Next Article in Journal
An Ultrahigh Sensitive Paper-Based Pressure Sensor with Intelligent Thermotherapy for Skin-Integrated Electronics
Next Article in Special Issue
Synthesis of Boron Nitride Nanotubes Using Plasma-Assisted CVD Catalyzed by Cu Nanoparticles and Oxygen
Previous Article in Journal
Reduction in Processing Time in Ca3Co4O9+δ Ceramics through Nanoprecursors Produced by an Easily Scalable and Environmentally Friendly Process
Previous Article in Special Issue
Generating Silicon Nanofiber Clusters from Grinding Sludge by Millisecond Pulsed Laser Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Silicon Nanotubes Fabricated by Wet Chemical Etching of ZnO/Si Core–Shell Nanowires

1
International Center for Materials Nanoarchitectonics, National Institute for Materials Science, 1-1 Namiki, Tsukuba 305-0044, Japan
2
Institute of Applied Physics, University of Tsukuba, 1-1-1 Tennodai, Tsukuba 305-8573, Japan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(12), 2535; https://0-doi-org.brum.beds.ac.uk/10.3390/nano10122535
Submission received: 27 November 2020 / Revised: 13 December 2020 / Accepted: 15 December 2020 / Published: 17 December 2020
(This article belongs to the Special Issue Advances in Nanowire)

Abstract

:
Silicon nanotubes (SiNTs) have garnered a great deal of interest for both their synthesis and their potential for application to high-capacity energy storage, biosensors, and selective transport. In this study, we report a convenient and low-cost route to the fabrication of vertically aligned SiNTs via a wet-etching process that enables the control of the wall thickness of SiNTs by varying the gas flux and growth temperature. Transmission electron microscopy (TEM) characterization showed the resultant SiNTs to have an amorphous nature and a hexagonal hollow core. These SiNTs can be crystallized by thermal annealing.

1. Introduction

Nanotubular structures are characterized by a large specific surface area with more exposed active sites, giving them unique physical and chemical properties. Silicon nanotubes (SiNTs) are a highly promising semiconductor material due to their potential application to energy storage, biosensors, and chemical transport. The axial hollow space within the SiNTs can offer improved electrochemical performance for lithium-ion batteries since it provides additional free space to accommodate the massive volume expansion that occurs during the lithiation/delithiation process, thus preventing the pulverization of silicon [1,2,3,4,5,6,7]. The large specific surface area of SiNTs also minimizes the lithium-ion diffusion length and increases the ion flux. In other applications, biomolecular species can directly flow into the inner cavity to gate a SiNT transistor for application as an intracellular sensor [8]. The inner and outer surfaces of nanotubular structures can also be functionalized differentially to enable selective transport, separation, and filtering [9,10].
Today, SiNTs are mainly synthesized using 1-D sacrificial templates. A Si shell layer is deposited on a template such as ZnO [2,5], Ge [9,11], or MgO [12] nanowires, followed by selectively etching away the templates to form the SiNTs. Top-down techniques using lithography [13,14,15] and self-assembled nanosphere bead templates [16,17,18] have more recently been developed to fabricate SiNTs, but they require a complex manufacturing process, long processing time, or the use of noble metals. Other methods such as chemical etching [19] and electrochemical formation [20] cannot produce vertically aligned SiNTs, so they have limited applications.
Sacrificial templating methods based on ZnO nanowires (NWs) are regarded as a convenient, low-cost, and controllable method of fabricating SiNTs. Vertically aligned ZnO nanowires are synthesized using the hydrothermal method at temperatures of under 100 °C, followed by the deposition of a Si shell layer on the ZnO template using chemical vapor deposition (CVD). Finally, the ZnO core is selectively etched via a reduction process at 600 °C with 50% H2 in N2 for 24 h [2]. However, this etching process is highly energy and time consuming, hindering the future large-scale production of SiNTs.
In this study, we propose the use of a wet chemical etching method to remove the ZnO template instead of the gas-phase etching process, resulting in a convenient, cheap, and timesaving approach to the fabrication of vertically aligned SiNTs. We investigated the Si shell growth rate as a function of the precursor gas flux and growth temperature to control the SiNTs’ wall thickness. The morphology, crystallinity, and elemental composition of the SiNTs were characterized. We also studied the crystallization of Si shell layers with different shell thickness by thermal annealing.

