Next Article in Journal
Determination of the Projected Atomic Potential by Deconvolution of the Auto-Correlation Function of TEM Electron Nano-Diffraction Patterns
Next Article in Special Issue
A Review on the Fabrication of Hierarchical ZnO Nanostructures for Photocatalysis Application
Previous Article in Journal
Structural and Theoretical Evidence of the Depleted Proton Affinity of the N3-Atom in Acyclovir
Previous Article in Special Issue
Effect of HNO3 Concentration on the Morphologies and Properties of Bi2WO6 Photocatalyst Synthesized by a Hydrothermal Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis, Characterization, and Visible-light Photocatalytic Activities of 3D Hierarchical Bi2S3 Architectures Assembled by Nanoplatelets

1
School of Materials Science and Engineering, Luoyang Institute of Science and Technology, Luoyang 471023, China
2
School of Materials Science and Engineering, Guilin University of Technology, Guilin 541004, China
3
School of Materials Science and Engineering, Henan Polytechnic University, Jiaozuo 45400, China
*
Author to whom correspondence should be addressed.
Crystals 2016, 6(11), 140; https://doi.org/10.3390/cryst6110140
Submission received: 8 August 2016 / Revised: 18 October 2016 / Accepted: 24 October 2016 / Published: 31 October 2016
(This article belongs to the Special Issue Mesocrystals and Hierarchical Structures)

Abstract

:
3D hierarchical Bi2S3 architectures have been successfully synthesized via a simple and effective hydrothermal process. The as-prepared Bi2S3 samples were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), nitrogen adsorption- desorption isotherms, and UV-vis diffuse reflectance spectrum (DRS). The observation of field emission scanning electron microscope (FESEM) images showed that numerous nanoplatelets are randomly arranged and interconnected with each other, which are assembled into 3D hierarchical Bi2S3 architectures. The photocatalytic activity of the as-prepared Bi2S3 samples was evaluated by the degradation of Rhodamine B (RhB) under visible light irradiation. The effect of hydrothermal temperature, reaction time, pH value and shape on the photocatalytic efficiency of the as-prepared Bi2S3 samples was investigated. The results showed that 3D hierarchical Bi2S3 architectures prepared at 165 °C for 12 h at a pH of 2.4 exhibits high photocatalytic efficiency, which could be ascribed to the synergetic effect of the shape, surface area, crystallinity, band gap and crystalline size.

1. Introduction

Nowadays, we are confronted with the problem related to various organic pollutants releasing from various textile and chemical plants accompanied with the continuously increasing industrialization. Most organic pollutants are difficult to effectively dispose of by natural means, leading to water sources and the environment to be contaminated and polluted. Subsequently, the contaminated water source and the polluted air have become an enormous threat to public health. Photocatalysis, considered an efficient and reliable technology for the environmental remediation of poisonous organic pollutants, has the merits of simple operation, ambient condition, low cost, high efficiency and reusability [1,2]. Over the past decade, TiO2 has been widely investigated as one of the most representative photocatalysts for degrading pollutants. However, traditional TiO2 exhibits ineffective photocatalytic performance under visible light irradiation due to its wide band gap and low quantum efficiency. Therefore, the development of novel visible-light-driven photocatalysts is required for the purpose of taking full use of solar energy and improving quantum efficiency. In order to achieve the goal, much research has been devoted to the modification of traditional TiO2 [3,4,5,6,7] and the pursuit of other materials such as bismuth sulfide (Bi2S3) [8,9,10].
With a direct narrow band gap of around 1.3 eV, bismuth sulfide (Bi2S3), emerging as one promising member of group V–VI binary semi-conductors, has been attracting considerable attention due to its wide range of application in the field of solar cells [8,9], photo-detectors [10,11,12,13,14], gas sensors [14], thermoelectric devices [15], X-ray computed tomography imaging [16], lithium-ion batteries [17,18,19,20], and photocatalysis [21,22,23]. Bi2S3 micro/nanostructures with various morphologies, including one-dimensional (1D) nanorods/nanowires, two-dimensional (2D) nanosheets/flakes, and three-dimensional (3D) hierarchical complex architectures, have been successfully fabricated in previous studies [10,11,12,13,14,15,17,18,19,23,24,25,26], since the properties of functional nanostructured materials are strongly dependent on the tailored morphology, microstructure and surface properties. Among them, 3D hierarchical complex architectures have exhibited excellent properties for applications in gas sensors, lithium batteries, and photocatalysts due to their large surface area and facile electron (ion) transport. Especially, 3D hierarchical complex architectures assembled by 2D nanostructures have porous structure, present more active sites at large surface area for adsorbing reactant molecules and photocatalytic reactions, thus are desirable for achieving superior photocatalytic activity [27,28]. To date, most of the reported 3D hierarchical Bi2S3 architectures are assembled by 1D nanostructure, while there are few reports on the synthesis and photocatalytic performance of 3D hierarchical Bi2S3 architectures assembled by 2D nanoplatelets. Therefore, it is essential to design a facile, highly efficient and environmentally friendly route to prepare 3D hierarchical Bi2S3 architectures assembled by 2D nanoplatelets for improving visible-light photocatalyitic activity.
Hydrothermal synthesis technique, an effectively and simply controlled solution-based method, involves the formation of well-crystallized products at relatively low temperatures, which has been widely adopted for the synthesis of nanostructured photocatalysts with tailored morphologies [29,30,31]. In this study, we reported a facile hydrothermal synthesis route for preparing 3D hierarchical Bi2S3 architectures assembled by 2D nanoplatelets. A series of techniques, including XRD, SEM, TEM, nitrogen adsorption–desorption isotherms, and UV-vis diffuse reflectance spectrum (DRS), were used to characterize the as-prepared products. The photocatalytic activities of the as-prepared samples were tested in the degradation of Rhodamine B (RhB) under visible light irradiation.

