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

In situ micro-Raman studies of laser-induced bismuth oxidation reveals metastability of β-Bi2O3 microislands

Open Access Open Access

Abstract

We report the laser irradiation-induced oxidation of bismuth metal investigated in situ by micro-Raman spectroscopy as a function of irradiation power and time. The purely optical synthesis and characterization of β-Bi2O3 oxide microislands on metallic Bi surfaces is shown to be stable over time, even at room-temperature. By closely examining possible reactions on simple Bi morphologies it is revealed for the first time that the ensuing oxide phase is critically dependent on the final oxide volume and follows a fixed kinetic transformation sequence: 32O2(g)+2Bi(l)βBi2O3(s)αBi2O3(s). These findings are unusual within the framework of traditional Bi2O3 thermal transformation relations. An electrostatic mechanism involving a changing Bi2O3 surface-to-volume ratio is proposed to explain the room-temperature metastability of small β-Bi2O3 volumes and the subsequent transformation sequence, as well as unifying the results of previous studies.

© 2014 Optical Society of America

1. Introduction

Bismuth trioxide (Bi2O3; also known as bismuth oxide and bismuthsesquioxide) polymorphs are fascinating optical materials with wide bandgap, high refractive index, large dielectric permittivity, remarkable photoconductivity, and ionic conductivity. The versatility of Bi2O3 nanomaterials has fostered considerable attention for scientific and technological advances over the past decade, with applications spanning electrical ceramics, solid-state electrolytes, gas sensors, medical disinfection/sterilization, and energy generation [14]. Within the context of current research on Bi2O3, surprisingly little is known about its phase properties at the nano- and micron-scale.

Of the four main Bi2O3 polymorphs [58] by far the most interesting are the high temperature metastable β and δ modifications, because of their extremely high oxygen ion conductivity (δ-Bi2O3 possessing the highest among all binary metal sesquioxides) and their strong visible-light-driven photocatalytic activity (β-Bi2O3 being the most active). While thermal oxidation of bismuth is inexpensive and easily controlled [911], synthesis of the desirable β-Bi2O3 and its stabilization at room temperature is not straightforward [1216]. Attempts to better understand phase transformation relations among Bi2O3 polymorphs are currently underway [17], however a comprehensive model of the transient stages of oxidation is yet to fully explain the varied and seemingly anomalous oxidation phase results reported in the literature [11, 1829]. From the nucleation of Bi2O3 molecules on the bismuth surface to the coalescence of a thermodynamically stable oxide represents a scientifically challenging and technologically important terra incognita. In most studies of bismuth oxidation to date, the only real-time in situ diagnostics have been through thermogravimetric analysis or X-ray diffraction (XRD) measurements. Usually, neither of these techniques is versatile enough to provide information suitable for process control or for elucidating transformation dynamics [3033].

In this paper, we shed light on the bismuth oxidation process through novel laser irradiation-induced oxidation experiments, monitored in situ by micro-Raman scattering. Our findings reveal that the resultant Bi2O3 phase is critically dependent on oxide volume (at the micron-scale) and follows a fixed transformation sequence. An electrostatic mechanism – based on a changing Bi2O3 surface-to-volume ratio – is invoked to explain how small volumes of un-doped β-Bi2O3 acquire room-temperature metastablity, as well as permitting the interpretation of previous studies.

2. Experiment

2.1. Sample details

Bismuth samples were prepared using a high purity (99.99% pure) bismuth ingot. Flakes scraped off by scalpel allowed measurements on larger (bulk) samples, while the production of smaller bismuth particles (0.3 – 3 μm) was achieved using a mortar and pestle, followed by selection under microscope. No further sample preparation was required.

2.2. Raman scattering experiments

Room-temperature micro-Raman spectra were acquired in a quasi-backscattering configuration using a Jobin-Yvon HR800 integrated micro-Raman system with confocal microscope, 20 mW HeNe 632.8 nm laser, and air-cooled CCD detector. Laser spot diameters of 1.7 μm and 4.2 μm were defined by use of either an Olympus ×100 or ×50 objective, respectively. Spot size was determined using a combination of highly-attenuated optical images and charting across a cleaved Si crystal edge. Laser power densities were controlled by a neutral variable density filter incorporated before the microscope optics. Dispersion was achieved through a 1800 g/mm diffraction grating (spectral resolution 0.2 cm−1) and instrument calibration was verified through checking the position of the Si band at ±520.7 cm−1. Implementing an appropriate notch filter to limit Rayleigh backscatter detection meant low-frequency portions of the Raman spectra (below 58 cm−1) are partially truncated. All spectra are presented as raw data (no background subtracting or smoothing has been performed) and are recorded using the Raman laser power densities indicated in corresponding figure captions.