2. Materials and Methods

An n-type Si(100) substrate was sputtered with 100 nm of ZnO to serve as a seed layer for nanowire growth. The growth solution consisted of 10 mM zinc nitrate, 5 mM hexamethylenetetramine (HMTA), and 0.6 M ammonium hydroxide. A small amount of sodium citrate (0.08 mM) was added to the solution to maintain the hexagonal cross-sectional shape of the nanowires. The substrate was then floated on the solution surface with its front side downwards and placed in an oven at 95 °C for 6 h to perform the hydrothermal growth. This growth condition was optimized to produce ZnO NWs with a large wire-to-wire spacing and high aspect ratio. All the chemicals were purchased from Wako Pure Chemical Industries (Tokyo, Japan).
After rinsing with deionized water and isopropyl alcohol (IPA), the samples were loaded into an ultra-high-vacuum chemical vapor deposition (UHV-CVD) chamber with a background pressure of 2 × 10−6 Pa. The Si shell layer was coated on the ZnO NWs using SiH4 as the precursor gas at temperatures of 650 and 700 °C. The growth time was increased from 6 to 20 min. The details of the UHV-CVD method are reported elsewhere [21,22,23]. The SiH4 gas flow was varied from 6 to 20 sccm, and the total pressure was set at 700 Pa by mixing with nitrogen gas.
To finally grow the Si nanotube structures, the ZnO cores were selectively wet-etched in phosphoric acid (85%) for 6 min at room temperature and then rinsed repeatedly with deionized water to remove any residual acid. Additionally, the ZnO/Si core–shell NWs were treated by rapid thermal annealing at 800 °C for 5 min under an N2 atmosphere to study the crystallization of the Si shell layers. The full fabrication procedure is illustrated in Figure 1.
Sample Characterization. Scanning electron microscopy (SEM) images were recorded using a Hitachi S-8000 SEM (Tokyo, Japan) at an acceleration voltage of 5 kV. Transmission electron microscopic (TEM) and energy-dispersive X-ray spectrometry (EDX) analyses were performed using a JEOL 2100F transmission electron microscope (Tokyo, Japan) at 200 kV. Micro-Raman scattering (Photon Design, Tokyo, Japan) measurements were carried out using a 355 nm excitation beam with a 100× objective, with the power set at 0.02 mW to prevent local heating effects [24,25]. X-ray diffraction (XRD) patterns were obtained using a PANalytical X’Pert Pro MRD system (Tokyo, Japan) with a parallel Cu Kα beam.