2. Results and Discussions

2.1. Phase Study

The crystallinity and phase structure of the as-prepared 3D hierarchical Bi2S3 architectures were investigated by XRD and Raman spectrum, as displayed in Figure 1. XRD result (Figure 1a) showed that the intensities of the diffraction peaks of the as-prepared Bi2S3 samples increased with increasing hydrothermal temperatures, indicating that the degree of crystallinity of the as-prepared Bi2S3 samples tended to be strengthened with increasing hydrothermal temperatures. Some other impurity peaks of the as-prepared Bi2S3 sample are detected when the hydrothermal temperature is lower than 165 °C. As for the representative Bi2S3 sample (BS-2#), all the diffraction peaks can be indexed to be the orthorhombic phase Bi2S3 (JCPDS No. 17-0320). No other characteristic peaks are observed, implying that the representative Bi2S3 sample (BS-2#) is of high purity. The sharp diffraction peaks of the XRD pattern for the BS-2# sample demonstrate that well-crystallized orthorhombic Bi2S3 was successfully achieved under the present synthetic conditions. Interestingly, the intensity ratio of (221) peak to (211) peak seemingly deviates from the standard value, implying that the Bi2S3 crystal tends to have special anisotropic growth. Figure 1b shows the Raman spectrum of the BS-2# sample, in which exhibits one sharp peak centered about 130 cm−1, and two mild peaks centered about 400 cm−1 and 600 cm−1, respectively. Different morphologies and structures can shift the wavenumbers of Raman spectrum due to the difference in the size and surface photo modes [32,33]; thus, the obtained result of Raman spectrum can also confirm the presence of the orthorhombic phase Bi2S3.

2.2. Morphology and Proposed Formation Process

Figure 2a shows the low-magnification FESEM image of the representative Bi2S3 sample (BS-2#). The BS-2# sample is composed of discrete 3D hierarchical architectures with inhomogeneous size. The higher magnification FESEM image (Figure 2b) reveals that 3D hierarchical architectures has porous structure, in which numerous nanoplatelets are randomly arranged and interconnected. Such a unique structure feature implies that the as-prepared 3D hierarchical Bi2S3 architectures are promising to exhibit effective photocatalytic activity for degrading the environmental pollutants. The S- and Bi-mappings of the BS-2# sample are shown in Figure 2d,e, revealing that the BS-2# sample only consists of Bi and S elements. The structure of the BS-2# sample was further investigated by TEM. The result of TEM observation (Figure 3a,b) shows the widths of nanoplatelets are inhomogeneous, while the thickness of nanoplatelets is thin (5–10 nm), which is in good agreement with that of FESEM observation. Figure 3c is a typical HRTEM image taken from the single nanoplatelet, in which an interplanar lattice of 0.374 nm between adjacent lattice fringes can be indexed to (101) plane of orthorhombic phase Bi2S3.
In order to investigate the effect of the surfactant on the morphology, the hydrothermal synthetic process was carried out without the addition of CTAB. Only Bi2S3 nanrods with an average diameter of 20 nm were obtained under such a synthetic condition, as is seen in Figure 3d. Thus, we can infer that the surfactant of CTAB plays a key role in the formation of 3D hierarchical Bi2S3 architectures assembled by nanoplatelets. Herein, a probable formation process of 3D hierarchical Bi2S3 architectures could be proposed as follows. At initial stage of hydrothermal process, Tu coordinated with Bi3+ and formed the Bi3+(III) complexes. Under high pressure and hydrothermal temperature, the Bi3+(III) complexes tended to aggregate and crystal nuclei of Bi2S3 came into being in situ. Then, Bi2S3 nanoplatelets were formed with the assistance of CTAB. Subsequently, Bi2S3 nanoplatelets grew larger and thicker with the ripening process going on. In order to reduce the surface energy, the as-grown Bi2S3 nanoplatelets with adsorbed CTAB molecules were likely to convolute and self-assemble loosely, resulting in the formation of 3D hierarchical Bi2S3 architectures. The probable formation process is illustrated in Figure 4.

2.3. Bet Surface Area and Pore Volume

The specific surface area and pore structure of the BS-2# sample were investigated by nitrogen adsorption–desorption isotherm. Figure 5 depicts the nitrogen adsorption–desorption isotherm and its corresponding pore size distribution of the BS-2# sample. From Figure 5a, we find that the BS-2# sample displays a type IV isotherm with a distinct hysteresis loop in the range of 0.8–1.0 (P/P0), implying the presence of mesopores, which is ascribed to the aggregation of Bi2S3 nanoplatelets [34,35,36]. The specific surface area of the BS-2# sample was calculated to be 49.5 m2·g−1 by Brunauer–Emmett–Teller equation. Due to its large specific area, the BS-2# sample can absorb more photons and reactants, which results in increasing photocatalytic active sites correspondingly. Moreover, it is clear from Figure 5b that the pore distribution of t the BS-2# sample exhibits two kinds of mesopores pores with an average diameter of 8.0 nm and 38.3 nm. Based on above results, the obtained mesoporous structure of photocatalyst provides efficient transport pathways to the interior, which may be extremely useful for improving its photocatalytic activity.

