Cu2O C
This work 933.8 eV 932.2 eV 927.6 eV 780.4 eV (2p3/2) 783.4 eV (2p3/2) 786.7 eV 529.9 eV 534.3 eV 531.2 eV 284.7 eV
790 eV
795.4 eV (2p1/2) 798.8 eV (2p1/2) 802.6 eV 532.4 eV
805.8 eV
Reference 933.6 eV 932.5 eV 927.0 eV 779.7 eV 781.3 eV 786.3 eV 531.1 eV 533.0 eV 530.9 eV 284.7 eV
788.2 eV
794.8 eV 796.8 eV 802.1 eV 532.2 eV
805.2 eV


The full survey spectra of pristine Co3O4, Co2.98Cu0.02O4 and Co2.94Cu0.06O4 are presented in Fig. 5a. The binding energy at 284.7 eV on the C 1s XPS spectrum corresponds to the carbon gel. The peaks located at 284.7, 532, and 780 eV are assigned to the characteristic peaks of C 1s, O 1s, and Co 2p, respectively. The peak located at 929 eV is assigned to Cu 2p, which becomes sharper at higher Cu doping concentration and confirms the existence of Cu in Co3O4. The further analysis of the Co2.98Cu0.02O4 nanorods shows the peaks of Cu 2p, O 1s, and Co 2p (Fig. 5b–d). Fig. 5b shows the Cu 2p peaks at 927.6, 932.2, and 933.8 eV, which are attributed to multi-copper complex, Cu+, and Cu2+, respectively.20–22 This reveals again that Cu has been doped into Co3O4 nanorods successfully, including Cu+ and Cu2+.

As expected in Fig. 5c, the O 1s peaks at 529.9 and 532.4 eV can be assigned to Co3O4, and 531.2 and 534.3 eV to Cu2O and CuO peaks, respectively. The representative XPS spectrum of Co 2p in Fig. 5d shows that the peaks at 780.6, and 795.6 eV correspond to Co(III) 2p3/2 and 2p1/2, while 783.7 and 798.8 eV correspond to Co(II) 2p3/2 and 2p1/2, respectively. The weak 2p3/2 satellite features were found at 786.7 and 790 eV, and those of 2p1/2 satellite features were found at 802.6 and 805.8 eV,23−25 respectively. Based on above XPS characterization, the Cu-doping in Co3O4 nanorods were further confirmed. The XPS results are summarized in Table 2.

Optical property analyses

The optical properties of the Cu-doped Co3O4 nanorods were analyzed by UV-Vis and cathodoluminescence (CL) spectra. From the UV-Vis absorption spectra in Fig. 6a for the pristine Co3O4, Co2.98Cu0.02O4 and Co2.94Cu0.06O4, it is noteworthy that the concentration of Cu dopant affects the absorption of nanorods significantly. In the visible light region, the absorption of Cu-doped Co3O4 nanorods increases with the increase of Cu dopant concentration. In order to quantify the enhancement of Cu-doping, the corresponding absorption coefficient was calculated according to Beer's law from different thicknesses of Cu-doped Co3O4. The absorption coefficients at 550 nm wavelength for the pristine Co3O4, Co2.98Cu0.02O4 and Co2.94Cu0.06O4 are 2.1 × 104, 3.1 × 104, and 4.5 × 104 cm−1, respectively, which also increase with the increase of Cu dopant concentration. In addition, the optical band gaps of Cu-doped Co3O4 nanorods could be adjusted by varying the concentrations of Cu dopant. The pristine Co3O4 nanorods exhibit the band gaps of 1.4 and 3.5 eV. With the increase of Cu-doping, the band gaps shift to 1.35 and 2.75 eV for the Co2.98Cu0.02O4 nanorods and to 1.3 and 2.5 eV for Co2.94Cu0.06O4 nanorods. These values were extracted to the energy (= ) axis at α = 0 according to the linear portion of (αhν)2 plots, which were inserted in Fig. 6b–d, where α, h, and ν represent absorption coefficient, Plank constant, and wavelength, respectively.26,27
image file: c4ra01463d-f6.tif
Fig. 6 (a) UV-Vis absorption spectrum of the pristine Co3O4, Co2.98Cu0.02O4 and Co2.94Cu0.06O4 nanorods. Plots of (αhν)2versus hν show the direct band gap of (b) pristine Co3O4, (c) Co2.98Cu0.02O4 and (d) Co2.94Cu0.06O4 nanorods.