3. Results and Discussion

Figure 2 shows scanning electron microscopy (SEM) images of the ZnO/Si core–shell NWs as a function of the Si shell growth temperature and SiH4 gas flux. The CVD process forms Si shell layers on the ZnO core, with thicknesses that vary from 12 to 21 nm. The Si shell growth rate increased sublinearly with SiH4 gas flux, which was caused by the higher density of the available source materials for its formation. Raising the growth temperature from 650 to 700 °C enhanced the Si shell growth rate due to the faster decomposition rate for the precursor gas at the higher temperature. To produce an appropriate shell thickness for each growth condition, the growth times for the samples shown in Figure 2a–f were set at 20, 20, 10, 15, 6, and 5 min, respectively.
The SEM and TEM images of the nanowires with Si shells grown at 700 °C using 20 sccm of SiH4 gas flux for 6 min are shown in Figure 3. The SEM image in Figure 3b shows a darker area at the center of every single nanowire compared to Figure 3a, indicating the formation of cavities after the wet-etching process. The XRD pattern recorded from the nanowires in Figure 3c shows an absence of ZnO peaks after etching, which confirms the removal of the ZnO core. The presence of the silicon peak is mainly due to the Si substrate, since the crystallinity of the Si shell is not perfect due to the growth on the ZnO surface being closer to amorphous. This amorphous nature will be also discussed later.
The TEM images in Figure 3d,e demonstrate the sealed tubular structure with a wall thickness of approximately 20 nm. The hazy halo pattern of the selected area electron diffraction (SAED) image in Figure 3e reveals the amorphous nature of the Si shell layer. The cross-sectional TEM image through the bottom of a nanotube in Figure 3f shows a hexagonal hollow space, which originates from the cross-sectional shape of ZnO nanowires. The elemental mapping and linescan analysis of the SiNTs are shown in Figure 4. The energy-dispersive X-ray (EDX) mapping and linescan for Si demonstrate the successful synthesis of SiNTs, and the TEM-EDX spectrum in Figure 4d shows that the ZnO core was completely etched out. Note that the Cu peaks in the spectrum come from the Cu grid. Moreover, the caps of these vertically aligned SiNTs can be etched using reactive-ion etching (RIE) [6] to produce open-cap SiNTs that have a wider potential range of applications.
To improve the electrical and optical properties, ZnO/Si core–shell NWs were thermally annealed to crystallize the silicon shell layers. Figure 5a–d show almost no change in the morphology of the nanowires after thermal annealing at 800 °C for 5 min. Figure 5e,f show the results of Raman measurements at an excitation wavelength of 355 nm. This wavelength was selected to shorten the penetration depth in Si, resulting in the suppression of the signal from the ZnO core. No Raman peaks were observed before annealing for both samples with thick (20 nm) or thin (12 nm) Si shell layers, showing that the Si shell layers are not crystalline but amorphous. After annealing, the Si optical phonon peak shifted to higher wavenumbers and reached the bulk value (520.1 cm−1) upon increasing the Si shell thickness. The downshift and asymmetric broadening to lower wavenumbers observed for core–shell NWs with thin shell layers can be explained by the phonon confinement effect [24,26,27,28,29].
Richter et al. [26] and Campbell and Fauchet [27] reported that the phonon confinement model and the Raman intensity is given by
I ( ω ) = | C ( 0 , q ) | 2 [ ω ω ( q ) ] 2 + ( Γ 0 / 2 ) 2 d 3 q
where C(0,q) is a Fourier coefficient of the confinement function, ω(q) is the Si phonon dispersion, and Γ0 is the full width at half maximum of the reference Si. Here, we used the following relations: ∣C(0,q)∣2 = exp(−q2d2/16π2) and ω(q) = [A + Bcos(/2)]0.5 + D, with A = 1.714 × 1015 cm−2 and B = 105 cm−2 [25]. D is an adjusting parameter for the bulk Si. These relations can be used for a cylindrical structure such as an NW. The Si shell structure after removing the ZnO core is an NT structure and is different from a typical NW structure. Considering the small diameter of the Si/ZnO core–shell NWs, we treated the Si shell structures as pseudo-NW structures. The fitting result is shown in Figure 6. The phonon correlation length estimated by the fitting was 7–8 nm. This value is close to the thickness of the Si shell layers. The fitting is, however, not so good on the high- and low-wavenumber sides. There are several possible reasons for this. First, the relationship used for the approximation is that employed for NWs. Second, there is the problem of crystallinity. This is considered the main reason for the incompleteness of the lower-wavenumber fitting. Third, the effect of the strain from the ZnO forming the heterojunction, mainly due to the incompleteness of the fitting on the higher-wavenumber side, may be a factor. The Si optical phonon peak did, in fact, show a higher Raman shift than the value obtained for bulk Si (520.1 cm−1), as shown in Figure 5g. The upshift means that compressive stress is induced from the ZnO core region.
This is expected to be a promising approach to the development of battery materials that require low-cost and high-capacity materials. Our method is applicable to materials that can be selectively etched. For example, by replacing ZnO with a material such as Ge [22,23], epitaxial growth on Ge nanowires is possible: it is expected that nanotube structures with a smoother surface can be formed. Impurity doping is also important in making materials funcstional [29,30,31]. Since this CVD method can also be used for doping impurities, multiple future applications are anticipated.
In summary, we have developed a convenient and low-cost route to fabricating vertically aligned SiNTs by applying a wet-etching process to the ZnO region of ZnO/Si core–shell NWs. The Si shell growth rate can be controlled by adjusting the precursor gas flux and growth temperature. The fabricated SiNTs showed an amorphous nature, with hexagonal hollow spaces, and were crystallized by thermal annealing, showing that the crystallinity can also be controlled.