2.4. Uv-Vis Spectra

As is well known, the photocatalytic activities of photocatalysts are strongly dependent on their energy band structure. Thus, UV-vis diffuse reflectance spectrum (Figure 6) was recorded to obtain optical properties of the BS-2# sample. It is clearly seen from Figure 6a that the BS-2# sample displays good optical absorption ability in the entire visible light region. Consulting the previous studies [37], the band gap (Eg) of Bi2S3 can be explored by the extrapolation method based on Kubelka–Munk equation. The above equation can be expressed as A(hν-Eg)2 = (αhν), where α is the adsorption coefficient, A is the constant, hν is the photon energy, n = 2 for an indirectly allowed transition semiconductor and n = 1/2 for directly allowed transition semiconductor. Figure 6b plots the (αhν)2 against photo energy () curve. The band gap energy of the BS-2# sample is estimated to be 1.48 eV, which is approximate to the reported band gap energy [32,38]. Based on the results, the as-obtained 3D hierarchical Bi2S3 architectures belong to one kind of relatively narrow band-gap semiconductors, indicating that more optical energy could be utilized to produce photo-generated carriers for taking part in the photocatalytic degradation reaction.

2.5. Photocatalytic Activities

The photocatalytic activity of the as-prepared Bi2S3 samples was evaluated by degrading RhB aqueous solution under visible light irradiation. The changes in the absorption spectra of RhB at 554 nm were used to monitor the photocatalytic performance. It is well established that synthetic conditions of photocatalysts are likely to affect the crystallinity and shape of the as-prepared samples. Thus, we compared the photodegradation of 3D hierarchical Bi2S3 architectures obtained under different hydrothermal temperature, time and pH value in this work. As is shown in Figure 7a, no obvious variation of RhB concentration was observed in the absence of either photocatalyst or light irradiation. Only 6% RhB can be degraded by pure TiO2 (P25) after irradiating for 60 min. The degradation rate of the BS-2# sample reaches 98.6% after irradiating for 80 min under visible light, while 72.3% and 74.9% degradation rates of RhB are observed for BS-1# and BS-3# sample, respectively. It can also be seen from Figure 7b that the BS-2# sample has the highest photocatalytic efficiency, while 70.5% and 90.1% degradation rates of RhB are observed for BS-4# and BS-5# sample, respectively. Figure 7c shows that the degradation rate of RhB over the as-prepared 3D hierarchical Bi2S3 architectures increased with increasing pH value, then decreased. Thus, the factors of temperature, time and pH value exert a significant effect on the photocatalytic efficiency of the as-prepared 3D hierarchical Bi2S3 architectures. The explanation can be summarized as follows. The crystallinity and crystalline size vary with the changes of hydrothermal temperature, time and pH value, implying that crystal defects and bandgaps will change correspondingly. The improvement of crystallinity implies that crystal defects acting as photogenerated electron–hole recombination centers will decrease in photocatalysts, while the spectral absorption ability will weaken with increasing crystalline size. Thus, the variation of photocatalytic efficiency of photocatalysts is mainly ascribed to the synergistic effect of the crystallinity and crystalline size. As is shown in Figure 7d, the degradation efficiency of the as-prepared Bi2S3 nanorods is only 63.1%. Obviously, the BS-2# sample exhibits much better photocatalytic activity than Bi2S3 nanorods. Figure 8 shows the variations of absorption spectra of RhB aqueous solution in the presence of the BS-2# sample and Bi2S3 nanorods. It can be found from Figure 8 that the intensity of the characteristic absorption peak of RhB decreased with the irradiation time prolonging. Meanwhile, the color of the suspension faded away gradually during the degradation process. The RhB aqueous solution was almost degraded by the BS-2# photocatalyst after irradiating for 80 min, while the photocatalytic activity of Bi2S3 nanorods was more inferior to that of the BS-2# photocatalyst.
It is generally accepted that photocatalytic activities are strongly dependent on the structural features of photocatalysts [39,40]. In this work, ultra thin Bi2S3 nanoplatelets can provide large specific surface area, which leads to enhanced absorption and light harvesting. In addition, nanoplatelets with ultra thickness will accelerate the transfer rate of photo-generated carriers from the inside to the surface of the crystal, in which they react with the organic molecules. Thus, the ultrathin structure facilitates the prompt separation of the photo-generated electron–hole, and the probability of the photo-generated electron–hole pairs will be reduced in turn. Base on the above demonstration, 3D hierarchical Bi2S3 architectures have porous structure, provide more active sites at large surface area for adsorbing reactant molecules and photocatalytic reactions, and present enhanced photo-generated electron–hole separation rate [7,27,28]. Accordingly, the representative 3D hierarchical Bi2S3 architectures exhibits higher photocatalytic efficiency, revealing that the photocatalytic activity of 3D hierarchical Bi2S3 architectures are strongly related to the factors of crystallinity, shape and surface area, and the effects of these factors on the photoactivity are interdependent.
To make the photodegradation reaction clear, ammonium oxalate (AO), methanol (ME), and benzoquinone (BQ) were used as scavenges of active holes (h+), hydroxyl radicals (•OH), and superoxide radicals (•O2) to investigate the effects of reactive species on the photodegradation process of RhB solution [41]. As shown in Figure 9, both AO and BQ cause a strong suppression on the photodegradation process of RhB solution while ME exhibited a relatively weak effect. The above results reveals that the reactive species of h+ and •O2 played a more significant role than •OH for the photodegradation of RhB, which is similar to the previous studies [42,43]. Because the stability has always been a critical concern for an excellent photocatalyst, it is essential to investigate the stability and repeatability of 3D hierarchical Bi2S3 architectures in the photodegradation of RhB solution. Thus, the BS-2# sample was recycled for five times under the same conditions, and the corresponding results are shown in Figure 10. The photocatalytic activity of the BS-2# sample exhibits no significant loss after five cycling runs of photodegradation of RhB, indicating that the BS-2# sample is an efficient visible-light-driven photocatalyst, good reusability and stability for potential practical application in wastewater treatment.

