Atomic layer deposition of crystalline Bi2O3 thin films and their conversion into Bi2S3 by thermal vapor sulfurization

H. F. Liu*a, K. K. Ansah Antwia, Y. D. Wangb, L. T. Onga, S. J. Chuaa and D. Z. Chia
aInstitute of Materials Research and Engineering (IMRE), A*STAR (Agency for Science, Technology and Research), 3 Research Link, Singapore 117602, Singapore. E-mail: liuhf@imre.a-star.edu.sg
bSchool of Engineering, Nanyang Polytechnic, 180 Ang Mo Kio Avenue 8, Singapore 569830, Singapore

Received 5th September 2014 , Accepted 22nd October 2014

First published on 22nd October 2014


Abstract

Crystalline Bi2O3 thin films have been grown on Si and quartz substrates at 300 °C in an atomic layer deposition (ALD) chamber using Bi(thd)3 (thd: 2,2,6,6-tetramethyl-3,5-heptanedionato) and H2O as precursors. Post-growth thermal sulfurization of such Bi2O3 films results in orthorhombic Bi2S3 at elevated temperatures. Atomic-force microscopy reveals that the surface roughness of Bi2O3 increases with growth cycles, and the roughness of Bi2O3/Si is generally larger than those of Bi2O3/quartz grown under the same conditions. However, after sulfurization, the surface morphologies became nearly similar in spite of their different substrates and/or thicknesses. The evolution of X-ray diffraction patterns as a function of growth cycles provide evidence that the growth starts with nonstoichiometric β-phase Bi2O2.3, and then transfers to α-Bi2O3 with the increase in growth cycles; the β-to-α growth transition occurs earlier on Si than that on a quartz substrate. Optical absorption spectroscopy reveals a bandgap of 3.0–3.5 eV for the as-grown Bi2O3 thin films, which is narrowed to 1.5 eV for the resultant Bi2S3 after sulfurization. X-ray photoelectron spectroscopy indicates that both the as-grown Bi2O3 and the sulfurized Bi2S3 thin films are n-type semiconductors with the valence band maxima located at about 2.1 and 1.0 eV below their Fermi levels, respectively. These observations, together with the large absorption coefficient of Bi2S3 in the wavelength range of visible light, suggest a significant potential for Bi2S3 thin films and/or Bi2S3/Bi2O3 heterojunctions in various optoelectronic devices, e.g., for solar energy conversions.


I. Introduction

Bi2S3, a direct bandgap semiconductor with the bandgap energy of 1.30 eV that locates between those of Si (1.12 eV) and GaAs (1.42 eV) and matches the potential requirement for water splitting (1.23 eV), has been attracting considerable research interest in recent years.1–3 Such a bandgap, together with the large optical absorption coefficient (>105 cm−1 in the wavelength range of visible light) of Bi2S3, makes it a promising absorber for photovoltaic and photoelectrochemical devices.4–6 In the nanoscale, Bi2S3 can form functional hybrid structures with metallic and/or semiconductor materials, e.g., carbon,7–9 gold,10,11 TiO2,12,13 ZnO,14 and In2S3,15 for various applications, including H- and Li-storage,7,8,16 nonlinear optical switching,11 photocatalysis,10,14,15 gas sensing and biomolecule detection.17,18 Although numerous optoelectronic devices and unique performances based on Bi2S3 and its hybrid structures have been demonstrated, the materials were generally synthesized by aqueous/solution-based methods.4–18 As a comparison, the methods of non-aqueous/solution growth of Bi2S3 thin films with respect to their photovoltaic applications that can be monolithically integrated into traditional thin film based solar cell technologies are less developed and generally limited to thermal vapor deposition and chemical vapor deposition (CVD) techniques.19–25

Atomic layer deposition (ALD) is a recently developed ultrathin film deposition technique controlled by the ‘surface saturation’ mechanism (i.e., self-limited surface reaction),26,27 which forms single atomic layers during each reaction cycle by alternatively feeding pulsed precursors. The by-products are purged with an inert gas in between each pulse in the reaction cycles, making the growth continuous. Recently, ultrathin α-Bi2O3 films have been synthesized on Si substrates by means of ALD using Bi(thd)3 (thd: 2,2,6,6-tetramethyl-3,5-heptanedionato) and H2O as the reaction precursors.28 Our previous studies show that Fe3S4, Mo/MoO3, and GaAs thin layers can be converted by thermal vapor sulfurization into crystalline FeS2 (pyrite), MoS2 and Ga2S3 films, respectively, with high purity in crystal phases.29 It has also been reported that Bi2S3 thin films can be iodinated into BiI3 at relatively low temperatures.30 These studies suggest that the ALD grown Bi2O3 might be able to convert into Bi2S3 under proper sulfurization conditions, while Bi2S3 can further convert back into Bi2O3 at elevated temperatures in the presence of oxygen.31 These conversion processes may have significant consequences in fabricating Bi2S3 thin films and Bi2S3/Bi2O3 hybrid structures for applications in various optoelectronic devices, e.g., photovoltaic solar cells,5 but are, unfortunately, seldom reported in the literature. In this regard, we have attempted to grow Bi2O3 thin films by employing ALD and to convert the ALD grown Bi2O3 thin films into Bi2S3 by thermal vapor sulfurization.