Author Contributions

Conceptualization, N.F.; methodology, Y.-L.S. and N.F.; validation, Y.-L.S. and N.F.; formal analysis, Y.-L.S., X.-D.Z., and N.F.; investigation, Y.-L.S., X.-D.Z., and W.J.; resources, W.J. and N.F.; data curation, Y.-L.S. and X.-D.Z.; writing—original draft preparation, Y.-L.S.; writing—review and editing, N.F.; visualization, Y.-L.S.; supervision, N.F.; project administration, N.F.; funding acquisition, N.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS Kakenhi, grant numbers 26246021 and 20K21135.

Acknowledgments

This work was supported by the World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Park, M.H.; Kim, M.G.; Joo, J.; Kim, K.; Kim, J.; Ahn, S.; Cui, Y.; Cho, J. Silicon nanotube battery anodes. Nano Lett. 2009, 9, 3844–3847. [Google Scholar] [CrossRef]
  2. Song, T.; Xia, J.; Lee, J.-H.; Lee, D.H.; Kwon, M.-S.; Choi, J.-M.; Wu, J.; Doo, S.K.; Chang, H.; Park, W.I.; et al. Arrays of Sealed Silicon Nanotubes As Anodes for Lithium Ion Batteries. Nano Lett. 2010, 10, 1710–1716. [Google Scholar] [CrossRef]
  3. Wu, H.; Chan, G.; Choi, J.W.; Ryu, I.; Yao, Y.; McDowell, M.T.; Lee, S.W.; Jackson, A.; Yang, Y.; Hu, L.; et al. Stable cycling of double-walled silicon nanotube battery anodes through solid–electrolyte interphase control. Nat. Nanotechnol. 2012, 7, 310–315. [Google Scholar] [CrossRef]
  4. Lotfabad, E.M.; Kalisvaart, P.; Kohandehghan, A.; Cui, K.; Kupsta, M.; Farbod, B.; Mitlin, D. Si nanotubes ALD coated with TiO2, TiN or Al2O3 as high performance lithium ion battery anodes. J. Mater. Chem. A 2014, 2, 2504–2516. [Google Scholar] [CrossRef]
  5. Tesfaye, A.T.; Gonzalez, R.; Coffer, J.L.; Djenizian, T. Porous Silicon Nanotube Arrays as Anode Material for Li-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 20495–20498. [Google Scholar] [CrossRef] [Green Version]
  6. Ha, J.; Paik, U. Hydrogen treated, cap-opened Si nanotubes array anode for high power lithium ion battery. J. Power Sources 2013, 244, 463–468. [Google Scholar] [CrossRef]
  7. Wang, C.; Wen, J.; Luo, F.; Quan, B.; Li, H.; Wei, Y.; Gu, C.; Li, J. Anisotropic expansion and size-dependent fracture of silicon nanotubes during lithiation. J. Mater. Chem. A 2019, 7, 15113–15122. [Google Scholar] [CrossRef] [Green Version]
  8. Gao, R.; Strehle, S.; Tian, B.; Cohen-Karni, T.; Xie, P.; Duan, X.; Qing, Q.; Lieber, C.M. Outside looking in: Nanotube transistor intracellular sensors. Nano Lett. 2012, 12, 3329–3333. [Google Scholar] [CrossRef] [Green Version]
  9. Ben-Ishai, M.; Patolsky, F. Wall-selective chemical alteration of silicon nanotube molecular carriers. J. Am. Chem. Soc. 2011, 133, 1545–1552. [Google Scholar] [CrossRef]
  10. Le, N.T.; Tian, Y.; Gonzalez-Rodriguez, R.; Coffer, J.L. Silicon nanotubes as potential therapeutic platforms. Pharmaceutics 2019, 11, 571. [Google Scholar] [CrossRef] [Green Version]
  11. Kwon, S.; Chen, Z.C.Y.; Noh, H.; Lee, J.H.; Liu, H.; Cha, J.N.; Xiang, J. Selective functionalization and loading of biomolecules in crystalline silicon nanotube field-effect-transistors. Nanoscale 2014, 6, 7847–7852. [Google Scholar] [CrossRef]