3. Experimental Section

3.1. Synthesis of 3D Hierarchical Bi2S3 Architectures

All the reagents obtained from Shanghai chemical industrial Co. Ltd. China were analytically pure and used without any further purification. In a typical synthesis procedure, 2 mmol of bismuth nitrate pentahydrate (Bi(NO3)3·5H2O) were dissolved in 20 mL mixture solution under vigorous stirring, which consisted of deionized water (8 mL) and dilute nitric acid (1M 12 mL). Then, 0.2 mmol of cetyltrimethyl ammonium bromide (CTAB) was added into the above solution. Next, 3 mmol of thioacetamide (Tu) was dissolved into 40 mL deionized water under stirring to form a homogeneous solution. After that, the Tu solution was slowly dropped into the above bismuth nitrate solution under constant magnetic stirring at room temperature. Subsequently, the pH value of the mixing solution was adjusted to be a certain value. After 1 h, the resulting mixture solution was transferred to Teflon-line stainless steel autoclave with 100 mL capacity and held at different temperatures for a certain time. After the autoclave was cooled naturally to room temperature, the final black precipitates were centrifuged and washed with deionized water and alcohol, respectively, followed by drying at 60 °C for 10 h in a vacuum oven. The detailed experimental conditions for the preparation of 3D hierarchical Bi2S3 architectures are listed in Table 1. As a comparison, Bi2S3 nanorods were also prepared by reacting Bi(NO3)3 and Tu without CTAB.

3.2. Characterization

Powder X-ray diffraction (XRD) measurements of the as-prepared 3D hierarchical Bi2S3 architectures were carried out with a Bruker AXS D8 Advance diffractometer with Cu Kα radiation (λ = 0.15406nm) over a range from 10° to 70° (Bruker, Billerica, MA ,USA). Surface morphology was observed by using a Hitachi S-4800 (Hitachi, Toyko, Japan) field emission scanning electron microscope. Transmission electron microscopy (TEM) was obtained by using a JEM-2100 microscope (JEOL Ltd., Tokyo, Japan). The Brunauer–Emmett–Teller (BET) surface area was determined on the basis of nitrogen adsorption and desorption isotherms at 77 K using a Quantachrome NOVA 2000e sorption analyzer (Quantachrome Instruments, Boynton Beach, FL, USA). Raman spectrum were carried out by using a Renishaw invia spectrometer (Renishaw, London, UK) at room temperature, and the excitation source was an Ar+ laser with 514 nm line. UV-vis diffuse reflectance spectrum was collected from a UV-vis spectrophotometer (TU 1901, Puxi, Beijing, China) with BaSO4 as reference.

3.3. Photocatalytic Activity

The photocatalytic activities of the as-prepared Bi2S3 samples were evaluated by the degradation of RhB aqueous solution under visible-light irradiation. As was reported in our previous work [29,30,31], 40 mL RhB aqueous solution with a concentration of 1.0 × 10−5 M was added into the culture dish with 0.05 g prepared photocatalyst. A 30 W fluorescent lamp was used as light source to trigger the photocatalytic reaction, and the distance between the lamp and the dish coated by the photocatalysts was about 10 cm. Prior to irradiation, an adsorption–desorption equilibrium was demanded to reach among the catalyst, RhB and water in the dark. At given time intervals, 5 mL of the reacted solution was centrifuged for concentration analysis. The concentration of RhB aqueous solution was determined by recording the variation of the intensity of absorption peak located at 554 nm using an UV-visible spectrophotometer (TU1901, Puxi, Beijing, China). The photo-degradation efficiency of RhB was evaluated by (C0-C)/C0, where C is the concentration of the RhB solution at reaction time t, C0 is the adsorption/desorption equilibrium concentration of RhB (at reaction time 0).

