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

The study of optical band edge property of bismuth oxide nanowires α-Bi2O3

Open Access Open Access

Abstract

The α-phase Bi2O3 (α-Bi2O3) is a crucial and potential visible-light photocatalyst material needless of intentional doping on accommodating band gap. The understanding on fundamental optical property of α-Bi2O3 is important for its extended applications. In this study, bismuth oxide nanowires with diameters from tens to hundreds nm have been grown by vapor transport method driven with vapor-liquid-solid mechanism on Si substrate. High-resolution transmission electron microscopy and Raman measurement confirm α phase of monoclinic structure for the as-grown nanowires. The axial direction for the as-grown nanowires was along <1¯22>. The band-edge structure of α-Bi2O3 has been probed experimentally by thermoreflectance (TR) spectroscopy. The direct band gap was determined accurately to be 2.91 eV at 300 K. Temperature-dependent TR measurements of 30-300 K were carried out to evaluate temperature-energy shift and line-width broadening effect for the band edge of α-Bi2O3 thin-film nanowires. Photoluminescence (PL) experiments at 30 and 300 K were carried out to identify band-edge emission as well as defect luminescence for the α-Bi2O3 nanowires. On the basis of experimental analyses of TR and PL, optical characteristics of direct band edge of α-Bi2O3 nanowires have thus been realized.

©2013 Optical Society of America

1. Introduction

Recently bismuth contained oxides serve as a promising candidate for a variety of applications in optoelectronics and microelectronics devices. Among them, nanostructral-photocatalyst bismuth oxide possesses an effective function for self-cleaned activity on architecture, assistance in germ and virus sterilization as well as for green and environmental protection use.

Thin film Bi2O3 usually forms a nanostructure with different polymorphisms of α-, β-, γ-, δ-, ε-, and ω-phases [1, 2], which may be a potential high-efficiency photocatalyst with a band gap below 3 eV. The values of direct band gap for the commonly forming phases of α- and β-Bi2O3 obtained by optical absorption [35] revealed lower energy than that of anatase TiO2 with an indirect gap of ~3.32 eV [6]. The smaller band gap of Bi2O3 means that the bismuth oxide nanostructures can generate electron-hole pairs more efficiently under visible light’s excitation beneath the sunlight shiny.

For the crystal structure of bismuth oxide, α-Bi2O3 is a stabilized low-temperature phase with monoclinic structure in the Bismites. The most stable form of the room temperature variety in Bi2O3 is the α monoclinic polymorph [1]. The high temperature phase of bismuth oxide is usually a cubic form, δ- Bi2O3. The polymorph of δ- Bi2O3 only stabilizes between 730 °C and the melting point 824 °C [2]. The γ-phase Bi2O3 appears in a body centered cubic structure formed below 639 °C. The β-Bi2O3 is a tetragonal structure formed below 650 °C. The formation of the β- polymorphism depends on impurities and reaction condition in the oxide. Both γ and β are metastable phases, which can be obtained by cooling during the growth process. If the crystalline nanomaterial was tempered at some established lower temperatures, the stabilization form is usually the α-Bi2O3 phase. Among all the polymorphisms of Bi2O3, α and β phases have been proven to be effective and sensitive photcatalysts operated in visible region [3, 7, 8]. α-Bi2O3 is the most stable phase existed in the Bismite photocatalysts, however, optical characteristic and band-edge structure of α-Bi2O3 have not yet been comprehensively studied hereto.

In this paper, we have characterized the band-edge optical property of α-Bi2O3 thin film nanowires using thermoreflectance (TR) and photoluminescence (PL) experiments. The α-Bi2O3 nanowires were grown by ambient controlled vapor transport process using vapor-liquid-solid (VLS) mechanism. The growth direction of the individual nanowire was along <1¯22>. The TR results show direct semiconductor behavior of α-Bi2O3 with a direct gap close to 2.91 eV at 300 K. PL measurements identify the band-edge emission and defect luminescence for the α-Bi2O3 nanowires. Temperature dependent TR measurements at several temperatures between 30 and 300 K were respectively carried out. The temperature dependences of energies and broadening parameters for the direct band gap of α-Bi2O3 nanowires are analyzed and discussed.