II. Experiment

The Bi2O3 thin films were synthesized on Si (001) and quartz substrates in an ALD chamber (Picosun Oy). Bi(thd)3 (thd: 2,2,6,6-tetramethyl-3,5-heptanedionato) and deionized water were used as bismuth and oxygen precursors, respectively. The former was evaporated from a booster bottle kept at 200 °C, while the latter was evaporated from a liquid-source bottle kept at room temperature. High-purity nitrogen (99.9995%) was used as the source carrier and purging gas. During growth, the chamber pressure was controlled at about 5–10 Torr, and the substrate temperature was maintained at 300 °C. In each ALD growth cycle, Bi(thd)3 and H2O vapor carried by N2 were alternately introduced into the ALD chamber, and the precursor pulses were separated by N2 purging. The flow rates of the carrier gas for Bi(thd)3 and H2O were set at 80 and 200 sccm, respectively. The pulse/purge durations for Bi(thd)3 and H2O were 1.6/6.0 s and 0.1/8.0 s, respectively. Before loading into the ALD chamber, the Si and quartz substrates were cleaned by acetone, methanol, and deionized water in sequence, which was followed by N2-flow drying. No additional chemical etching was adopted. The thin film samples in this study were grown for 500 to 1500 cycles.

Post-growth thermal vapor sulfurization of the ALD grown Bi2O3 thin films was carried out by thermal annealing in a sulfur vapor ambient. The sulfur vapor was evaporated from sulfur powder (99.995%) and carried by N2 in a tube-furnace. Details of the annealing setup and the general sulfurization procedures can be found in ref. 29. In this work, the sulfurization time was 20 min, and the sulfurization temperature was varied from 500 to 700 °C.

Atomic-force microscopy (AFM), X-ray diffraction (XRD, Cu-Kα1), optical absorption spectroscopy, Raman scattering spectroscopy, X-ray photoelectron spectroscopy (XPS), and photoluminescence (PL) have been used to study the structural and optical properties of the ALD grown Bi2O3 thin films before and after sulfurization. The AFM was operated in an air environment using a tapping mode. The film thickness was measured by AFM addressing on an intentionally scratched area, where a step is formed between the film and the surface of the substrate. The XRD was carried out in a general-area-detector diffraction system (GADDS, Bruker-D8), which has the significant advantage of being highly sensitive to crystal phase structures. Transmittance and reflectance were measured in an UV-vis-NIR scanning spectrophotometer using a bare quartz wafer and a standard aluminum mirror as the references, respectively. Raman scattering and PL measurements were conducted using a backscattering configuration with the 532 nm line of a Nd:YAG laser as the excitation source. XPS was carried out in a VG ESCALAB 220i-XL system equipped with a monochromatic Al-Kα1 X-ray source (1486.6 eV). The spectra were collectively shifted with the binding energy of C1s (due to contaminants, physically adsorbed on the surface, from the environment) at 285.0 eV as the reference.

III. Results and discussion

A. ALD growth of Bi2O3

Fig. 1(a) and (b) show the AFM images recorded from the Bi2O3 thin films grown for 1000 cycles (∼45 nm thick) on the Si and quartz substrates, respectively. Similarly, Fig. 1(c) and (d) show the AFM images recorded from the Bi2O3 thin films grown for 1500 cycles (∼110 nm thick) on the Si and quartz substrates, respectively. The insets in Fig. 1(b) and (d) are the optical absorbance spectra collected, using a bare quartz substrate as the reference, from the 1000- and 1500-cycle Bi2O3/quartz samples, respectively. It is clearly seen that the grain sizes in Fig. 1(c) and (d) are larger than those in Fig. 1(a) and (b), respectively. A detailed analysis revealed that the root-mean-square roughnesses are 4.9, 2.0, 9.2, and 5.8 nm as imaged by AFM in Figs. 1(a)–(d), respectively. These results indicate that the Bi2O3 thin films grown on quartz substrates are generally smoother than those grown on Si substrates under similar growth conditions; the grain size and the surface roughness of the Bi2O3 thin films increase with their growth cycles (i.e., the nominal thicknesses) on both the Si and quartz substrates.
image file: c4ra09896j-f1.tif
Fig. 1 AFM images recorded from ALD grown Bi2O3 thin films: (a) 1000 cycles on Si, (b) 1000 cycles on quartz, (c) 1500 cycles on Si, and (d) 1500 cycles on quartz. The insets in (b) and (d) are the absorbance spectra measured from the 1000- and 1500-cycle Bi2O3/quartz samples, respectively.

The absorbance spectra in Fig. 1(b) and (d) show that the onsets of apparent increment, i.e., the optical bandgap EOpt, locate at about 3.0 and 3.5 eV, respectively. These values are consistent with those reported for α-Bi2O3 at 3.04–3.89 eV, depending on the growth conditions.32 It is also seen that the absorbance ratio between the 1500- and 1000-cycle samples, for example, at the photon energy of 6.0 eV is about 3.3, which is considerably larger than the nominal thickness ratio (i.e., 1.5) of the two Bi2O3 thin film samples. This deviation is mainly caused by the variations in the effect of light reflections due to the significantly changed surface morphologies [see Fig. 1(b) and (d)].32,33 These morphology change induced variations in light reflection were not taken into account in the absorbance measurements (see a later section for further discussions).

Fig. 2(a) presents typical XRD patterns collected from the as-grown Bi2O3/Si samples. In terms of the angles, the diffraction peaks well match those of α-Bi2O3 (JCPDS 71-2274) and β-Bi2O2.3 (JCPDS 76-2477) with the former dominant in the thicker films. The intensity and the linewidth correlations between the diffraction peaks of α-Bi2O3 (012) and α-Bi2O3 (024), together with their spotty distributions in the two-dimensional imaging frames (see Fig. S1, ESI), indicate that the α-Bi2O3 grains have a collective growth orientation. These results, except for the incorporation of the nonstoichiometric β-phase, are consistent with those reported by Shen et al. in ref. 28. In fact, the coexistence of α- and β-phase crystals has already been observed in Bi2O3 nanostructures grown by oxidative metal vapor deposition,34 where the β-phase is, however, stoichiometric and largely dominant over the α-phase in the nanostructures. It is worth to note that although the α-Bi2O3 related diffractions increased with the growth cycles, the β-phase Bi2O2.3 related diffractions do not exhibit any significant variation. Such an evolution of XRD patterns indicates that the nonstoichiometric β-phase mainly locates at the initial grown region, i.e., in the area close to the Bi2O3/Si interface, probably due to the presence of the native amorphous SiO2 and the non-optimal ALD growth conditions.27 In comparison, the onset of β-to-α growth transition occurs later on the quartz substrate, which is only observed in the 1500-cycle Bi2O3/quartz sample (see Fig. S2, ESI). These results are consistent with the morphological comparisons shown in Fig. 1.


image file: c4ra09896j-f2.tif
Fig. 2 (a) XRD (Cu-Kα1) patterns of the ALD Bi2O3/Si thin film samples as a function of growth cycles and (b) Raman spectra collected at room temperature from a Bi2O3/quartz sample and a bare quartz substrate. Spectral fittings using Lorentzian functions are also presented for comparison.

Fig. 2(b) presents the typical Raman spectrum taken from a bare quartz substrate and a 1500-cycle Bi2O3/quartz sample along with its spectral fittings. The use of a quartz substrate is to avoid the complex photon features of the Si substrate in the studied frequency range. By fitting to the experimental spectrum using Lorentzian functions and comparing to the spectrum of the quartz substrate, we have unambiguously deconvoluted five additional features as indicated by P1–P4 and Pa in Fig. 2(b) that originated from the Bi2O3 thin film. The Pa feature (at about 205 cm−1), when compared to P1–P4, is much larger in linewidth (∼130 cm−1), which is probably a PL-emission feature due to defects and/or Bi-ions/clusters rather than a phonon mode.35,36 The features of P1–P4 are well consistent with those of α-Bi2O3 at 101, 151, 314, and 598 cm−1, respectively.36 In contrast, the typical modes of β-Bi2O3 at 124, 311, and 462 cm−1 are not distinguishable in Fig. 2(b),37 probably due to the nonstoichiometry-induced disordering and/or the low crystal quality of β-Bi2O2.3 that was grown at the initial stage directly on the surface of the quartz substrate. A further comparison with those experimental observations and theoretical calculations reported by Narang et al. in ref. 36 reveals that only P1, P2, and P3 are the Raman active phonon modes out of the 14/22 observed/predicted ones in the frequency range of 100–600 cm−1. This is possibly due to the structural disorders and/or impurity defects of the α-Bi2O3 crystals; moreover, the collective growth orientation of the thin film as revealed by XRD discussed above could also result in the limitation of phonon detections in the backscattering configuration.38 The feature of P4 at ∼598 cm−1 is an infrared active and Raman silent mode; its presence in the Raman spectrum can be attributed to the disorder- and/or defect-induced phonon activation, i.e., the infrared-active to Raman-active transition.3,39

B. Sulfurizing Bi2O3 into Bi2S3

Figures 3(a)–(c) present the AFM images recorded after sulfurization at 500 °C from the 500-, 1000-, and 1500-cycle Bi2O3/Si samples, respectively. For comparison, an AFM image recorded from the 1500-cycle Bi2O3/Si sample after sulfurization at 600 °C is presented in Fig. 3(d). Unlike the as-grown samples (see Fig. 1), after sulfurization, the surface morphologies of the samples on the quartz substrates (see Fig. S3, ESI) are nearly similar to those of the samples on the Si substrates grown and sulfurized under identical conditions. For the samples sulfurized at 500 °C, it can be observed in Figs. 3(a)–(c) that their morphologies are also quite similar to one another. However, in Fig. 3(d), it is seen that the grain sizes are apparently increased by increasing the sulfurization temperature to 600 °C; the morphology clearly exhibits an onset of surface evaporation. In fact, we have observed that the original Bi2O3 films were completely evaporated when the sulfurization temperature was further increased to 700 °C, even under a sulfur-rich condition. This is because of the sublimation behavior and the high vapor pressure of Bi2S3 at temperatures higher than 550 °C.40
image file: c4ra09896j-f3.tif
Fig. 3 AFM images recorded from the sulfurized Bi2S3 samples, i.e., the Bi2O3 thin film samples after sulfurization on Si substrates: (a) 500 cycles sulfurized at 500 °C, (b) 1000 cycles sulfurized at 500 °C, (c) 1500 cycles sulfurized at 500 °C, and (d) 1500 cycles sulfurized at 600 °C.

Fig. 4(a) shows the XRD patterns collected from the sulfurized samples on Si together with that of JCPDS 75-1036 (orthorhombic Bi2S3). The excellent match between the experimental spectra and JCPDS 75-1036, in terms of peak positions and intensity ratios, provides clear-cut evidence that orthorhombic Bi2S3 were obtained with high phase purities from the ALD grown Bi2O3 thin films. It is also seen in Fig. 4(a) that the diffraction intensities increase with the nominal film thickness after sulfurization at 500 °C. However, an increase in the sulfurization temperature to 600 °C turns to reduce the diffraction intensities, which is consistent with the AFM observations (see Fig. 3) and confirms the onset of surface evaporation of Bi2S3 at 600 °C. It has to be noted that, after sulfurization, identical XRD patterns of the thin films (i.e., Bi2S3) on the Si and quartz substrates are observed (see Fig. S4, ESI). This is consistent with the morphological similarities observed in Fig. 3 (on Si substrates) and Fig. S2 (on quartz substrates) after sulfurization (see discussions above).


image file: c4ra09896j-f4.tif
Fig. 4 (a) XRD (Cu-Kα1) patterns collected from the Bi2S3 thin film samples obtained by thermal vapor sulfurizing the ALD grown Bi2O3 on Si substrates. The patterns of orthorhombic Bi2S3 (JCPDS-ICDD 75-1306) are also presented for comparison. (b) Raman spectra collected at room temperature from the sulfurized Bi2S3 thin films grown on quartz substrates using the 532 nm line of a Nd:YAG laser as the excitation source. The spectra of an as-grown Bi2O3 sample and a bare quartz substrate are also presented for easier comparison.

Furthermore, to avoid the complicated photon features of Si in the studied frequency range, Raman spectroscopy studies were carried out for the sulfurized thin film samples on the quartz substrates, and the results are shown in Fig. 4(b). For simple comparisons, the Raman spectra collected from an as-grown sample and a bare substrate are also presented. It is seen that, after sulfurization, the typical modes of α-Bi2O3 at 314 and 598 cm−1, corresponding to the displacements of oxygen atoms with respect to the bismuth atoms,36 disappeared; instead, five phonon modes (indicated by P1–P5) emerged in the Raman spectra of the samples sulfurized at 500 °C, and an additional mode (indicated by P6) appeared in the spectrum of the sample sulfurized at 600 °C. The modes P1–P5 are consistent with the Raman active phonons of Bi2S3 observed by Zhao et al. at 100.0, 168.7, 186.0, 237.1, and 254.5 cm−1, respectively.3 The mode of P6 at 117.0 cm−1 does not correspond to any of the theoretically predicted Raman active modes but well matches the infrared active mode of Bi2S3 at 119.3 cm−1.2,3 We have already mentioned above that structural disorders and/or impurity defects could induce such infrared-active to Raman-active transitions. Since the transition only occurs at a higher sulfurization temperature, it is reasonable that impurity atoms, e.g., Si, were doped into the Bi2S3 crystals from the substrate due to the increased sulfurization temperature. Nevertheless, both the XRD and Raman spectroscopy studies show that crystalline orthorhombic Bi2S3 thin films with high phase purity have been obtained by sulfurizing the ALD grown Bi2O3.