4. Conclusions

In summary, we demonstrated a facile hydrothermal route for preparing 3D hierarchical Bi2S3 architectures. The as-prepared 3D hierarchical Bi2S3 architectures in which abundant nanoplatelets are randomly arranged and interconnected with each other have porous structure and large specific surface area. The formation process of 3D hierarchical Bi2S3 architectures was proposed. The enhanced photocatalytic activity of the representative 3D hierarchical Bi2S3 architectures could be ascribed to the synergetic effect of the shape, surface area, crystallinity, band gap and crystalline size. Considering the facile synthesis method and superior photocatalytic activity, the as-prepared 3D hierarchical Bi2S3 architectures are promising to speed up the application of visible-light photocatalysts in the field of environmental remediation.

Acknowledgments

The authors express grateful thanks to the National Natural Science Foundation of China (Grant U1304520, and U1404613), Education Department of Henan Province for the fund support (2013JJGS-185), and the program for New Century Excellent Talents in University (NECT-12-0655).

Author Contributions

Tiekun Jia conceived and designed the experiments; Jili Li and Zhihua Kang performed the synthetic experiment; Xiaofeng Wang and Fei Long analyzed the data; Guang Sun gave basic and important ideas as well as precise instruction; Fang Fu and Jian Chen performed the photo-catalytic activity of the as-prepared samples; and Tiekun Jia wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
XRDX-ray diffraction
SEMScanning electron microscopy
TEMTransmission electron microscopy
HRTEMHigh resolution transmission electron microscopy
BETBrunauer-Emmett-Teller
DRSDiffuse reflectance spectrum
1DOne dimensional
2DTwo-dimensional
3DThree-dimensional
RhBRhodamine B
CTABCetyltrimethyl ammonium bromide
BETBrunauer-Emmett-Teller
DRSDiffuse reflectance spectrum