3. Results and discussion

The melting point of bismuth is much lower than that of Bi2O3 (see Fig. 3), so atomic rearrangement immediately after oxide formation is inhibited. The tendency for Bi2O3 to be glass-forming on metallic Bi surfaces [34] means it is useful to investigate small-scale surface reactions. Figure 1 shows optical microscope images of the three kinds of bismuth surfaces examined in this study, along with a typical bismuth Raman spectrum exhibiting two modes at 71 cm−1 and 98 cm−1, which are consistent with the two first-order optical bands of rhombohedral bismuth corresponding to doubly degenerate Eg and non-degenerate A1g phonon modes, respectively [35]. Enlarged in the inset are the weak second-order bands at ∼188 cm−1 consisting of three overtones of similar frequencies [36].

 figure: Fig. 1

Fig. 1 Optical micrographs of representative Bi surfaces examined in this study. (a) Smooth, (b) rough, and (c) small particulates on a glass slide. (d) The micro-Raman spectra measured at very low laser power density (5×102 W· cm−2) from a smooth Bi surface with the inset expanded over the second-order bismuth harmonics.

Download Full Size | PDF

The development of laser-induced oxidation on a smooth Bi surface is plotted in Fig. 2(a) for increasing laser power densities from 0.6 × 104 to 12.5 × 104 W· cm−2, where additional modes at 128, 315, and 461 cm−1 contribute to the Raman lineshape for irradiances in excess of 2.7 × 104 W· cm−2. These three bands correspond to unique Bi-O stretches and are attributed to the β-phase Bi2O3 [37]. Their appearance confirms laser irradiation-induced oxidation of the bismuth surface (using ambient oxygen) and indicates that, for sufficiently large power densities, there are two induced effects: an oxidation reaction followed by the arranging into β-Bi2O3. The second-order scattering mode (with heavy A1g phonon composition [36]) also intensifies and appears connected to the surface oxidation; however, it is not. As the laser power increases the probed area rises in temperature and causes first-order bands to grow. Consequently, this also increases the intensity of the second-order harmonics. Figure 2(b) presents similar data for a decreasing probe irradiance. Here the β-Bi2O3 signatures represent a prominent and unchanging contribution to the overall Raman lineshape. At the relatively low power density of 0.5 × 104 W· cm−2, where temperature estimates based on non-degenerate A1g red-shifts place the bismuth crystal close to room-temperature [38, 39], no changes are observed.

 figure: Fig. 2

Fig. 2 Raman spectra of smooth bismuth surface (see Fig. 1(a)) for (a) increasing laser power, and (b) decreasing laser power. The insets expand these data over the first-order bismuth optical phonon range. For decreasing power, time was spent between each measurement to allow for cooling. The three solid vertical lines at 128, 315, and 461 cm−1 represent the three modes of β-Bi2O3. Spectra have been scaled and shifted vertically for clarity.

Download Full Size | PDF

Given 632.8 nm HeNe laser light is used for excitation, the photolytically activated influences are small compared to the isothermally driven process [40]. Many metal oxidation reactions promoted by laser irradiation are largely thermally driven, and isolating non-thermal radiation effects is difficult. For the specific case of a specular bismuth surface [41] irradiated by low intensity 632.8 nm laser light, Raman measurements demonstrated no evidence of oxidation following long exposures; only once temperatures in the laser heated volume surpass the bismuth melting point will an oxidation reaction occur. This is verified by the insets of Figs. 2(a) and 2(b), where phonon mode redshifting is attributed to large changes in local temperature [38,39]. Decreasing laser powers reveal these phonon bands returning to their original energies. This is important, as it removes factors such as laser-induced defects, resulting in quantum confinement, from influencing the observed redshift; the shifting is driven by nondestructive and purely thermal mechanisms.

The stability of the transformed β-Bi2O3 material was investigated over time. Raman spectra were recorded before and after exposure to high irradiance and in both cases an extremely low power density of IR < 5 × 102 W· cm−2 was used. The absence of Bi-O stretches before irradiation and their presence two weeks after irradiation (spectra not shown) reveal a stable synthesis of β-Bi2O3. Furthermore, no decrease in the relative Raman scattering intensities of these peaks was observed. Thus laser-induced β-Bi2O3 oxidation is not only demonstrated, but shown to be stable at room temperature.