2. Experimental details

The growth of Bi2O3 nanowires was carried out using a simple and effective ambient controlled vapor transport process. A horizontal tube furnace with three independent heating zones was used for the thin film growth. Bismuth powder (1.5 g, 99.95%) was the source material. A thin layer (5 nm) of Au was pre-deposited on Si (100) substrate by DC sputtering using for catalyst of the VLS growth. The tube furnace was initially heated to 600 °C and then oxygen and argon mixtures (100 SCCM) were fed into the quartz tube. The growth temperature was set at ~400 °C with background pressure of 0.3 Torr. The growth time was about two hours. After the growth, the chamber tube was cooled down to room temperature under ambient pressure. X-ray diffraction measurement revealed α phase of the as-grown Bi2O3 nanowires. The lattice constants were analyzed and determined to be a = 5.85 Å, b = 8.17 Å, c = 7.51 Å, and β = 113°, respectively.

TR experiments were carried out by an indirect heating manner with a gold-evaporated quartz plate as the heating element [9, 10]. The thin sheet-type sample was closely attached on the heating element by silicone grease. The on-off heating disturbance was uniformly modulated the individual nanorods periodically. An 150 W xenon-arc lamp filtered by a PTI 0.2-m monochromator provided the monochromatic light. The incident light was focusing onto the sample with a spot size of ~100 μm2. The reflected and scattering lights from the thin-film nanorods were collected and detected by a photomultiplier tube. The signal was detected and recorded via an EG&G 7265 lock-in amplifier and personal computer. A RMC model 22 closed-cycle cryogenic refrigerator with model 4075 thermometer controller facilitated the temperature-dependent measurements. PL spectra of the nanowires samples were detected by a QE65000 charge-couple-devices imaging spectrometer. A Q-switched diode-pumped solid-state laser (λ = 266 nm) acted as the pumping light source. The measurements were done from 1.25 to 4.5 eV at 300 and 30 K, repectively.

3. Results and discussion

Figure 1(a) shows the scanning electron microscopy (SEM) image of the top view of the Bi2O3 nanowires. The cross section view of the thin-film nanowires is also displayed in Fig. 1(b), which indicates the Bi2O3 thin film is consisted of a compact continuous film (~960 nm) initially deposited on Si (100) substrate, and then the nanowires of high density were grown on the continuous film. The average thickness of the Bi2O3 thin-film nanowires was about 24.5 μm. In order to verify crystal structure and stochiometry of the nanowires, transmission electron microscopy (TEM) and energy dispersive X-ray spectroscopy (EDS) were carried out. Figures 1(c) and 1(d) show the TEM image and EDS spectrum of one single nanowire of Bi2O3. The wire width was ~298 nm and the EDS result depicted some related X-ray peaks coming from the Bi and O elements which can approximately verify the stoichiometric content of the Bi2O3 nanowire after analysis [see Fig. 1(d)]. Figure 1(e) shows the high resolution TEM image of the Bi2O3 nanowire derived from Fig. 1(c). The picture reveals that the interplanar spacing of c-axis estimated from the TEM image is d(001) = 0.75 nm, which is identical with that of the α-phase Bi2O3. The TEM image of Fig. 1(e) also indicates the axial direction of the single nanowire is grown along <1¯22> direction similar to that of the previous nanowires obtained by a reaction of trimethylbismuth and oxygen mixture on Si (100) substrate [11]. Some of the α-Bi2O3 nanowires were usually found to grow along the other direction of <010> depending on growth condition [12, 13]. The selected-area electron diffraction (SAED) pattern of the single nanowire {i.e. with [210] zone axis} in Fig. 1(f) also indicates α crystalline phase of the Bi2O3 nanowire with the growth direction along [1¯22].

 figure: Fig. 1

Fig. 1 SEM images of (a) top view and (b) cross-section view for as-grown α-Bi2O3 nanowires on Si substrate. (c) TEM image of a single nanowire with diameter of ~298 nm, the corresponding EDS spectrum and the content data are shown in (d). (e) High-resolution TEM image of the α-Bi2O3 nanowire, the interplannar distance of d(001) is indicated. (f) The related SAED pattern of the α-Bi2O3 nanowire with zone axis along [210].