  12. Epur, R.; Hanumantha, P.J.; Datta, M.K.; Hong, D.; Gattu, B.; Kumta, P.N. A simple and scalable approach to hollow silicon nanotube (h-SiNT) anode architectures of superior electrochemical stability and reversible capacity. J. Mater. Chem. A 2015, 3, 11117–11129. [Google Scholar] [CrossRef]
  13. Kim, Y.Y.; Kim, H.J.; Jeong, J.H.; Lee, J.; Choi, J.H.; Jung, J.Y.; Lee, J.H.; Cheng, H.; Lee, K.W.; Choi, D.G. Facile Fabrication of Silicon Nanotube Arrays and Their Application in Lithium-Ion Batteries. Adv. Eng. Mater. 2016, 18, 1349–1353. [Google Scholar] [CrossRef]
  14. Jeong, H.; Lee, J.; Bok, C.; Lee, S.H.; Yoo, S. Fabrication of Vertical Silicon Nanotube Array Using Spacer Patterning Technique and Metal-Assisted Chemical Etching. IEEE Trans. Nanotechnol. 2017, 16, 130–134. [Google Scholar] [CrossRef]
  15. Laney, S.K.; Li, T.; Michalska, M.; Ramirez, F.; Portnoi, M.; Oh, J.; Tiwari, M.K.; Thayne, I.G.; Parkin, I.P.; Papakonstantinou, I. Spacer-Defined Intrinsic Multiple Patterning. ACS Nano 2020, 14, 12091–12100. [Google Scholar] [CrossRef]
  16. Soleimani-Amiri, S.; Gholizadeh, A.; Rajabali, S.; Sanaee, Z.; Mohajerzadeh, S. Formation of Si nanorods and hollow nano-structures using high precision plasma-treated nanosphere lithography. RSC Adv. 2014, 4, 12701–12709. [Google Scholar] [CrossRef]
  17. He, Y.; Che, X.; Que, L. A top-down fabrication process for vertical hollow silicon nanopillars. J. Microelectromech. Syst. 2016, 25, 662–667. [Google Scholar] [CrossRef]
  18. Tseng, Y.M.; Gu, R.Y.; Cheng, S.L. Design and fabrication of vertically aligned single-crystalline Si nanotube arrays and their enhanced broadband absorption properties. Appl. Surf. Sci. 2020, 508, 145223. [Google Scholar] [CrossRef]
  19. Zhang, C.; Cheng, H.; Liu, X. A convenient way of manufacturing silicon nanotubes on a silicon substrate. Mater. Chem. Phys. 2016, 177, 479–484. [Google Scholar] [CrossRef]
  20. Weng, W.; Yang, J.; Zhou, J.; Gu, D.; Xiao, W. Template-Free Electrochemical Formation of Silicon Nanotubes from Silica. Adv. Sci. 2020, 7, 2001492. [Google Scholar] [CrossRef]
  21. Fukata, N.; Sato, K.; Mitome, M.; Bando, Y.; Sekiguchi, T.; Kirkham, M.; Hong, J.I.; Wang, Z.L.; Snyder, R.L. Doping and Raman characterization of boron and phosphorus atoms in germanium nanowires. ACS Nano 2010, 4, 3807–3816. [Google Scholar] [CrossRef]
  22. Fukata, N.; Mitome, M.; Sekiguchi, T.; Bando, Y.; Kirkham, M.; Hong, J.-I.; Wang, Z.L.; Snyder, R.L. Characterization of Impurity Doping and Stress in Si/Ge and Ge/Si Core–Shell Nanowires. ACS Nano 2012, 6, 8887–8895. [Google Scholar] [CrossRef]
  23. Fukata, N.; Yu, M.; Jevasuwan, W.; Takei, T.; Bando, Y.; Wu, W.; Wang, Z.L. Clear Experimental Demonstration of Hole Gas Accumulation in Ge/Si Core-Shell Nanowires. ACS Nano 2015, 9, 12182–12188. [Google Scholar] [CrossRef]
  24. Fukata, N.; Oshima, T.; Murakami, K.; Kizuka, T.; Tsurui, T.; Ito, S. Phonon confinement effect of silicon nanowires synthesized by laser ablation. Appl. Phys. Lett. 2005, 86, 213112. [Google Scholar] [CrossRef]
  25. Piscanec, S.; Cantoro, M.; Ferrari, A.C.; Zapien, J.A.; Lifshitz, Y.; Lee, S.T.; Hofmann, S.; Robertson, J. Raman spectroscopy of silicon nanowires. Phys. Rev. B 2003, 68, 241312. [Google Scholar] [CrossRef] [Green Version]