References

  1. Zou, Z.G.; Ye, J.H.; Sayama, K.; Arakawa, H. Direct splitting of water under visible light irradiation with an oxide semiconductor photocatalyst. Nature 2001, 414, 625–627. [Google Scholar] [CrossRef] [PubMed]
  2. Hernandez-Alonso, M.D.; Fresno, F.; Suarez, S.; Cornnado, J.M. Development of alternative photocatalysts to TiO2: Challenges and opportunities. Energy Environ. Sci. 2009, 2, 1231–1257. [Google Scholar] [CrossRef]
  3. Zhang, Z.G.; Huang, Z.F.; Cheng, X.D.; Wang, Q.L.; Chen, Y.; Dong, P.M.; Zhang, X.W. Product selectivity of visible-light photocatalytic reduction of carbon dioxide using titanium dioxide doped by different nitrogen-sources. Appl. Sur. Sci. 2015, 355, 45–51. [Google Scholar] [CrossRef]
  4. Tu, W.G.; Zhou, Y.; Zou, Z.G. Photocatalytic conversion of CO2 into renewable hydrocarbon rules: State-of-the-art accomplishment, challenges, and prospects. Adv. Mater. 2014, 26, 4607–4626. [Google Scholar] [CrossRef] [PubMed]
  5. Truong, Q.D.; Le, T.H.; Liu, J.Y.; Chung, C.C.; Ling, Y.C. Synthesis of TiO2 nanoparticles using novel titanium oxalate complex towards visible light-driven photocatalytic reduction of CO2 to CH3OH. Appl. Cata. A Gen. 2012, 437, 28–35. [Google Scholar] [CrossRef]
  6. Paola, A.D.; Bellardita, M.; Palmisano, L. Brookite, the least known TiO2 photocatalyst. Catalysts 2013, 3, 36–73. [Google Scholar] [CrossRef]
  7. Liu, B.S.; Nakata, K.; Sakai, M.; Saito, M.; Ochiai, T.; Murakami, T.; Takagi, K.; Fujishima, A. Hierarchical TiO2 spherical nanostructures with tunable pore size, pore volume, and specific surface area: Facile preparation and high-photocatalytic performance. Catal. Sci. Technol. 2012, 2, 1933–1939. [Google Scholar] [CrossRef]
  8. Rath, A.K.; Bernechea, M.; Martinez, L.; Konstantatos, G. Solution-processed heterojunction solar cells based on p-type PbS quantum dots and n-type Bi2S3 nanocrystals. Adv. Mater. 2011, 23, 3712–3717. [Google Scholar] [CrossRef] [PubMed]
  9. Zhang, H.; Yang, L.; Liu, Z.; Ge, M.; Zhou, Z.; Chen, W.; Li, Q.; Liu, L. Facet-dependent activity of bismuth sulfide as low-cost counter-electrode materials for dye-sensitized solar cells. J. Mater. Chem. 2012, 22, 18572–18577. [Google Scholar] [CrossRef]
  10. Li, Y.P.; Wei, F.; Ma, Y.G.; Zhang, H.; Gao, Z.W.; Dai, L.; Qin, G.G. Selected-control hydrothermal synthesis and photoresponse properties of Bi2S3 micro/nanocrystals. CryEngComm 2013, 15, 6611–6616. [Google Scholar] [CrossRef]
  11. Li, R.; Yang, J.; Fan, C.; Lu, F.; Wei, Z.; Li, J. Effect of electrical contact on the performance of Bi2S3 nanowire photodetectors. ChemPhysChem 2014, 15, 2510–2516. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, G.H.; Yu, Y.Q.; Zheng, K.; Ding, T.; Wang, W.L.; Jiang, Y.; Yang, Q. Fabrication of ultrathin Bi2S3 nanosheets for high performance, flexible, visible-NIR photodetectors. Small 2015, 11, 2848–2855. [Google Scholar] [CrossRef] [PubMed]
  13. Xiao, G.J.; Dong, Q.F.; Wang, Y.N.; Sui, Y.M.; Ning, J.J.; Liu, Z.Y.; Tian, W.J.; Liu, B.B.; Zou, G.T.; Zou, B. One-step solution synthesis of bismuth sulfide (Bi2S3) with various hierarchical architectures and their photoresponse properties. RSC Adv. 2012, 2, 234–240. [Google Scholar] [CrossRef]
  14. Li, H.H.; Yang, J.; Zhang, J.Y.; Zhou, M. Facile synthesis of hierarchical Bi2S3 nanostructures for photo-detector and gas sensor. RSC Adv. 2012, 2, 6258–6261. [Google Scholar] [CrossRef]
  15. Schricker, A.D.; Sigman, M.B.; Korgel, B.A. Electrical transport, Meyer-Neldel rule and oxygen sensitivity of Bi2S3 nanowire. Nanotechnology 2005, 16, S508–S513. [Google Scholar] [CrossRef]
  16. Rabin, O.; Perez, J.M.; Grimm, J.; Weissleder, R. An X-ray computed tomography imaging agent based on long-circulating bismuth sulphide nanoparticles. Nat. Mater. 2006, 5, 118–122. [Google Scholar] [CrossRef] [PubMed]
  17. Ma, J.M.; Liu, Z.F.; Lian, J.B.; Duan, X.C.; Kim, T.; Peng, P.; Liu, X.; Yao, G.; Zheng, W.J. Ionic liquids-assisted synthesis and electrochemical properties of Bi2S3 nanostructures. CrystEngComm 2011, 13, 3072–3079. [Google Scholar] [CrossRef]
  18. Jin, R.C.; Xu, Y.B.; Li, G.H.; Liu, J.S.; Chen, G. Hierarchical chlorophytum-like Bi2S3 architectures with high electrochemical performance. Int. J. Hydroge. Energy 2013, 38, 9137–9144. [Google Scholar] [CrossRef]
  19. Zhang, Z.; Zhou, C.K.; Lu, H.; Jia, M.; Lai, Y.Q.; Li, J. Facile synthesis of dandelion-like Bi2S3 microspheres and their electrochemical properties for lithium-ion batteries. Mater. Lett. 2013, 91, 100–102. [Google Scholar] [CrossRef]
  20. Jung, H.; Park, C.M.; Sohn, H.J. Bismuth sulfide and its carbon nanocomposite for rechargeable lithium-ion batteries. Electrochim. Acta 2011, 56, 2135–2139. [Google Scholar] [CrossRef]
  21. Vinu, R.; Madras, G. Kinetics of simultaneous photocatalytic degradation of phenolic compounds and reduction of metal ions with nano-TiO2. Environ. Sci. Technol. 2008, 42, 913–919. [Google Scholar] [CrossRef] [PubMed]