Figure 3 displays the traditional existence domains and transformation relations (broken light arrows) for the four main Bi2O3 polymorphs [58], represented as α- (monoclinic), β- (tetragonal), γ- (bcc), and δ-phase (fcc), along with the oxidation sequence observed in this study (solid heavy arrows). The existence of at least three other modifications have been realized (ω, ε and η), although these are synthesized under atypical conditions [4244] and are omitted for clarity. A striking feature of Fig. 3 is a tendency for all polymorphs to transform into the δ-phase at sufficiently high temperatures, and ultimately to transform into the α-phase when cooled to room temperature. Of the phases, α-Bi2O3 is thermodynamically stable from room temperature up to 1003 K, where it transforms into δ-Bi2O3 phase, which is stable until melting at 1097 K. When cooling from δ-Bi2O3, large thermal hysteresis effects regulate the transformation into one of two high-temperature metastable phases, β-Bi2O3 or γ-Bi2O3. Below approximately 641–923 K the metastable phase transforms into α-Bi2O3, unless chemically stabilized [12]. Within this framework, our findings are unusual for two reasons. Firstly, the laser-induced oxidation results reveal a purely β-Bi2O3 oxide layer. This is directly verified through no observation of α-Bi2O3 vibrations (see Fig. 4(a)) in the Raman spectra. Secondly, the laser-induced β-Bi2O3 violates the traditional temperature domains for Bi2O3 by exhibiting phase stability at room temperature.

 figure: Fig. 3

Fig. 3 Existence domains and transformation relations of Bi and the four main Bi2O3 polymorphs [1, 7]. The light broken arrows represent bulk isotherm transformations, while the heavy solid arrows represent the micro-probe laser-induced transformations observed in this study. The processes of heating and cooling are indicated here by Δ and ∇, respectively. Temperature ranges without parentheses represent transition temperatures while the values in parentheses represent required minimum initial temperatures.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 (a) Representative unpolarized Raman spectra from bismuth metal and resultant laser-induced Bi2O3 phases. Traces labeled ome-4-10-2133-i001.JPG, ome-4-10-2133-i002.JPG and ome-4-10-2133-i003.JPG represent the pure phase spectra, while the spectra labeled ome-4-10-2133-i004.JPG portray an initial (and sometimes final) combined phase. (b) Chosen ome-4-10-2133-i005.JPG transitional dual spectra for longer irradiation exposures. Arrows between traces indicate the observed oxidation phase sequence and all spectra have been normalized (normalization factors given) and offset for clarity. (c) Schematic of oxide aggregation leading to ome-4-10-2133-i005.JPG transition.

Download Full Size | PDF

To understand these results, we next investigate the laser-induced changes which are possible with the more textured Bi surfaces of Figs. 1(a) – 1(c). Laser-induced oxidations were carried out using a maximum irradiance of 12.5 × 104 W· cm−2 and 3.2 × 104 W· cm−2 for 1.7 μm and 4.2 μm laser spots, respectively, and spectra recorded at fixed time intervals. Final resultant phases were determined in situ only after temporal changes in the Raman spectra ceased. Due to the micro-beam profile, laser spot sizes are not equal to the size of the oxide formed, but represent the size of the probed area. Depending on incident power densities and surface morphology, the laser-induced oxides can be much smaller.

Plotted in Fig. 4(a) are representative Raman spectra of bismuth [35] and the two single β- [45] and α-phase [46] oxides achievable with our experimental setup. The arrows between traces here indicate observed sequential intermediates. Optical imaging of pure α-phase oxidation (not shown) exhibited the characteristic light yellow color associated with α-Bi2O3. While a laser-induced δ-phase transformation was recently realized [30] in Bi2O3, the characteristic 625 cm−1 Raman signature was never observed in our data. A stunning aspect of these results is that we observe an initial β-phase (or ome-4-10-2133-i004.JPG) in all transformations; laser-induced oxidation of α-Bi2O3 is never achieved without first passing through an intermediate β-phase. This is highlighted by the temporal data shown in Fig. 4(b), where for longer exposure times the ome-4-10-2133-i002.JPG Raman spectra is shown to evolve into single-phase ome-4-10-2133-i003.JPG (these spectra have been normalized to the bottom ome-4-10-2133-i002.JPG trace by the factors given). The phase transition is matched by broadband temporal growth of the Raman signal, increasing by a factor of approximately 15. The increase in Raman signal here is evenly weighted for both existing β-Bi2O3 modes and transitioning α-Bi2O3 modes (for example, the peak just above 300 cm−1 undergoes a continuous intensification during the transformation), indicating a dramatic change in the scattering volume and the formation of new oxide material, rather than a pronounced shift in scattering cross-section. The tipping point between phases appears to be connected to the size of Bi2O3 produced and is schematically represented in Fig. 4(c); after some well-defined critical volume, the Bi2O3 transforms its structure from β-phase to α-phase.