  26. Richter, H.; Wang, Z.P.; Ley, L. The one phonon Raman spectrum in microcrystalline silicon. Solid State Commun. 1981, 39, 625–629. [Google Scholar] [CrossRef]
  27. Campbell, I.H.; Fauchet, P.M. The effects of microcrystal size and shape on the one phonon Raman spectra of crystalline semiconductors. Solid State Commun. 1986, 58, 739–741. [Google Scholar] [CrossRef]
  28. Fukata, N.; Oshima, T.; Okada, N.; Murakami, K.; Kizuka, T.; Tsurui, T.; Ito, S. Phonon confinement and self-limiting oxidation effect of silicon nanowires synthesized by laser ablation. J. Appl. Phys. 2006, 100, 024311. [Google Scholar] [CrossRef] [Green Version]
  29. Fukata, N. Impurity doping in silicon nanowires. Adv. Mater. 2009, 21, 2829–2832. [Google Scholar] [CrossRef]
  30. Fukata, N.; Subramani, T.; Jevasuwan, W.; Dutta, M.; Bando, Y. Functionalization of Silicon Nanostructures for Energy-Related Applications. Small 2017, 13, 1–13. [Google Scholar] [CrossRef] [Green Version]
  31. Fukata, N. Impurity doping in semiconductor nanowires. In Fundamental Properties of Semiconductor Nanowires; Fukata, N., Rurali, R., Eds.; Springer Nature: Singapore, 2020; pp. 143–181. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the fabrication procedure for silicon nanotubes (SiNTs).
Figure 1. Schematic illustration of the fabrication procedure for silicon nanotubes (SiNTs).
Nanomaterials 10 02535 g001
Figure 2. (ac) 30°-tilted SEM images of vertically aligned ZnO NWs with a Si shell layer grown at 650 °C using 10, 20, and 40 sccm of SiH4 gas flux, respectively. (df) 30°-tilted SEM images of ZnO NWs with Si shell grown at 700 °C using 6, 20, and 40 sccm of SiH4 gas flux, respectively. (g) Growth rate as a function of the SiH4 gas flux at different growth temperatures. The growth time was varied for each growth condition to ensure an appropriate diameter.
Figure 2. (ac) 30°-tilted SEM images of vertically aligned ZnO NWs with a Si shell layer grown at 650 °C using 10, 20, and 40 sccm of SiH4 gas flux, respectively. (df) 30°-tilted SEM images of ZnO NWs with Si shell grown at 700 °C using 6, 20, and 40 sccm of SiH4 gas flux, respectively. (g) Growth rate as a function of the SiH4 gas flux at different growth temperatures. The growth time was varied for each growth condition to ensure an appropriate diameter.
Nanomaterials 10 02535 g002
Figure 3. Top-view SEM images of ZnO/Si core–shell NWs (a) before and (b) after the wet-etching process. (c) Corresponding XRD pattern before and after the wet-etching process. (d) Low- and (e) high-magnification TEM images of the SiNTs. Corresponding selected area electron diffraction (SAED) pattern in the inset of (e) showing its amorphous nature. (f) Cross-sectional TEM image showing the hexagonal hollow space of the SiNTs.
Figure 3. Top-view SEM images of ZnO/Si core–shell NWs (a) before and (b) after the wet-etching process. (c) Corresponding XRD pattern before and after the wet-etching process. (d) Low- and (e) high-magnification TEM images of the SiNTs. Corresponding selected area electron diffraction (SAED) pattern in the inset of (e) showing its amorphous nature. (f) Cross-sectional TEM image showing the hexagonal hollow space of the SiNTs.