Download Full Size | PDF

To further verify the structure of the Bi2O3 nanowires, Raman spectroscopy has been implemented at 300 K. Figure 2 shows the Raman spectrum of the as-grown Bi2O3 nanowires between 100 and 700 cm−1. A lot of peak features observed due to the α-Bi2O3 is a double refraction biaxial crystal with a lower symmetry of monoclinic structure [14]. Ten Raman peaks above 100 cm−1 (see Fig. 2) have been observed in good agreements with the Raman modes seen in previous results [1214], which further confirm the nanowires are comprising in monoclinic α-Bi2O3. As shown in Fig. 2, the 119 cm−1 mode is coming from Ag symmetry caused by mainly the Bi atoms’ participations. Modes of 138 (Ag) and 153 cm−1 (Bg) may come from the displacements of both Bi and O atoms in the α-Bi2O3 lattice. The Raman peaks of the higher frequency modes 183, 211, 279, 313, 410, 446, and 521 cm−1 are attributed to the displacements of the O atoms in α-Bi2O3. These vibration modes possess Ag, Bg or both symmetries, which can be defined and distinguished by tensor selection rule using polarized micro-Raman spectroscopy [12, 14]. The Raman peak pattern in Fig. 2 also identifies the low-temperature α phase of the as-grown bismuth oxide nanowires.

 figure: Fig. 2

Fig. 2 Raman peak pattern of α-Bi2O3 thin-film nanowires with the vibration frequencies between 100 and 700 cm−1.

Download Full Size | PDF

Figure 3(a) shows the temperature-dependent TR spectra of the α-Bi2O3 thin-film nanowires from 300 K down to 30 K near the band edge. TR has been proven to be more effective for the characterization of optical properties in semiconductor nanostructures [9, 15]. The TR can keep uniformity of periodically thermal modulation of each nanowire to result in easily resolved derivative line-shape spectrum that emphasizes at direct critical-point transition [16]. As shown in Fig. 3(a) by dashed lines are the experimental TR spectra of band-edge transition of α-Bi2O3 nanowires. The solid lines are the least-square fits of the experimental data using a first derivative Lorentzian line-shape function appropriate for direct band gap transition expressed as [17]:

ΔRR=Re[Aejφ(EEg+jΓ)0.5] (1),
where Aand φ are amplitude and phase of the line shape, and Eg and Γ are the energy and broadening parameter for the direct band gap feature of α-Bi2O3. The fits yield direct transition energies are indicated with arrows in Fig. 3(a). The obtained values of direct band gap for the α-Bi2O3 nanowires are 2.91 eV at 300 K and 3.144 eV at 30 K, respectively. The value of direct gap at 300 K is close to the direct absorption edge of α-Bi2O3 measured by optical absorption [5,7]. The temperature-dependent TR spectra in Fig. 3(a) reveal energy red shift and line width broadened character with the increase of temperatures from 30 to 300 K such as the general semiconductor behavior. Displayed in Figs. 3(b) and 3(c) are, respectively, the temperature-dependent band gaps Eg(T) and broadening parameters Γ(T) for α-Bi2O3 nanowires with representative error bars. The solid line in Fig. 3(b) is the fitting result using a Bose-Einstein expression Eg(T) = EB-aB·{1 + 2/[exp(ωB/kT)-1]}, where aB represents the electron-phonon interaction and ωB is the average phonon energy [18]. The obtained values of fitting parameters for the direct band gap Eg(T) are EB = 2.980 ± 0.003 eV, aB = 160 ± 30 meV, and ωB = 23 ± 3 meV, respectively. Also displayed with hollow triangles in Fig. 3(c) are the temperature-dependent broadening parameters Γ(T) of the TR features for α-Bi2O3. The solid line is a least-square fit to a Bose-Einstein type equation appropriate for line-width broadening expressed as Γ(T) = Γ0 + ΓLO/[exp(ωLO/kT)-1], where Γ0 is invoked from the mechanisms of impurity, dislocation, electron interaction and Auger processes. ΓLO is caused by electron-longitudinal optical (LO) phonon (Fröhlich) interaction and ωLO is the LO phonon energy. The obtained results of fitting parameters in Fig. 3(c) are Γ0 = 76 ± 2 meV, ΓLO = 158 ± 30 meV, and ωLO = 23 ± 3 meV, respectively. It is reasonable that the analyses of temperature-dependent band gaps and fitting parameters in Figs. 3(b) and 3(c) show comparable electron-phonon interaction strength (i.e. aB ≅ ΓLO) and equivalent average phonon energy (i.e. ωBωLO). The value of Γ0 ≅ 76 meV is much larger than that of a bulk single crystal such as ReS2 with Γ0 = 7.8 ± 1.0 meV [19] due to certain defects like oxygen vacancies (VO) existed inside the oxide nanowires. The temperature-dependent relationship of direct band gap Eg(T) derived from Fig. 3(b) can also be the estimate of transition energies of α-Bi2O3 when the nanowires operated at higher temperatures above 300 K.