Table 1 summarizes the laser-induced oxidation results for all bismuth surface types examined. Here ‘smooth’ corresponds to a highly specular surface (Fig. 1(a)) and ‘rough’ represents a granulated surface on the scale of 0.2 – 4 μm (Fig. 1(b)). Smaller particles ranging from roughly 0.3 to 3 μm were also examined (Fig. 1(c)). Differences in the relative texture of the particulates were insignificant (when compared to size) in determining oxide phase and are not considered further. The oxidation results in Table 1 suggest an inability for Bi2O3 to retain β-phase structure above the critical dimension (∼1.5 μm). This is highlighted by a complete absence of α-Bi2O3 when irradiating with a 1.7 μm laser spot. This is a consequence of the small (< 1.7 μm) and limited volume of oxides formed when employing this beam spot size. Raman spectra revealing a combination of ome-4-10-2133-i001.JPG and ome-4-10-2133-i002.JPG were most favored, with the notable exception of using a 4.2 μm beam on smaller particles. To this point, measurements made with the larger spot size proved to be a better gauge for assessing size dependencies; a distinct shift in subsequent phase was observed for particles transitioning across ∼1.5 μm range. Furthermore, rough surfaces with elevated structure were more likely to produce α-Bi2O3.

Tables Icon

Table 1. Laser-induced bismuth oxide phase results for varying bismuth surfaces (corresponding to those shown in Figs. 1(a)–1(c)) irradiated by a large (4.2 μm) and small (1.7 μm) focused 632.8 nm laser beam. Representative Raman spectra of resulting phase(s) match with those given in Fig. 4(a).

Given the 632.8 nm laser-induced oxidation process is thermally driven, the observation of a new governing oxidation/transformation sequence

32O2(g)+2Bi(l)βBi2O3(s)αBi2O3(s),
extends the tradtional phase relations of Fig. 3 and should hold true for all thermal oxidations of bismuth. It is emphasized that a laser-induced β-Bi2O3α-Bi2O3 transition occurs in direct opposition to conventional isothermal existence domains, since longer laser exposures generally increase local temperatures rather than decrease them.

Examining the varied Bi oxidation studies in the literature we make the observation that the sequence described by Eq. 1 is both accurate for high temperature annealing methods [1823] and laser-induced techniques [11, 2426]. For the specific cases of thermally-treated nanoparticles [18, 21] and nanotextured films [22, 23], lower oxidation temperatures produced single-phase β-Bi2O3, while higher temperatures resulted in single-phase α-Bi2O3. More importantly, scanning electron microscope (SEM) images revealed an increased tendency for bismuth oxide microislands to coalesce and form larger aggregated oxide volumes (see Fig. 4(c)) with increasing temperature, depicting a critical oxide size comparable to ours, for which the β-Bi2O3α-Bi2O3 transition is observed. It was even demonstrated by XRD measurements that for temperatures between the synthesis of pure β-phase or α-phase oxides, a transitional mixed phase is observed – reminiscent of Fig. 4(b) – which bridged the two single-phases [22]. Huang et al. [23] also inadvertently reported similar results, however, accounting for their findings using a model based on the differing crystallographic orientations of electrodeposited Bi films. In the pursuit of laser-induced oxidation of Bi nanowires [26] and nanotextured films [11,24], several authors have pointed to the role laser power plays in determining the resulting oxide phase(s). In brief, all studies demonstrated an initial β-phase and oxidations under higher irradiance favored α-Bi2O3 formation. Similar to elevated temperatures in thermal annealing processes, the increased α-Bi2O3 formation demonstrated an enhanced capacity for the laser-heated area to produce larger oxides [11, 24]. The consistency of results across the very different oxidation schemes means oxide/metal interface effects (i.e. mismatch or strain relief/dislocations) can be neglected. In light of these findings, the one feature which unifies the synthesis of room-temperature metastable β-Bi2O3 becomes clear: size.

The fundamental relationships between bismuth oxidation and ensuing phases are still unresolved. For example, why is an initial β-phase favorable over an α-phase and how does undoped β-Bi2O3 oxide remain stable at room temperature? From the view point of joining small β-Bi2O3 volumes to form an aggregated α-phase, a likely explanation lies in their respective molecular structures; the β-phase Bi2O3 polyhedra are proposed to be BiO6 octahedra with all Bi-O bonds equivalent at 2.4Å and very similar to that of α-Bi2O3. However, given the volumic nature of results, a detailed answer presumably resides in how the electrostatics of the ionic Bi2O3 crystal change with an evolving surface-to-volume ratio [47].