  22. Cao, J.; Xu, B.Y.; Lin, H.L.; Luo, B.; Chen, S.F. Chemical etching preparation of BiOI/BiOBr heterostructures with enhanced photocatalytic, properties for organic dye removal. Chem. Eng. J. 2012, 185, 91–99. [Google Scholar] [CrossRef]
  23. Cui, Y.M.; Jia, Q.F.; Li, H.Q.; Han, H.G.; Zhu, L.J.; Li, S.G.; Zou, Y.; Yang, J. Photocatalytic activities of Bi2S3/BiOBr nanocomposites synthesized by a facile hydrothermal process. Appl. Surf. Sci. 2014, 290, 233–239. [Google Scholar] [CrossRef]
  24. Sigman, M.B., Jr.; Korgel, B.A. Solventless synthesis of Bi2S3 (bismuthinite) nanorods, nanowires, and nanofabric. Chem. Mater. 2005, 17, 1655–1660. [Google Scholar] [CrossRef]
  25. Liu, Z.P.; Peng, S.; Xie, Q.; Hu, Z.; Yang, Y.; Zhang, S.; Qian, Y. Large-scale synthesis of ultra long Bi2S3 nanoribbons via a solvothermal process. Adv. Mater. 2003, 15, 936–940. [Google Scholar] [CrossRef]
  26. Zhou, H.Y.; Xiong, S.L.; Wei, L.Z.; Xi, B.J.; Zhu, Y.C.; Qian, Y.T. Acetylacetone-directed controllable synthesis of Bi2S3 nanostructures with tunable morphology. Cryst. Growth Des. 2009, 9, 3862–3867. [Google Scholar] [CrossRef]
  27. Li, G.; Liu, J.; Lan, J.; Li, G.; Chen, Q.; Jiang, G. 3D hierarchical anatase TiO2 superstructures constructed by “nanobricks” built nanosheets with exposed{001}facets: Facile synthesis, formation mechanism and superior photocatalyticactivity. CrystEngComm 2014, 16, 10547–10552. [Google Scholar] [CrossRef]
  28. Zhang, H.; Du, G.; Lu, W.; Cheng, L.; Zhu, X.; Jiao, Z. Porous TiO2 hollow nanospheres: Synthesis, characterization and enhanced photocatalytic properties. CrystEngComm 2012, 14, 3793–3796. [Google Scholar] [CrossRef]
  29. Jia, T.K.; Wang, W.M.; Long, F.; Fu, Z.Y.; Wang, H.; Zhang, Q.J. Synthesis, characterization, and photocatalytic activity of Zn-doped SnO2 hierarchical architectures assembled by nanocones. J. Phys. Chem. C 2009, 113, 9701–9077. [Google Scholar] [CrossRef]
  30. Jia, T.K.; Wang, W.M.; Long, F.; Fu, Z.Y.; Wang, H.; Zhang, Q.J. Fabrication, Characterization and Photocatalytic Activity of La-doped ZnO Nanowires. J. Alloys Compd. 2009, 484, 410–415. [Google Scholar] [CrossRef]
  31. Jia, T.K.; Fu, F.; Long, F.; Min, Z.Y.; Zhao, J.W.; Chen, J.; Li, J.L. Synthesis, characterization andenhanced visible-light photocatalytic activity of Zn2SnO4-C nanocomposites with truncated octahedron morphology. Ceram. Int. 2016, 142, 13893–13899. [Google Scholar] [CrossRef]
  32. Helal, A.; Harraz, F.A.; Iamail, A.A.; Sami, T.M.; Ibrahim, I.A. Controlled synthesis of bismuth sulfide nanorods by hydrothermal method and their photocatalytic activity. Mater. Design 2016, 102, 202–212. [Google Scholar] [CrossRef]
  33. Zhang, Y.Y.; Guo, Y.P.; Fang, B.J.; Chen, Y.J.; Duan, H.N.; Li, H. Continuously enhanced photoactivity of hierarchical β–Bi2O3/Bi2S3 heterostructure derived from novel BiO2CH3 octagonal nanoplates. Appl. Catal. A Gen. 2016, 514, 146–153. [Google Scholar] [CrossRef]
  34. Wang, X.; Li, S.; Yu, H.; Yu, J. In situ anion-exchange synthesis and photocatalytic activity of Ag8W4O16/AgCl-nanoparticle core-shell nanorods. J. Mol. Catal. A Chem. 2011, 334, 52–59. [Google Scholar] [CrossRef]
  35. Yuan, Z.Y.; Su, B.L. Surfactant-assisted nanoparticle assembly of mesoporous β-FeOOH (akaganeite). Chem. Phys. Lett. 2003, 381, 259–267. [Google Scholar]
  36. Ren, T.Z.; Yuan, Z.Y.; Su, B.L. Surfactant-assisted preparation of hollow microspheres of mesoporous TiO2. Chem. Phys. Lett. 2003, 374, 170–175. [Google Scholar] [CrossRef]
  37. Butler, M.A. Hotoelectrolysis and physical properties of the semiconducting electrode WO3. J. Appl. Phys. 1977, 48, 1914–1920. [Google Scholar] [CrossRef]
  38. Miller, D.D.; Heller, A. Semiconductor liquid junction solar cells based on anodic sulphide films. Nature 1976, 262, 680–681. [Google Scholar] [CrossRef]
  39. Lu, H.B.; Wang, S.M.; Zhao, L.; Li, J.C.; Dong, B.H.; Xu, Z.X. Hierarchical ZnO microarchitectures assembled by ultrathin nanosheets: Hydrothermal synthesis and enhanced photocatalytic activity. J. Mater. Chem. 2011, 21, 4228–4234. [Google Scholar] [CrossRef]
  40. Zhang, L.; Yu, J.C. A snochemical approach hierarchical porous titania spheres with enhanced photocatalytic activity. Chem. Comm. 2003, 2078–2079. [Google Scholar] [CrossRef]
  41. Di, J.; Xia, J.X.; Ji, M.X.; Wang, B.; Li, X.W.; Zhang, Q.; Chen, Z.G.; Li, H.M. Nitrogen-doped carbon quantum dots/BiOBr ultrathin nanosheets: In Situ strong coupling and improved molecular oxygen activation ability under visible light irradiation. ACS Sustain. Chem. Eng. 2016, 4, 136–146. [Google Scholar] [CrossRef]
  42. Cao, J.; Xu, B.Y.; Lin, H.L.; Luo, B.D.; Chen, S.F. Novel Bi2S3-sensitized BiOCl with highly visible light photocatalytic activity for the removal of Rhodamine B. Catal. Comm. 2012, 26, 204–208. [Google Scholar] [CrossRef]
  43. Chen, X.; Li, H.K.; Wu, Y.X.; Wu, H.S.; Wu, L.D.; Tan, P.F.; Pan, J.; Xiong, X. Facile fabrication of novel porous graphitic carbon nitride/copper sulfide nanocomposites with enhanced visible light driven photocatalytic performance. J. Coll. Inter. Sci. 2016, 476, 132–143. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD patterns of 3D hierarchical Bi2S3 architectures prepared at different hydrothermal temperatures; and (b) the Raman spectrum of the representative 3D hierarchical Bi2S3 architectures (BS-2#).
Figure 1. (a) XRD patterns of 3D hierarchical Bi2S3 architectures prepared at different hydrothermal temperatures; and (b) the Raman spectrum of the representative 3D hierarchical Bi2S3 architectures (BS-2#).
Crystals 06 00140 g001
Figure 2. (a,b) Low magnification and higher magnification FESEM image of the BS-2# sample; (c) SEM-EDS elemental mapping of the the BS-2# sample; and the corresponding elemental mappings of (d) S and (e) Bi.
Figure 2. (a,b) Low magnification and higher magnification FESEM image of the BS-2# sample; (c) SEM-EDS elemental mapping of the the BS-2# sample; and the corresponding elemental mappings of (d) S and (e) Bi.
Crystals 06 00140 g002
Figure 3. (a,b) Low-magnification and higher magnification TEM images of the BS-2# sample; (c) a typical HRTEM image taken from the single nanoplatelet; and (d) TEM image of Bi2S3 nanorods.
Figure 3. (a,b) Low-magnification and higher magnification TEM images of the BS-2# sample; (c) a typical HRTEM image taken from the single nanoplatelet; and (d) TEM image of Bi2S3 nanorods.
Crystals 06 00140 g003
Figure 4. Schematic illustrating the formation process of 3D hierarchical Bi2S3 architectures.
Figure 4. Schematic illustrating the formation process of 3D hierarchical Bi2S3 architectures.
Crystals 06 00140 g004
Figure 5. (a) Nitrogen adsorption–desorption isotherm; and (b) the corresponding pore size distribution of the BS-2# sample.
Figure 5. (a) Nitrogen adsorption–desorption isotherm; and (b) the corresponding pore size distribution of the BS-2# sample.
Crystals 06 00140 g005
Figure 6. (a) UV-vis diffuse reflectance spectrum of the BS-2# sample; and (b) the corresponding (αhν)2 vs. photon energy (hν) plots of the BS-2# sample.
Figure 6. (a) UV-vis diffuse reflectance spectrum of the BS-2# sample; and (b) the corresponding (αhν)2 vs. photon energy (hν) plots of the BS-2# sample.
Crystals 06 00140 g006
Figure 7. (a) The photocatalytic activities of RhB solution in the dark, without catalyst and over the samples of P25, BS-1#, BS-2# and BS-3#; (b) the comparison of photocatalytic activity over the samples of BS-2#, BS-4# and BS-5#; (c) the comparison of photocatalytic activity of over the samples of BS-2#, BS-6# and BS-7#; and (d) the comparison of photocatalytic activity over the samples of BS-2# and Bi2S3 nanorods.
Figure 7. (a) The photocatalytic activities of RhB solution in the dark, without catalyst and over the samples of P25, BS-1#, BS-2# and BS-3#; (b) the comparison of photocatalytic activity over the samples of BS-2#, BS-4# and BS-5#; (c) the comparison of photocatalytic activity of over the samples of BS-2#, BS-6# and BS-7#; and (d) the comparison of photocatalytic activity over the samples of BS-2# and Bi2S3 nanorods.
Crystals 06 00140 g007
Figure 8. The variations of adsorption spectra of aqueous RhB over different phototocatalysts: (a) the BS-2# sample; and (b) Bi2S3 nanorods.
Figure 8. The variations of adsorption spectra of aqueous RhB over different phototocatalysts: (a) the BS-2# sample; and (b) Bi2S3 nanorods.
Crystals 06 00140 g008
Figure 9. The effects of reactive species involved in the photodegradation of RhB solution over the BS-2# sample.
Figure 9. The effects of reactive species involved in the photodegradation of RhB solution over the BS-2# sample.
Crystals 06 00140 g009
Figure 10. Cycling performance of the photodegradation of RhB solution over the BS-2# sample.
Figure 10. Cycling performance of the photodegradation of RhB solution over the BS-2# sample.
Crystals 06 00140 g010
Table 1. Summary of 3D hierarchical Bi2S3 architectures experimental conditions.
Table 1. Summary of 3D hierarchical Bi2S3 architectures experimental conditions.
SampleTemperature (°C)Reaction time (h)pH value
BS-1#16582.4
BS-2#165122.4
BS-3#165162.4
BS-4#105122.4
BS-5#135122.4
BS-6#165121.0
BS-7#165124.0

Share and Cite

MDPI and ACS Style

Jia, T.; Wang, X.; Long, F.; Li, J.; Kang, Z.; Fu, F.; Sun, G.; Chen, J. Facile Synthesis, Characterization, and Visible-light Photocatalytic Activities of 3D Hierarchical Bi2S3 Architectures Assembled by Nanoplatelets. Crystals 2016, 6, 140. https://doi.org/10.3390/cryst6110140

AMA Style

Jia T, Wang X, Long F, Li J, Kang Z, Fu F, Sun G, Chen J. Facile Synthesis, Characterization, and Visible-light Photocatalytic Activities of 3D Hierarchical Bi2S3 Architectures Assembled by Nanoplatelets. Crystals. 2016; 6(11):140. https://doi.org/10.3390/cryst6110140

Chicago/Turabian Style

Jia, Tiekun, Xiaofeng Wang, Fei Long, Jili Li, Zhihua Kang, Fang Fu, Guang Sun, and Jian Chen. 2016. "Facile Synthesis, Characterization, and Visible-light Photocatalytic Activities of 3D Hierarchical Bi2S3 Architectures Assembled by Nanoplatelets" Crystals 6, no. 11: 140. https://doi.org/10.3390/cryst6110140

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

Article Metrics

Back to TopTop