 figure: Fig. 3

Fig. 3 (a) Temperature-dependent TR spectra of α-Bi2O3 nanowires between 300 and 30 K near band edge. (b) Temperature dependence of direct band gaps of α-Bi2O3 from 30 to 300 K. (c) Temperature-dependent broadening parameter of the TR feature for the direct band gap derived from spectral analysis in (a). The solid line is fitted to a Bose-Einstein expression containing electron-phonon interaction.

Download Full Size | PDF

To evaluate below- and near-band-edge electronic structure of α-Bi2O3 nanowires, PL measurements of 30 and 300 K are respectively carried out and the PL spectra are shown in the lower part of Fig. 4 with dashed line (30 K) and solid line (300 K). For comparison purpose, the corresponding TR spectra at 30 and 300 K are also included in the upper part of Fig. 4 as a reference. The energy position of direct band gap Eg (i.e. 2.91 eV at 300 K and 3.144 eV at 30 K) measured by TR is close to a shoulder peak in the corresponding PL spectrum at 300 and 30 K. The shoulder peak may correlate with direct band-edge emission coming from Eg of α-Bi2O3 nanowires. It also verifies the direct band-edge character of α-Bi2O3. A broadened PL peak Ed that centered at ~2.53 eV at 300 K and ~2.645 eV at 30 K was also detected in the PL spectra in Fig. 4. The broadened peak may be a defect luminescence (Ed) comprising a lot of defect states (by oxygen vacancy VO) to valence band emissions. The oxygen vacancies in the oxide nanostructures usually form a defect donor band such as that observed in the other oxide nanostructure of Ga2O3 [15]. The VO states in Bi2O3 may also have the possibility to contain Bi+, Bi2+, or Bi3+ centers, which can emit visible luminescences from red to blue region [20]. The defect emission Ed in Fig. 4 therefore appears in a broadened emission peak with wider line width. As shown in Fig. 4, the temperature-energy shift of Ed between 30 and 300 K is ~115 meV, which is approximately one half of the energy separation (~230 meV) of band-edge emission Eg. This result verifies that Eg is a band-to-band transition and Ed is a defect to valence-band recombination. The temperature insensitive energy shift (only one half of Eg) is a general character for a defect transition with dangling bond such as the oxygen vacancies in the crystals. The representative band-edge scheme of α-Bi2O3 nanowires is also depicted in the inset of Fig. 4 for comparison. The α-Bi2O3 is a direct semiconductor, which can easily emit and absorb a photon with hν = Eg. The defect luminescence (hν = Ed) from the α-Bi2O3 nanowire may be a broadened emission band caused by defect donor band consisted of VO [15]. Both TRand PL results show α-Bi2O3 nanowire not only a visible-active semiconductor but also a white-light luminescent material with broad-band defect emissions.

 figure: Fig. 4

Fig. 4 The PL and TR spectra of α-Bi2O3 nanowires at 300 and 30 K. The inset depicts a representative band-edge scheme of α-Bi2O3 nanowire derived from the experimental analysis of TR and PL results.