The reason for band gap narrowing is because Cu dopant replaces Co place and bonds with O. CuO and Cu2O band gaps (∼1.21 eV and ∼2.10 eV, respectively) are smaller than Co3O4 band gaps (1.40 eV and 3.50 eV, respectively). Besides, Cu+ and Cu2+ could occupy Co2+ or Co3+ place in Co3O4, which generates one more hole carriers.

The absorption coefficient of ∼104 cm−1 and direct band gaps of 1.4, 1.35, and 1.3 eV for Cu-doped Co3O4 nanorods suggest that they could serve as an efficient light harvesting agent for potential photovoltaic and photodiode applications.

Their CL properties were determined in Fig. 7a–c. The spectra are basically composed of two peaks and the left peak can be deconvoluted into two sub-peaks (peak 1 and peak 2). The Co3O4 absorption bands have been reported to be related to the ligand field transitions of Co3+ and Co2+ ions in octahedral and tetrahedral coordination, respectively, and the charge transfer process between Co2+ and Co3+ ions and those from oxygen ligands to Co ions.28 The peak 1 and 2 are related to the transition between Co2+ and Co3+ bands, which are located at 2.5–2.9 eV and ∼2.0 eV respectively. These peaks could correspond to the larger band gap transition in UV-Vis analyses. Comparing them with the UV-Vis band gap fitting results, the corresponding energies of these peaks are smaller than that of the larger band gap observed in UV-Vis analyses, which could be attributed to the deep-level emissions associated with defects.29 The peak 3 could be corresponded to the band gap (∼1.4 eV) transition of Cu-doped Co3O4 nanorods and it is consistent with the previous UV-Vis measurement. It is noteworthy that as the Cu-doping concentration increases, the peak 1 shifts to UV region while peak 2 and peak 3 shifts to infrared light region. The results and their comparison with other works are also summarized in Table 3 with the detailed analysis and description.


image file: c4ra01463d-f7.tif
Fig. 7 Cathodoluminescence (CL) spectra of (a) pristine Co3O4, (b) Co2.98Cu0.02O4 and (c) Co2.94Cu0.06O4 nanorods.
Table 3 The summary of CL analyses
  Center of peak 1 (nm) Center of peak 2 (nm) Center of peak 3 (nm)
Pristine Co3O4 488 610 890
Co2.98Cu0.02O4 473 616 891
Co2.94Cu0.06O4 472 617 893


Conclusions

In summary, tunable band gaps of Co3−xCuxO4 nanorods, were synthesized via solution process and first reported in this study. Detailed characterizations of the nanorods show a crystalline structure and that the Cu has been doped into Co3O4 successfully, including Cu+ and Cu2+. The optical properties were also carefully examined, showing that the absorption coefficient of nanorods increases with the increase of Cu dopant. The Cu-doped Co3O4 nanorods exhibit dual direct bandgaps and could be adjusted at 1.3–1.5 eV, and 2.55–3.5 eV by Cu-doping using different Cu-doping concentration (the higher the doping, the lower the band gap). The Cu-doped Co3O4 nanorods also reveal an absorption coefficient about 104 cm−1, which suggests that they could serve as a light harvesting agent for potential applications in photovoltaics and photodiodes.

Acknowledgements

This work was supported by National Science Council under project number NSC 100-2221-E-007-002-MY3. The authors thank C. H. Chen, Y. H. Chang, C. H. Hsu and S. Y. Wei at National Tsing-Hua University for optical property measurements. The authors also thank Prof. J. Y. Gan, C. N. Yeh, C. W. Juan, Y. J. Hong, and Y. M. Hsu for the discussion of the nanorod synthesis.