Nanomaterials 10 02535 g003
Figure 4. (a) TEM image and (b) EDX mapping for Si of the SiNTs. (c) EDX linescan for Si and Zn along the radial direction of a SiNT. The linescan position is shown as the dashed line in (a). (d) TEM-EDX spectrum recorded from the SiNTs, showing the absence of a Zn peak.
Figure 4. (a) TEM image and (b) EDX mapping for Si of the SiNTs. (c) EDX linescan for Si and Zn along the radial direction of a SiNT. The linescan position is shown as the dashed line in (a). (d) TEM-EDX spectrum recorded from the SiNTs, showing the absence of a Zn peak.
Nanomaterials 10 02535 g004
Figure 5. 30°-tilted SEM images of ZnO/Si core–shell NWs with a thick (20 nm) shell layer (a) before and (b) after thermal annealing at 800 °C for 5 min. SEM images of the nanowires with a thin (12 nm) shell layer (c) before and (d) after thermal annealing. Corresponding Raman spectra for the nanowires with (e) thick and (f) thin shell layers. The green dashed line shows the bulk value. (g) Raman shift of Si optical phonon peak for nanowires with different shell thicknesses after annealing. The error bar indicates the standard deviation of the shell thickness and Raman measurements.
Figure 5. 30°-tilted SEM images of ZnO/Si core–shell NWs with a thick (20 nm) shell layer (a) before and (b) after thermal annealing at 800 °C for 5 min. SEM images of the nanowires with a thin (12 nm) shell layer (c) before and (d) after thermal annealing. Corresponding Raman spectra for the nanowires with (e) thick and (f) thin shell layers. The green dashed line shows the bulk value. (g) Raman shift of Si optical phonon peak for nanowires with different shell thicknesses after annealing. The error bar indicates the standard deviation of the shell thickness and Raman measurements.
Nanomaterials 10 02535 g005
Figure 6. Si optical phonon peak observed for ZnO/Si core–shell NWs with a thin (~12 nm) Si shell layer. The dashed line indicates the fitting result using phonon confinement theory.
Figure 6. Si optical phonon peak observed for ZnO/Si core–shell NWs with a thin (~12 nm) Si shell layer. The dashed line indicates the fitting result using phonon confinement theory.
Nanomaterials 10 02535 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sun, Y.-L.; Zheng, X.-D.; Jevasuwan, W.; Fukata, N. Silicon Nanotubes Fabricated by Wet Chemical Etching of ZnO/Si Core–Shell Nanowires. Nanomaterials 2020, 10, 2535. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10122535

AMA Style

Sun Y-L, Zheng X-D, Jevasuwan W, Fukata N. Silicon Nanotubes Fabricated by Wet Chemical Etching of ZnO/Si Core–Shell Nanowires. Nanomaterials. 2020; 10(12):2535. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10122535

Chicago/Turabian Style

Sun, Yong-Lie, Xiang-Dong Zheng, Wipakorn Jevasuwan, and Naoki Fukata. 2020. "Silicon Nanotubes Fabricated by Wet Chemical Etching of ZnO/Si Core–Shell Nanowires" Nanomaterials 10, no. 12: 2535. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10122535

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop