Next Article in Journal
Improved Brønsted to Lewis (B/L) Ratio of Co- and Mo-Impregnated ZSM-5 Catalysts for Palm Oil Conversion to Hydrocarbon-Rich Biofuels
Previous Article in Journal
Extremely Stable and Durable Mixed Fe–Mn Oxides Supported on ZrO2 for Practical Utilization in CLOU and CLC Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of a Novel 1D/2D Bi2O2CO3–BiOI Heterostructure and Its Enhanced Photocatalytic Activity

1
The Key Laboratory of Functional Molecular Solids, Ministry of Education, College of Chemistry and Materials Science, Anhui Normal University, Wuhu 241000, China
2
Anhui Provincial Engineering Laboratory of Water and Soil Pollution Control and Remediation, School of Ecology and Environment, Anhui Normal University, Wuhu 241002, China
3
College of International Education, Anhui Normal University, Wuhu 241000, China
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(11), 1284; https://doi.org/10.3390/catal11111284
Submission received: 28 September 2021 / Revised: 16 October 2021 / Accepted: 19 October 2021 / Published: 25 October 2021

Abstract

:
A novel 1D/2D Bi2O2CO3–BiOI heterojunction photocatalyst with high-quality interfaces was synthesized through a hydrothermal method by using Bi2O2CO3 nanorods and KI as raw materials. Two-dimensional (2D) BiOI nanosheets uniformly and vertically grow on the 1D porous Bi2O2CO3 rods. Bi2O2CO3–BiOI heterojunctions exhibit better photocatalytic activity than pure Bi2O2CO3 nanorods and BiOI nanosheets. Cr(VI) (30 mg/L), MO (20 mg/L) and BPA (20 mg/L) can be completely degraded in 8–15 min. The superior photocatalytic performance of 1D/2D Bi2O2CO3–BiOI heterojunction is ascribed to the synergistic effects: (a) vertical 2D on 1D multidimensional structure; (b) the formation of the Bi2O2CO3–BiOI p–n heterojunction; (c) high-quality interfaces between Bi2O2CO3 and BiOI.

1. Introduction

Nowadays, one-dimensional (1D)/two-dimensional (2D) heterogeneous photocatalysts have been extensively studied for their unique dimensional advantages in environmental and sustainable energy applications [1,2,3]. A large number of studies have shown that 1D nanostructures can provide a short diffusion length perpendicular to its axis and a fast carrier transfer path along its axis [4,5,6,7]. Two-dimensional (2D) materials usually have a large surface area with good electrical conductivity and superior electron mobility [8,9,10,11]. So, compared with other heterostructures, 1D/2D heterostructures have unique advantages because of their intrinsic structural features [12,13,14,15]. Firstly, 1D/2D heterojunction has a larger surface/interface area, which can provide more photocatalytic active sites. Secondly, the photogenerated charge carriers on 2D nanosheets will transfer to 1D nanorods, which make electron–hole pairs separation more effectively and charge lifetimes longer [12,16,17,18]. At present, there are mainly two types of 1D/2D heterojunction based on the interfacial contact and morphology: the growth of 1D material on 2D nanomaterials and the growth of 2D nanosheets on 1D nanomaterials. Among them, the vertical epitaxial growth of 2D nanosheets on 1D nanomaterial can expose most of their overall surface [19,20]. Although 1D/2D heterojunction materials have many advantages, there are very few studies on bismuth-based 1D/2D heterojunction materials at present [21,22,23]. In recent years, our group has reported the synthesis of photocatalysts which 2D nanosheets vertically loaded on 1D bismuth-based materials [19,20,24,25]. These 1D/2D bismuth-based heterojunction photocatalysts exhibit superior photocatalytic activity in the removal of organic pollutants. However, the research in this field is still insufficient. So, it is necessary to extensively explore the synthesis of this kind of 1D/2D bismuth-based heterojunction photocatalysts with excellent photocatalytic performance.
As known, the unique layered structure of bismuth-based materials makes it has the characteristics of high connection quality between interfaces, internal electric field formation, asymmetric polarization effect [26,27,28,29]. Bi2O2CO3 is a typical n-type semiconductor with wide band gap (3.55 eV) and only responds to ultraviolet light [30,31,32]. BiOI is a typical p-type semiconductor with narrow band gap (1.8–2.1 eV) [33,34,35,36]. BiOI is widely used to eliminate the pollutants because it can improve the absorption capacity of visible light [34,37,38]. BiOI and Bi2O2CO3 have the similar layered structure which can couple into p-n typical heterostructure [39,40,41,42,43,44,45,46,47,48]. All of the above BiOI–Bi2O2CO3 heterostructures display improved activities compared to pristine BiOI and Bi2O2CO3. However, the reported BiOI–Bi2O2CO3 photocatalysts show the 0D/2D, 0D/3D or 2D/2D coupling of BiOI with Bi2O2CO3, and 1D/2D Bi2O2CO3–BiOI heterojunction has not been reported yet.
In this article, a novel 1D/2D Bi2O2CO3–BiOI heterojunction was synthesized through a hydrothermal method by using porous Bi2O2CO3 nanorods as Bi source and KI as I source. BiOI nanosheets vertically grow on the Bi2O2CO3 porous rods. This 1D/2D Bi2O2CO3–BiOI heterostructure displays superior photocatalytic activity for degrading Cr (VI), methyl orange (MO) and bisphenol A (BPA) under solar light irradiation, and Cr (VI) (30 mg/L) could be completely reduced in 8 min. This excellent photocatalytic performance is due to the synergistic effect of some factors: (a) Unique 1D/2D nanostructure; (b) the formation of the p–n junction; (c) the high-quality interfaces between Bi2O2CO3 and BiOI.