Notes and references

  1. J. Wollenstein, M. Burgmair, G. Plescher, T. Sulima, J. Hildenbrand, H. Bottner and I. Eisele, Sens. Actuators, B, 2003, 93, 442–448 CrossRef CAS.
  2. W. Y. Li, L. N. Xu and J. Chen, Adv. Funct. Mater., 2005, 15, 851–857 CrossRef CAS.
  3. T. Maruyama and S. Arai, J. Electrochem. Soc., 1996, 143, 1383–1386 CrossRef CAS PubMed.
  4. L. H. Hu, Q. Peng and Y. D. Li, J. Am. Chem. Soc., 2008, 130, 16136–16137 CrossRef CAS PubMed.
  5. K. Chidambaram, L. K. Malhotra and K. L. Chopra, Thin Solid Films, 1982, 87, 365–371 CrossRef CAS.
  6. Y. G. Li, P. Hasin and Y. Y. Wu, Adv. Mater., 2010, 22, 1926–1929 CrossRef CAS PubMed.
  7. X. H. Xia, J. P. Tu, Y. J. Mai, X. L. Wang, C. D. Gu and X. B. Zhao, J. Mater. Chem., 2011, 21, 9319–9325 RSC.
  8. X. W. Lou, D. Deng, J. Y. Lee, J. Feng and L. A. Archer, Adv. Mater., 2008, 20, 258 CrossRef CAS.
  9. Y. G. Li, B. Tan and Y. Y. Wu, Nano Lett., 2008, 8, 265–270 CrossRef CAS PubMed.
  10. F. Nobili, F. Croce, R. Tossici, I. Meschini, P. Reale and R. Marassi, J. Power Sources, 2012, 197, 276–284 CrossRef CAS PubMed.
  11. Q. Zhang, Z. D. Wei, C. Liu, X. Liu, X. Q. Qi, S. G. Chen, W. Ding, Y. Ma, F. Shi and Y. M. Zhou, Int. J. Hydrogen Energy, 2012, 37, 822–830 CrossRef CAS PubMed.
  12. Z. Sheikhi Mehrabadi, A. Ahmadpour, N. Shahtahmasebi and M. M. Bagheri Mohagheghi, Phys. Scr., 2011, 84, 015801 CrossRef.
  13. L. Run, D. Ying and H. J. Baibiao, Phys. Chem. Lett., 2013, 4, 2223–2229 CrossRef.
  14. L. Run and N. J. English, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 155209–155213 CrossRef.
  15. L. Run and N. J. English, Chem. Mater., 2010, 22, 1626 Search PubMed.
  16. V. G. Hadjiev, M. N. Iliev and I. V. Vergilov, J. Phys. C: Solid State Phys., 1988, 21(7), L199–L201 CrossRef.
  17. H. Shirai, Y. Morioka and I. Nakagawa, J. Phys. Soc. Jpn., 1982, 51(2), 592–597 CrossRef CAS.
  18. B. C. Cheng, Y. H. Xiao, G. S. Wu and L. D. Zhang, Appl. Phys. Lett., 2004, 84(3), 416–418 CrossRef CAS PubMed.
  19. S. K. Sharma and G. J. Exarhos, Solid State Phenom., 1997, 55, 32–37 CrossRef CAS.
  20. R. A. Zarate, F. Hevia, S. Fuentes, V. M. Fuenzalida and A. Zuniga, J. Solid State Chem., 2007, 180(4), 1464–1469 CrossRef CAS PubMed.
  21. E. Abelev, D. Starosvetsky and Y. Ein-Eli, Electrochim. Acta, 2007, 52(5), 1975–1982 CrossRef CAS PubMed.
  22. J. P. Chang, Z. Zhang, H. Xu, H. H. Sawin and J. W. Butterbaugh, J. Vac. Sci. Technol., A, 1997, 15(6), 2959–2967 CAS.
  23. S. C. Petitto and M. A. Langell, J. Vac. Sci. Technol., A, 2004, 22(4), 1690 CAS.
  24. T. He, D. R. Chen, X. L. Jiao, Y. L. Wang and Y. Z. Duan, Chem. Mater., 2005, 17(15), 4023–4030 CrossRef CAS.
  25. W. L. Guo, Y. F. E, L. Gao, L. Z. Fan and S. H. Yang, Chem. Commun., 2010, 46(8), 1290–1292 RSC.
  26. H. Yamamoto, S. Tanaka and K. Hirao, J. Appl. Phys., 2003, 93, 4158–4162 CrossRef CAS PubMed.
  27. G. X. Wang, X. P. Shen, J. Horvat, B. Wang, H. Liu, D. Wexler and J. Yao, J. Phys. Chem. C, 2009, 113, 4357–4361 CAS.
  28. A. Llavona, C. Diaz-Guerra, M. C. Sanchez and L. Perez, Mater. Chem. Phys., 2010, 124(2–3), 1177–1181 CrossRef CAS PubMed.
  29. Z. G. Chen, J. Zou, G. Liu, F. Li, Y. Wang, L. Z. Wang, X. L. Yuan, T. Sekiguchi, H. M. Cheng and G. Q. Lu, ACS Nano, 2008, 2, 2183–2191 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.