The electrostatic energy of a unit cell deep within the interior of a Bi2O3 crystal is composed of two parts. One part is morphology-independent, and the other is morphology-dependent and varies with the structure/phase concerned and the distribution of ions within a unit cell. Morphology dependencies are conditional to crystal shape and are equal to zero if the dipole moment of the unit cell zero. In Tasker’s model of ionic crystal surfaces [48], the surface of Bi2O3 has a dipole moment perpendicular to the surface which is exposed favorably in a practical environment. Moreover the surface is electrostatically unstable. The electrostatic energy of the whole Bi2O3 volume therefore is the electrostatic energy of a unit cell multiplied by the number of unit cells plus the corrections proportional to the surface area of the crystal [49]. Thus, the bulk energy of diminutive Bi2O3 is heavily dependent on surface-to-volume ratio. Harwig and Weenk [50] have already pointed out that there is only a very small difference in the electrostatic energy of α-phase and β-phase Bi2O3, with a Madelung constant marginally lower in the former. The surface stability has also been shown to relate to the stability of different crystal planes and, if polar or nonpolar terminations are present, due to alternate stacking of Bi layers and O layers. However, since α-Bi2O3 is never seen in the initial oxidation of bismuth, it is not likely that crystal-plane effects are significant.

The temporal data shown in Fig. 4(b) permits the evaluation of real-time phase and structure, as well as allowing the general assessment of mode shifting magnitude and direction. Through a purely empirical interpretation of phonon frequencies, the Bi-O stretching frequency is inversely related to the bond lengths [37]. While α-phase nucleation points take hold in the β-Bi2O3 volume, new interactions are established between Bi and O atoms in the transitioning oxide. Early in the transformation, new bonds formed within the oxide experience very little of the pressures exerted by a surrounding external network. New vibrational frequencies are set up and appear at lower frequencies due to this lack of internal pressure and longer bond lengths. With the notable exception of the Bi-Bi vibration just below 70 cm−1, all modes appear to strongly harden to higher frequencies for longer laser exposure times. This is paralleled by the entire spectra intensifying over the course of the transformation, until a pure α-phase is reached. It is reasonable to assume, therefore, a correlation between an increasing oxide volume and the direction of the transformation. Rising pressure introduced by a growing crystal network reduces bond lengths and results in phonon mode hardening. As the budding α-Bi2O3 condenses, the electrostatic potential of unit cells is further reduced and the dipole moment is lowered toward a final equilibrium.

For reasons that are at present unclear, the synthesis and stabilization of β-Bi2O3 is much easier than that of γ-Bi2O3. This suggests β-Bi2O3 has a strong tendency to remain stable when its electrostatic energy occupies a local minimum. As far as we know, there is little information about the relative stability, geometries, and electronic structures of Bi2O3 polymorph surfaces, which hinders the understanding of their properties and reactivity at small scales. While a comprehensive physical model to aid explanation is lacking, our results and conclusions present merely a plausibility argument; one that points toward oxide volume at the micron-scale as playing a central role in defining Bi2O3 phase. This aspect of bismuth oxidation has thus far been overlooked and is particularly significant. In terms of engineering practical Bi2O3 elements, fine control over the transformation process is favorable to the exacting preparation of functional materials with specific crystal structure.

4. Conclusion

In conclusion, we have elucidated the dynamics of Bi oxidation through micro-Raman scattering experiments. Our optical results reveal the unexpected and seemingly fixed bismuth oxidation/transformation sequence: 32O2(g)+2Bi(l)βBi2O3αBi2O3. The phenomenon of room-temperature metastable β-Bi2O3 is examined within this sequence and the tipping point is attributed to Bi2O3 volume dependencies. We invoke an electrostatic model based on the changing Bi2O3 surface-to-volume ratio to account for our findings and also to unlock the interpretation to previous studies. Considering the wide variety of applications for bismuth oxide nanomaterials, these findings represent a timely stimulus to further theoretical and experimental investigation of bismuth oxidation, as well as possible methods of material synthesis/fabrication.

Acknowledgments

We thank the Australian Research Council for support and the Australian National Fabrication Facility (ANFF) - Materials Node for their provision of research facilities.

References and links

1. N. M. Sammes, G. A. Tompsett, H. Näfea, and F. Aldingera, “Bismuth based oxide electrolytes – structure and ionic conductivity,” J. Eur. Ceram. Soc. 19(10), 1801–1826 (1999). [CrossRef]  

2. J. W. Fergus, “Electrolytes for solid oxide fuel cells,” J. Power Sources 162(1), 30–40 (2006). [CrossRef]  

3. H. J. Fan, P. Werner, and M. Zacharias, “Semiconductor nanowires: from self-organization to patterned growth,” Small 2(6), 700–717 (2006). [CrossRef]   [PubMed]  

4. Q. Hu, J. Wang, Y. Zhao, and D. Li, “A light-trapping structure based on Bi2O3 nano-islands with highly crystallized sputtered silicon for thin-film solar cells,” Opt. Express 19(S1), A20–A27 (2011). [CrossRef]  

5. H. A. Harwig, “On the structure of bismuthsesquioxide: the α, β, γ, and δ-phase,” Z. Anorg. Allg. Chem. 444(1), 151–166 (1978). [CrossRef]  

6. M. Miyayama, S. Katsuta, Y. Suenaga, and H. Yanagida, “Electrical conduction in β-Bi2O3 doped with Sb2O3,” J. Am. Ceram. Soc. 66, 585 (1983). [CrossRef]  

7. H. A. Harwig and A. G. Gerards, “Electrical properties of the α, β, γ, and δ phases of bismuth sesquioxide,” J. Solid State Chem. 26(3), 265–274 (1978). [CrossRef]  

8. J. W. Medermach and R. L. Snyder, “Powder diffraction patterns and structures of the bismuth oxides,” J. Am. Ceram. Soc. 61(11–12), 494–497 (1978). [CrossRef]  

9. L. Leontie, M. Caraman, M. Alexe, and C. Harnagea, “Structural and optical characteristics of bismuth oxide thin films,” Surf. Sci. 507–510, 480–485 (2002). [CrossRef]  

10. R. A. Ismail, “Characteristics of bismuth trioxide film prepared by rapid thermal oxidation,” Surf. Sci. Nanotech. 4, 563–565 (2006). [CrossRef]  

11. L. Kumari, J.-H. Lin, and Y.-R. Ma, “Laser oxidation and wide-band photoluminescence of thermal evaporated bismuth thin films,” J. Phys. D 41(2), 025405 (2008). [CrossRef]  

12. T. B. Tran and A. Navrotsky, “Energetics of disordered and ordered rare earth oxide-stabilized bismuth oxide ionic conductors,” Phys. Chem. Chem. Phys. 16(5), 2331–2337 (2014). [CrossRef]  

13. Y. Wang, Y. Wen, H. Ding, and Y. Shan, “Improved structural stability of titanium-doped β-Bi2O3 during visible-light-activated photocatalytic processes,” J. Mater. Sci. 45(5), 1385–1392 (2010). [CrossRef]  

14. M. Shlesinger, S. Schylze, M. Hietschold, and M. Mehring, “Metastable β-Bi2O3 nanoparticles with high photocatalytic activity from polynuclear bismuth oxide clusters,” Dalton Trans. 42(2), 1047–1056 (2013). [CrossRef]  

15. V. V. Zyryanov, “Mechanochemical synthesis of complex oxides,” Russ. Chem. Rev. 77, 105 (2008). [CrossRef]  

16. V. V. Zyryanov and N. F. Uvarov, “Mechanochemical synthesis and electrical conductivity of Bi1.6M0.4O3x (M = Ca, Ca0.5Sr0.5, In, Y, La) metastable fluorite solid solutions,” Inorg. Mater. 40(7), 729–734 (2004). [CrossRef]  

17. H. -Y. Deng, W.-C. Hao, and H.-Z. Xu, “A transition phase in the transformation from α-, β- and ε- to δ-bismuth oxide,” Chin. Phys. Lett. 28(5), 056101 (2011). [CrossRef]  

18. H. Fitouri, R. Boussaha, A. Rabey, and B. El Jani, “Oxidation of bismuth nanodroplets deposit on GaAs substrate,” Appl. Phys. A 112(3), 701–710 (2013). [CrossRef]  

19. B. J. Yang, M. S. Mo, H. M. Hu, C. Li, X. G. Yang, Q. W. Li, and Y. T. Qian, “A rational self-sacrificing template route to β-Bi2O3 nanotube arrays,” Eur. J. Inorg. Chem. 2004(9), 1785–1787 (2004). [CrossRef]  

20. M. Gotić, S. Popović, and S. Musić, “Influence of synthesis procedure on the morphology of bismuth oxide particles,” Materials Letters 61(3), 709–714 (2007). [CrossRef]  

21. A. J. Salazar-Pérez, M. A. Camacho-López, R. A. Morales-Luckie, and V. Sánchez-Mendieta, “Structural evolution of Bi2O3 prepared by thermal oxidation of bismuth nano-particles,” Superficies y Vacío 18(3), 4–8 (2005).

22. T. P. Gujar, V. R. Shinde, and C. D. Lokhande, “The influence of oxidation temperature on structural, optical and electrical properties of thermally oxidized bismuth oxide films,” Appl. Surf. Sci. 254(13), 4186–4190 (2008). [CrossRef]  

23. C.-C. Huang, T.-Y Wen, and K.-Z. Fung, “Orientation-controlled phase transformation of Bi2O3 during oxidation of electrodeposited Bi film,” Mater. Res. Bull. 41(1), 110–118 (2006). [CrossRef]  

24. M. A. Zepeda, M. Picquart, and E. Haro-Poniatowski, “Laser induced oxidation effects in bismuth thin films,” MRS Proceedings1477, (2012).

25. K. Trentelman, “A note on the characterization of bismuth black by Raman microspectroscopy,” J. Raman Spectrosc. 40(5), 585–589 (2009). [CrossRef]  

26. W. Wang, M. Liu, Z. Yang, W. Mai, and J. Gong, “Synthesis and Raman optical properties of single-crystalline Bi nanowires,” Physica E 44(7–8), 1142–1145 (2012). [CrossRef]  

27. S. Condurache-Bota, G. I. Rusu, N. Tigau, R. Drasozean, and C. Gheorghies, “Structural and optical characterization of thermally oxidized bismuth films,” Rom. J. Chem. 54(3), 205–211 (2009).

28. L. Kumari, J.-H. Lin, and Y.-R. Ma, “One-dimensional Bi2O3 nanohooks: synthesis, characterization and optical properties,” J. Phys.: Condens. Matter 19(40), 406204 (2007).

29. K. J. Stevens, B. Ingham, M. F. Toney, S. A. Brown, J. Partridge, A. Ayeshb, and F. Natalie, “Structure of oxidized bismuth nanoclusters,” Acta Cryst. B63, 569–576 (2007). [CrossRef]  

30. M. Vila, C. Diaz-Guerra, and J. Piqueras, “Laser irradiation-induced α to δ phase transformation in Bi2O3 ceramics and nanowires,” Appl. Phys. Lett. 101(7), 071905 (2012). [CrossRef]  

31. D. Y. Lu, J. Chen, J. Zhou, S. Z. Deng, N. S. Xu, and J. B. Xu, “Raman spectroscopic study of oxidation and phase transition in W18O49 nanowires,” J. Raman Spectrosc. 38(2), 176–180 (2007). [CrossRef]  

32. O. N. Shebanova and P. Lazor, “Raman study of magnetite (Fe3O4): laser-induced thermal effects and oxidation,” J. Raman Spectrosc. 34(11), 845–852 (2003). [CrossRef]  

33. M. A. Camacho-López, L. Escobar-Alarcón, M. Picquart, R. Arroyo, G. Córdoba, and E. Haro-Poniatowski, “Micro-Raman study of the m-MoO2 to α-MoO3 transformation induced by cw-laser irradiation,” Opt. Mater. 33(3), 480–484 (2011). [CrossRef]  

34. M. S. R. Hynes and H. Rawson, “Bismuth trioxide glasses,” J. Soc. Glass Technol. 41, 347–349 (1957).

35. M. G. Mitch, S. J. Chase, J. Fortner, R. Q. Yu, and J. S. Lannin, “Phase transition in ultrathin Bi films,” Phys. Rev. Lett. 67(7), 875–878 (1991). [CrossRef]   [PubMed]  

36. J. S. Lannin, J. M. Calleja, and M. Cardona, “Second-order Raman scattering in the group-Vb semimetals: Bi, Sb, and As,” Phys. Rev. B 12(2), 585–593 (1975). [CrossRef]  

37. F. D. Hardcastle and I. E. Wachs, “The molecular structure of bismuth oxide by Raman spectroscopy,” J. Solid State Chem. 97(2), 319–331 (1992). [CrossRef]  

38. J. L. Yarnell, J. L. Warren, R. G. Wenzel, and S. H. Koenig, “Phonon dispersion curves in bismuth,” IBM J. Res. Dev. 8(3), 234 (1964). [CrossRef]  

39. J. Höhne, U. Wenning, H. Shulz, and S. Hüfner, “Temperature dependence of the k=0 optical phonons of Bi and Sb,” Z. Phys. B 27(4), 297–302 (1977). [CrossRef]  

40. L. Nánai, R. Vajtai, and T. F. George, “Laser-induced oxidation of metals: state of the art,” Thin Solid Films 298, 160–164 (1997). [CrossRef]  

41. M. K. Zayed and H. E. Elsayed-Ali, “Condensation on (002) graphite of liquid bismuth far below its bulk melting point,” Phys. Rev. B. 72(20), 205426 (2005). [CrossRef]  

42. A. F. Gualtieri, S. Immovilli, and M. Prudenzianti, “Powder X-ray diffraction data for the new polymorphic compound omega-Bi2O3,” Powder Diffr. 12(2), 90 (1997). [CrossRef]  

43. N. Cornei, N. Tancet, F. Abraham, and O. Mentré, “New ε-Bi2O3 polymorph,” Inorg. Chem. 45(13), 4886–4888 (2006). [CrossRef]   [PubMed]  

44. T. Atou, H. Faqir, M. Kikuchi, H. Chiba, and Y. Syono, “A new high-pressure phase of bismuth oxide,” Mater. Res. Bull. 33(2), 289–292 (1998). [CrossRef]  

45. 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,” Chem. Eur. J. 17(4), 1304–1309 (2011). [CrossRef]   [PubMed]  

46. 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 (1997).

47. Y.-H Lei and Z.-X. Chen, “Density functional study of the stability of various α-Bi2O3 surfaces,” J. Chem. Phys. 138(5), 054703 (2013). [CrossRef]   [PubMed]  

48. P. W. Tasker, “The stability of ionic crystal surfaces,” J. Phys. C: Solid State Phys. 12(22), 4977 (1979). [CrossRef]  

49. E. R. Smith, “Electrostatic energy in ionic crystals,” Proc. R. Soc. Lond. A 375, 475–505 (1981). [CrossRef]  

50. H. A. Harwig and J. W. Weenk, “Phase relations in bismuthsesquioxide,” Z. Anorg. Allg. Chem. 444(1), 167–177 (1978). [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 Optical micrographs of representative Bi surfaces examined in this study. (a) Smooth, (b) rough, and (c) small particulates on a glass slide. (d) The micro-Raman spectra measured at very low laser power density (5×102 W· cm−2) from a smooth Bi surface with the inset expanded over the second-order bismuth harmonics.
Fig. 2
Fig. 2 Raman spectra of smooth bismuth surface (see Fig. 1(a)) for (a) increasing laser power, and (b) decreasing laser power. The insets expand these data over the first-order bismuth optical phonon range. For decreasing power, time was spent between each measurement to allow for cooling. The three solid vertical lines at 128, 315, and 461 cm−1 represent the three modes of β-Bi2O3. Spectra have been scaled and shifted vertically for clarity.
Fig. 3
Fig. 3 Existence domains and transformation relations of Bi and the four main Bi2O3 polymorphs [1, 7]. The light broken arrows represent bulk isotherm transformations, while the heavy solid arrows represent the micro-probe laser-induced transformations observed in this study. The processes of heating and cooling are indicated here by Δ and ∇, respectively. Temperature ranges without parentheses represent transition temperatures while the values in parentheses represent required minimum initial temperatures.
Fig. 4
Fig. 4 (a) Representative unpolarized Raman spectra from bismuth metal and resultant laser-induced Bi2O3 phases. Traces labeled ome-4-10-2133-i001.JPG, ome-4-10-2133-i002.JPG and ome-4-10-2133-i003.JPG represent the pure phase spectra, while the spectra labeled ome-4-10-2133-i004.JPG portray an initial (and sometimes final) combined phase. (b) Chosen ome-4-10-2133-i005.JPG transitional dual spectra for longer irradiation exposures. Arrows between traces indicate the observed oxidation phase sequence and all spectra have been normalized (normalization factors given) and offset for clarity. (c) Schematic of oxide aggregation leading to ome-4-10-2133-i005.JPG transition.

Tables (1)

Tables Icon

Table 1 Laser-induced bismuth oxide phase results for varying bismuth surfaces (corresponding to those shown in Figs. 1(a)–1(c)) irradiated by a large (4.2 μm) and small (1.7 μm) focused 632.8 nm laser beam. Representative Raman spectra of resulting phase(s) match with those given in Fig. 4(a).

Equations (1)

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

3 2 O 2 ( g ) + 2 Bi ( l ) β Bi 2 O 3 ( s ) α Bi 2 O 3 ( s ) ,
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.