Download Full Size | PDF

4. Conclusions

We prove and evaluate, for the first time, the accurate direct band-edge nature of α-Bi2O3 thin-film nanowires grown by vapor transport method. TEM and Raman spectroscopy confirm the α-monoclinic phase of the nanowire with an axial direction grown along <1¯22>. TR measurement determines a direct band gap of Eg = 2.91 eV for the α-Bi2O3 nanowires at 300 K. Temperature dependences of energies and broadening parameters of direct band gap in α-Bi2O3 verify optoelectronic semiconducting properties of the oxide nanowires. The temperature-energy shift of the direct band gap has been characterized. Both PL and TR measurements verify direct band-edge nature of the α-Bi2O3 nanowires, and a broad-band white light luminescence occurs owing to a defect donor band caused by oxygen vacancies may exist in the oxide nanowires. On the basis of PL and TR results, a probable near-band-edge scheme of α-Bi2O3 nanowires is proposed. It shows that α-Bi2O3 nanostructure is a visible active semiconductor, which may also be applied in solid-state white lightening optoelectronics.

Acknowledgments

The authors would like to acknowledge the financial support from the National Science Council of Taiwan under the grant Nos. NSC 101-2221-E-011-052-MY3 and NSC 99-2112-M-259-006-MY3.

References and links

1. M. Drache, P. Roussel, and J.-P. Wignacourt, “Structures and oxide mobility in Bi-Ln-O materials: heritage of Bi2O3.,” Chem. Rev. 107(1), 80–96 (2007). [CrossRef]   [PubMed]  

2. M. Mehring, “From molecules to bismuth oxide-based materials: Potential homo- and heterometallic precursors and model compounds,” Coord. Chem. Rev. 251(7-8), 974–1006 (2007). [CrossRef]  

3. H. Cheng, B. Huang, J. Lu, Z. Wang, B. Xu, X. Qin, X. Zhang, and Y. Dai, “Synergistic effect of crystal and electronic structures on the visible-light-driven photocatalytic performances of Bi2O3 polymorphs,” Phys. Chem. Chem. Phys. 12(47), 15468–15475 (2010). [CrossRef]   [PubMed]  

4. Y. Qiu, M. Yang, H. Fan, Y. Zuo, Y. Shao, Y. Xu, X. Yang, and S. Yang, “Nanowires of α- and β-Bi2O3: phase-selective synthesis and application in photo catalysis,” CrystEngComm 13(6), 1843–1850 (2011). [CrossRef]  

5. T. Saison, N. Chemin, C. Chanéac, O. Durupthy, V. Ruaux, L. Mariey, F. Mauge, P. Beaunier, and J.-P. Jolivet, “Bi2O3, BiVO4, and Bi2WO6: impact of surface properties on photocatalytic activity under visible light,” J. Phys. Chem. C 115(13), 5657–5666 (2011). [CrossRef]  

6. C. H. Ho, M. C. Tsai, and M. S. Wong, “Characterization of indirect and direct interband transitions of anatase TiO2 by thermoreflectance spectroscopy,” Appl. Phys. Lett. 93(8), 081904 (2008). [CrossRef]  

7. M. Muruganandham, R. Amutha, G.-J. Lee, S.-H. Hsieh, J. J. Wu, and M. Sillanpää, “Facile fabrication of tunable Bi2O3 self-assembly and its visible light photocatalytic activity,” J. Phys. Chem. C 116(23), 12906–12915 (2012). [CrossRef]  

8. A. Hameed, T. Montini, V. Gombac, and P. Fornasiero, “Surface phases and photocatalytic activity correlation of Bi2O3/Bi2O4-x nanocomposite,” J. Am. Chem. Soc. 130(30), 9658–9659 (2008). [CrossRef]   [PubMed]  

9. C. H. Ho, C. H. Chan, L. C. Tien, and Y. S. Huang, “Direct optical observation of band edge excitons, band gap, and Fermi level in degenerate semiconducting oxide nanowires In2O3,” J. Phys. Chem. C 115(50), 25088–25096 (2011). [CrossRef]  

10. C. H. Ho, “Enhanced photoelectric-conversion yield in niobium-incorporated In2S3 with intermediate band,” J. Mater. Chem. 21(28), 10518–10524 (2011). [CrossRef]  

11. H. W. Kim, J. H. Myung, S. H. Shim, and C. Lee, “Growth of Bi2O3 rods using a trimethylbismuth and oxygen mixture,” Appl. Phys., A Mater. Sci. Process. 84(1-2), 187–189 (2006). [CrossRef]  

12. J. In, I. Yoon, K. Seo, J. Park, J. Choo, Y. Lee, and B. Kim, “Polymorph-tuned synthesis of α- and β-Bi2O3 nanowires and determination of their growth direction from polarized Raman single nanowire microscopy,” Chemistry 17(4), 1304–1309 (2011). [CrossRef]   [PubMed]  

13. B. Ling, X. W. Sun, J. L. Zhao, Y. Q. Shen, Z. L. Dong, L. D. Sun, S. F. Li, and S. Zhang, “One-dimensional single-crystalline bismuth oxide micro/nanoribbons: morphology-controlled synthesis and luminescent properties,” J. Nanosci. Nanotechnol. 10(12), 8322–8327 (2010). [CrossRef]   [PubMed]  

14. V. N. Denisov, A. N. Ivlev, A. S. Lipin, B. N. Mavrin, and V. G. Orlov, “Raman spectra and lattice dynamics of single-crystal α-Bi2O3,” J. Phys. Condens. Matter 9(23), 4967–4978 (1997). [CrossRef]  

15. C. H. Ho, C. Y. Tseng, and L. C. Tien, “Thermoreflectance characterization of β-Ga2O3 thin-film nanostrips,” Opt. Express 18(16), 16360–16369 (2010). [CrossRef]   [PubMed]  

16. D. E. Aspnes, in Handbook on Semiconductors, edited by M. Balkanski, (North Holland, 1980).

17. F. H. Pollak and H. Shen, “Modulation spectroscopy of semiconductors: bulk/thin film, microstructures, surfaces/interfaces and devices,” Mater. Sci. Eng. R10, 275–374 (1993).

18. P. Lautenschlager, M. Garriga, S. Logothetidis, and M. Cardona, “Interband critical points of GaAs and their temperature dependence,” Phys. Rev. B Condens. Matter 35(17), 9174–9189 (1987). [CrossRef]   [PubMed]  

19. C. H. Ho, P. C. Liao, Y. S. Huang, and K. K. Tiong, “Temperature dependence of energies and broadening parameters of the band-edge excitons of ReS2 and ReSe2,” Phys. Rev. B 55(23), 15608–15613 (1997). [CrossRef]  

20. S. Zhou, N. Jiang, B. Zhu, H. Yang, S. Ye, G. Lakshminarayana, J. Hao, and J. Qiu, “Multifunctional bismuth-doped nanoporous silica glasses: from blue-green, orange, red, and white light sources to ultrabroadband infrared amplifiers,” Adv. Funct. Mater. 18(9), 1407–1413 (2008). [CrossRef]  

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 (4)

Fig. 1
Fig. 1 SEM images of (a) top view and (b) cross-section view for as-grown α-Bi2O3 nanowires on Si substrate. (c) TEM image of a single nanowire with diameter of ~298 nm, the corresponding EDS spectrum and the content data are shown in (d). (e) High-resolution TEM image of the α-Bi2O3 nanowire, the interplannar distance of d(001) is indicated. (f) The related SAED pattern of the α-Bi2O3 nanowire with zone axis along [210].
Fig. 2
Fig. 2 Raman peak pattern of α-Bi2O3 thin-film nanowires with the vibration frequencies between 100 and 700 cm−1.
Fig. 3
Fig. 3 (a) Temperature-dependent TR spectra of α-Bi2O3 nanowires between 300 and 30 K near band edge. (b) Temperature dependence of direct band gaps of α-Bi2O3 from 30 to 300 K. (c) Temperature-dependent broadening parameter of the TR feature for the direct band gap derived from spectral analysis in (a). The solid line is fitted to a Bose-Einstein expression containing electron-phonon interaction.
Fig. 4
Fig. 4 The PL and TR spectra of α-Bi2O3 nanowires at 300 and 30 K. The inset depicts a representative band-edge scheme of α-Bi2O3 nanowire derived from the experimental analysis of TR and PL results.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

ΔR R =Re[ A e jφ (E E g +jΓ) 0.5 ]
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.