C. Optical properties of Bi2S3

Fig. 5(a) shows the absorbance spectra collected from the sulfurized Bi2S3 thin films on the quartz substrates. The spectra exhibit a knee-like feature with an exciton resonance absorption peak at 2.96 eV.41 The values of absorbance, e.g., at the photo energy of 2.96 eV, have a strong linear correlation with the nominal thickness of the thin films. This observation indicates that the surface and the internal light reflections have similar effects on the absorbance measurements due to the similar morphologies after sulfurization, different from those of the as-grown thin films discussed above. In general, the absorption coefficient of a thin film can be determined via measuring the transmittance T, reflectance R, and the film thickness d. In the case of negligible light reflections, the absorption coefficient reads α = ln[1/T]/d;32 however, when only the surface reflection is taken into account, the absorption coefficient reads α = ln[(1 − R)/T]/d; furthermore, when both the surface and the internal reflections are taken into account, the absorption coefficient reads image file: c4ra09896j-t1.tif,33,42 which is obviously more accurate. To verify the effect of reflections and to measure the absorption coefficient of the resultant Bi2S3 with an increased accuracy, we have measured and shown, in Fig. 5(b), the T and R for the Bi2S3 thin film obtained by sulfurizing the 1500-cycle Bi2O3/quartz sample at 500 °C. The film thickness, d = 70 nm, was measured by AFM from an intentionally scratched area (see Fig. S5, ESI). In consequence, the absorption coefficient spectra were obtained, as shown in Fig. 5(c), with and without counting the internal light reflections. It is seen that the EOpt is reduced from 3.0 to 3.5 eV of the as-grown Bi2O3 to 1.50 eV after sulfurization. The overestimation of the absorption coefficients in the range of photon energies larger than EOpt due to the uncalculated internal reflections is less than 10%. The most important observation in Fig. 5(c) is that the optical absorption coefficients in the wavelength range of 400–700 nm are larger than 1.5 × 105 cm−1, which confirms that the resultant Bi2S3 could have significant potentials for solar energy conversion applications.
image file: c4ra09896j-f5.tif
Fig. 5 (a) Optical absorbance spectra measured from the sulfurized Bi2S3/quartz samples, (b) reflectance, R, and transmittance, T, of a sulfurized Bi2S3/quartz sample with the initial Bi2O3 grown for 1500 cycles, and (c) optical absorption coefficients derived from the measured R and T with and without counting the internal light reflections.

Fig. 6 shows the PL spectra collected at low-temperatures from a sulfurized Bi2S3 thin film sample. Basically, the decrease in PL emission intensity with the increase in sample temperature is due to the activation of nonradiative recombination centers. The PL peaks are asymmetric and can be deconvoluted into two distinct emissions with the peak splitting of ∼23 meV. In general, these PL emission energies are a few tens of meV smaller than the bandgap energy of EOpt = 1.50 eV obtained at room temperature [see Fig. 5(c)], which clearly indicate that the PL emissions are primarily transitions associated with shallow defects rather than the band-to-band emissions reported by Sträter et al. in ref. 43. However, the intact PL emission energies as the sample temperature increased from 10 to 60 K and the redshift of ∼10 meV induced by a further temperature increase to 110 K are consistent with the reported temperature-dependent bandgap shift of Bi2S3 (see Fig. S6, ESI).43 It is evident that more work is needed to identify the detailed shallow defects and the recombination mechanisms involved in the PL emissions. Nevertheless, the PL spectroscopy, together with the optical absorption measurements, provides clear evidence that the bandgap energy of the crystalline orthorhombic Bi2S3 thin films is about 1.5 eV, which is slightly larger than the predicted value of 1.3 eV.


image file: c4ra09896j-f6.tif
Fig. 6 Low-temperature PL spectra collected from a sulfurized Bi2S3 thin film sample with the initial Bi2O3 grown for 1500 cycles on Si substrate.

D. Electronic structures of Bi2O3 and Bi2S3

Electronic states of the as-grown Bi2O3 thin films and the sulfurized Bi2S3 ones were further compared using XPS. Fig. 7(a) shows the XPS survey spectra which revealed only O and Bi along with C (due to contaminants from the environment) for the as-grown Bi2O3 sample, while S presented in the sulfurized Bi2S3 samples together with largely reduced O. The high-resolution XPS spectra of the Bi4f and O1s/S2s core levels are presented in Fig. 7(b) and (c), respectively. It is clearly seen in Fig. 7(b) that the core levels of Bi4f were shifted to lower binding energies by ∼1.1 eV after sulfurization due to the smaller electronegativity of S than that of O. However, it can be observed that, after sulfurization, small shoulders emerged at the higher energy sides of both the Bi4f5/2 and Bi4f7/2 peaks. The peak energies of the small Bi4f shoulders of the Bi2S3 samples are nearly the same as those of Bi4f of the as-grown Bi2O3 sample. Similarly, O1s is also detected in the sulfurized Bi2S3 samples and can be deconvoluted into three distinct peaks with one located at 530 eV, i.e., the same binding energy of the lattice O2− of the as-grown Bi2O3 [see Fig. 7(c)]. Similar to Bi2O3 grown by a solution method,5 the presence of O–H (∼531.2 eV) in the as-grown Bi2O3, which is significantly reduced in intensity after sulfurization, could be due to the use of a water precursor and the relatively low substrate temperature during the ALD growth in this study. The other O1s peak (∼533 eV) of the sulfurized Bi2S3 samples is possibly due to adsorbates on the surface, e.g., chemically adsorbed oxygen, Oad, at 532.5 eV and H2O at 533 eV.5 On the other hand, the symmetric line shape of the S2s peaks in Fig. 7(c), as well as their binding energy at 225.5 eV, indicates the absence of elemental sulfur in the sulfurized Bi2S3 samples. The combination of Bi4f and O1s spectra of the sulfurized samples in Fig. 7(b) and (c) provides evidence for the incorporation of Bi2O3 in the Bi2S3 samples. It is well known that XPS is a surface sensitive technique, which can only detect the information from a depth of a few nanometers below the film surface. It is thus unclear, whether Bi2O3 was formed after sulfurization due to a native oxidation of Bi2S3 or it remained in Bi2S3 due to a limited sulfurization time. However, as we have discussed above that both XRD and Raman scattering measurements could not detect the presence of Bi2O3 in the sulfurized Bi2S3 thin films, it is more reasonable to suggest that Bi2O3 was formed due to the native surface oxidation of Bi2S3.
image file: c4ra09896j-f7.tif
Fig. 7 XPS spectra collected from the ALD grown Bi2O3/Si thin films samples before and after sulfurization: (a) survey spectra, (b) high-resolution spectra of Bi4f and S2p core levels, (c) high-resolution spectra of O1s and S2s core levels, and (d) high-resolution spectra of the valence bands.

Fig. 7(d) shows the valence band spectra collected from the as-grown Bi2O3 and the sulfurized Bi2S3 thin films. The sulfurization induced increase in the sharpness of the valence band maximum, together with the reduced linewidths of Bi4f in Fig. 7(b), indicates the increased crystallinity and the reduced density of defect states in the bandgap. Linear fittings to the valence band maxima reveal that they have been raised from 2.1 to 1.0 eV below the Fermi level after sulfurization. Taking into account the optical bandgaps of 3.0–3.5 eV [see Fig. 1(b) and (d) and Fig. S7, ESI] and 1.5 eV for the as-grown Bi2O3 and the sulfurized Bi2S3 thin films, respectively, we may conclude that they are both n-type semiconductors. These observations clearly indicate that type-I Bi2S3/Bi2O3 heterojunctions can be formed by partially sulfurizing Bi2O3 for effective photocatalyst and/or photoelectrochemical applications.

IV. Conclusions

In conclusion, Bi2O3 crystalline thin films were grown on Si and quartz substrates by atomic layer deposition at 300 °C. The films in smaller thickness are dominated by nonstoichiometric β-Bi2O2.3, but the component of stoichiometric α-Bi2O3 increases with the grown thickness and becomes dominant when the growth period is larger than 1000 cycles on Si. Such a β-to-α growth transition also occurs on a quartz substrate but later than that on the Si substrate. Morphological evolutions of the thin films show that clustering of Bi2O3 crystallites occurred when the growth period is increased over 1000 cycles, leading to the larger surface roughnesses. As a result, the effect of surface and internal light reflections on the absorbance measurements significantly varied by the increase in film thickness. Optical bandgaps of 3.0–3.5 eV, depending on the film thickness, were obtained for the as-grown Bi2O3 thin films. After sulfurization at elevated temperatures, the Bi2O3 thin films were converted into orthorhombic Bi2S3, and the surface morphologies became similar in spite of different substrates and/or thicknesses. The optical bandgap was reduced to 1.5 eV after sulfurization, and this value is fairly constant among the samples with different film thicknesses. Absorption coefficients of larger than 105 cm−1 in the wavelength range of visible light are observed for the sulfurized Bi2S3 thin films, confirming their significant potentials in solar energy conversion applications. Electronic structural studies, together with the absorption spectroscopies, provide evidence that both the as-grown Bi2O3 and the sulfurized Bi2S3 are n-type semiconductors with the valence band maxima located at 2.1 and 1.0 eV below their Fermi levels, respectively. These results suggest that type-I Bi2S3/Bi2O3 heterojunctions can be formed via partially sulfurizing Bi2O3 for effective photocatalyst and/or photoelectrochemical applications.

Acknowledgements

The authors would like to thank Lim Poh Chong for his help in XRD data collection and Dr Dong Zhaogang for his help in the Raman scattering experiments.

References

  1. H. Koc, H. Ozisik, E. Deligöz, A. M. Mamedov and E. Ozbay, J. Mol. Model., 2014, 20, 2180 CrossRef PubMed.
  2. I. Efthimiopoulos, J. Kemichick, X. Zhou, S. V. Khare, D. Ikuta and Y. Wang, J. Phys. Chem. A, 2014, 118, 1713 CrossRef CAS PubMed.
  3. Y. Zhao, K. T. E. Chua, C. K. Gan, J. Zhang, B. Peng, Z. Peng and Q. Xiong, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 205330 CrossRef.
  4. D. Becerra, M. T. S. Nair and P. K. Nair, J. Electrochem. Soc., 2011, 158, H741 CrossRef CAS PubMed.
  5. X. Lu, F. Pu, Y. Xia, W. Huang and Z. Li, Appl. Surf. Sci., 2014, 299, 131 CrossRef CAS PubMed.
  6. Y. Lu, J. Jia and G. Yi, CrystEngComm, 2012, 14, 3433 RSC.
  7. Y. Zhao, D. Gao, J. Ni, L. Gao, J. Yang and Y. Li, Nano Res., 2014, 7, 765 CrossRef CAS PubMed.
  8. Y. Zhao, T. Liu, H. Xia, L. Zhang, J. Jiang, M. Shen, J. Ni and L. Gao, J. Mater. Chem. A, 2014, 2, 13854 CAS.
  9. G. Li, X. Chen and G. Gao, Nanoscale, 2014, 6, 3283 RSC.
  10. G. Manna, R. Bose and N. Pradhan, Angew. Chem., 2014, 126, 6861 CrossRef.
  11. J. L. T. Chen, V. Nalla, G. Kannaiyan, V. Mamidala, W. Ji and J. J. Vittal, New J. Chem., 2014, 38, 985 RSC.
  12. I. Zumeta-Dubé, V. Ruiz-Ruiz, D. Díaz, S. Rodil-Posadas and A. Zeinert, J. Phys. Chem. C, 2014, 118, 11495 Search PubMed.
  13. H. Yu, J. Huang, H. Zhang, Q. Zhao and X. Zhong, Nanotechnology, 2014, 25, 215702 CrossRef PubMed.
  14. W. Xitao, L. Rong and W. Kang, J. Mater. Chem. A, 2014, 2, 8304 Search PubMed.
  15. J. Zhou, G. Tian, Y. Chen, Y. Shi, C. Tian, K. Pan and H. Fu, Sci. Rep., 2014, 4, 4027 Search PubMed.
  16. B. Zhang, X. Ye, W. Hou, Y. Zhao and Y. Xie, J. Phys. Chem. B, 2006, 110, 8978 CrossRef CAS PubMed.
  17. K. Yao, Z. Zhang, X. Liang, Q. Chen, L. Peng and Y. Yu, J. Phys. Chem. B, 2006, 110, 21408 CrossRef CAS PubMed.
  18. L. Cademartiri, F. Scotognella, P. G. O'Brien, B. V. Lotsch, J. Thomson, S. Petrov, N. P. Kherani and G. A. Ozin, Nano Lett., 2009, 9, 1482 CrossRef CAS PubMed.
  19. S. T. Haaf, H. Sträter, R. Brüggemann, G. H. Bauer, C. Felser and G. Jakob, Thin Solid Films, 2013, 535, 394 CrossRef PubMed; S. T. Haaf, B. Balke, C. Felser and G. Jakob, J. Appl. Phys., 2012, 112, 053705 CrossRef PubMed.
  20. X. Yu and C. Cao, Cryst. Growth Des., 2008, 8, 3951 CAS.
  21. M. E. Rincón, M. Sánchez, P. J. George, A. Sánchez and P. K. Nair, J. Solid State Chem., 1998, 136, 167 CrossRef.
  22. J. Lukose and B. Pradeep, Solid State Commun., 1991, 78, 535 CrossRef CAS.
  23. H. Zhang, J. Ge and Y. Li, Chem. Vap. Deposition, 2005, 11, 147 CrossRef CAS.
  24. J. Waters, D. Crouch, J. Raftery and P. O'Brien, Chem. Mater., 2004, 16, 3289 CrossRef CAS.
  25. A. A. Tahir, M. A. Ehsan, M. Mazhar, K. G. U. Wijayantha, M. Zeller and A. D. Hunter, Chem. Mater., 2010, 22, 5084 CrossRef CAS.
  26. C. Marichy, M. Bechelany and N. Pinna, Adv. Mater., 2012, 24, 1017 CrossRef CAS PubMed.
  27. S. M. George, Chem. Rev., 2010, 110, 111 CrossRef CAS PubMed.
  28. Y. D. Shen, Y. W. Li, W. M. Li, J. Z. Zhang, Z. G. Hu and J. H. Chu, J. Phys. Chem. C, 2012, 116, 3449 CAS.
  29. H. F. Liu and D. Z. Chi, J. Vac. Sci. Technol., A, 2012, 30, 04D102 Search PubMed; H. F. Liu, K. K. Ansah Antwi, S. J. Chua and D. Z. Chi, Nanoscale, 2014, 6, 624 RSC; H. F. Liu, K. K. Ansah Antwi, J. F. Ying, S. J. Chua and D. Z. Chi, Nanotechnology, 2014, 25, 405702 CrossRef CAS PubMed; H. F. Liu, K. K. Ansah Antwi, N. L. Yakovlev, H. R. Tan, L. T. Ong, S. J. Chua and D. Z. Chi, ACS Appl. Mater. Interfaces, 2014, 6, 3501 Search PubMed.
  30. B. B. Nayak, S. K. Singh and B. S. Acharya, Phys. Status Solidi A, 1988, 105, 455 CrossRef CAS.
  31. H. F. Liu, A. Huang and D. Z. Chi, J. Phys. D: Appl. Phys., 2010, 43, 455405 CrossRef CAS; H. F. Liu, D. Z. Chi and W. Liu, CrystEngComm, 2012, 14, 7140 RSC.
  32. L. Leontie, M. Caraman, M. Delibaş and G. I. Rusu, Mater. Res. Bull., 2001, 36, 1629 CrossRef CAS.
  33. J. I. Pankove, Optical Processes in Semiconductors, Dover, New York, 1971 Search PubMed.
  34. L. Kumari, J. Lin and Y. Ma, J. Phys.: Condens. Matter, 2007, 19, 406204 CrossRef PubMed.
  35. B. Ling, X. W. Sun, J. L. Zhao, Y. Q. Shen, Z. L. Dong, L. D. Sun, S. F. Li and S. Zhang, J. Nanosci. Nanotechnol., 2010, 10, 8322 CrossRef CAS PubMed.
  36. S. N. Narang, N. D. Patel and V. B. Kartha, J. Mol. Struct., 1994, 327, 221 CrossRef CAS.
  37. F. D. Hardcastle and I. E. Wachs, J. Solid State Chem., 1992, 97, 319 CrossRef CAS.
  38. V. N. Denisov, A. N. Ivlev, A. S. Lipin, B. N. Mavrin and V. G. Orlov, J. Phys.: Condens. Matter, 1997, 9, 4967 CrossRef CAS.
  39. H. F. Liu and S. J. Chua, Appl. Phys. Lett., 2010, 96, 091902 CrossRef PubMed.
  40. V. Piacente, V. D. Gianfreda and G. Bardi, J. Chem. Thermodyn., 1983, 15, 7 CrossRef CAS.
  41. H. F. Liu and S. J. Chua, J. Appl. Phys., 2009, 106, 023511 CrossRef PubMed.
  42. J. I. Pankove, Phys. Rev., 1965, 140, A2059 CrossRef.
  43. H. Sträter, S. T. Haaf, R. Brüggemann, G. Jakob, N. Nilius and G. H. Bauer, Phys. Status Solidi B, 2014, 251, 2247–2256 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra09896j

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