2. Results and Discussion

1D/2D Bi2O2CO3–BiOI heterojunction was synthesized by hydrothermal method using porous Bi2O2CO3 nanorods and KI as raw materials. The as-obtained Bi2O2CO3–BiOI heterojunctions were labelled as S1, S2, S3 and S4 when the molar ratios of Bi2O2CO3: KI are 10:1, 2:1,1:2 and 1:20, respectively. The SEM images and XRD pattern of the obtained samples are shown in Figure 1. Bi2O2CO3 is 1D rod-like structure with rough surfaces and porosity (Figure 1a). Few nanosheets grow vertically on the surface of Bi2O2CO rods when the molar ratio of Bi2O2CO3: KI is 10:1 (Figure 1b). With the molar ratio of Bi2O2CO3: KI increase, more and more nanosheets are loaded on the surface of Bi2O2CO rods (Figure 1b–d). When the molar ratio of Bi2O2CO3: KI is 1:20, only out-of-order nanosheets can be obtained (Figure 1e). Figure 1f shows the XRD patterns of the obtained samples. It can be seen that the main diffraction peaks in S1–S3 samples are indexed to Bi2O2CO3 (PDF#25-1464). In samples S2 and S3, a new diffraction peak (2θ = 31.65°) is found clearly, which is indexed to the tetragonal BiOI (PDF#10-0445). The diffraction peaks of Bi2O2CO3 become weaker and weaker from S1 to S3. The results indicate Bi2O2CO3 reacts with KI and form 1D/2D Bi2O2CO3–BiOI heterostructure. In S4 sample, only BiOI peaks can be found, indicating all the Bi2O2CO3 nanorods is consumed completely. The Bi2O2CO3 nanorod skeleton is disappeared, which results in the collapse of 1D structure. So, BiOI nanosheets is formed when the molar ratio of Bi2O2CO3:KI is 1:20.
The S2 sample was characterized by using TEM and HRTEM. As shown in Figure 2a, it can be seen that BiOI nanosheets grow vertically on the surface of Bi2O2CO3 nanorods, which is consistent with the SEM images. In Figure 2b, the lattice spacing of 0.685 and 0.915 nm indexes to the (002) lattice plane of Bi2O2CO3 and the (001) lattice plane of BiOI, respectively. BiOI nanosheets grow out from the Bi2O2CO3 rod by oriented epitaxial nucleation and growth, which is beneficial for the formation of a high-quality interface [24].
The full XPS spectra of S2, S4 and Bi2O2CO3 nanorods is shown in Figure 3a. Bi, C, O and I elements are co-existence in S2 sample, indicating the formation of Bi2O2CO3–BiOI. There is almost no C element in the S4 sample, confirming that S4 is pure BiOI. The high resolution XPS spectra of the Bi 4f, I3d and C1s are shown in Figure 3b–d. Two peaks centered at about 164.50 and 159.00 eV are attributed to Bi 4f7/2 and Bi 4f5/2, respectively, indicating that the Bi element in all samples is in the form of Bi3+ ion (Figure 3b). Compared with Bi2O2CO3 nanorods and S4 (pure BiOI), the peak of Bi 4f moves to higher binding energies, which proves that high-quality interface forms between Bi2O2CO3 and BiOI. The same phenomenon is also observed in the I 3d spectra (Figure 3c). The peak centered at 289.08 eV is indexed to C1s (Figure 3d). The intensity of C1s in S2 sample is obviously lower than in Bi2O2CO3. This result confirms Bi2O2CO3 is consumed to form BiOI during the reaction.
The FTIR spectra of Bi2O2CO3, S2 and S4 samples were measured, as shown in Figure 4. The intensive peak which centered at 847.1 cm−1 is attributed to the ν2 modes of the CO32− group. After Bi2O2CO3 and BiOI are combined with each other, a new peak at 493.6 cm−1 appears in S2 sample, further proving the formation of the strong interfacial junction between Bi2O2CO3 and BiOI. Depending on the increase of the loaded content of BiOI, the main characteristic peak of the CO32− (847.1 cm1) almost disappears in the S4. This result is consistent with XRD and XPS results, verifying that the S4 sample is pure BiOI.
UV-vis diffuse reflectance spectroscopies (DRS) of Bi2O2CO3, S2 and S4 were performed to study their optical absorption properties. The absorption band edge of Bi2O2CO3 and BiOI are ~450 and 670 nm (Figure 5a), respectively, indicating the wider band gap of Bi2O2CO3 than that of BiOI. The absorption band edge of the S2 heterostructure red-shifts compared with Bi2O2CO3 due to the loaded-BiOI with narrower band gap (Figure 5b). The optical band gap of Bi2O2CO3 and BiOI is obtained using the following equation:
αhν = A(Eg)n/2
where α is the absorption coefficient, h is Planck’s constant, ν is the light frequency, A is the constant and Eg is the bandgap energy [49]. In our study, both Bi2O2CO3 and BiOI possess indirect band gaps, so n = 4 [25,50]. The band gap energies are estimated to 2.96 eV for pure Bi2O2CO3, and 1.75 eV for pure BiOI (Figure 5b).
The photocatalytic activity of the Bi2O2CO3–BiOI heterostructures are tested using the Cr (VI) (30 mg/L), MO (20 mg/L) and BPA (20 mg/L) as model pollutants under solar light irradiation. The degradation curves of the different photocatalysts and the UV-vis absorption spectra of Cr (VI), MO and BPA are shown in Figure 6. The results show that S2 owns the highest photocatalytic activity among all the samples, and Cr(VI) (30 mg/L) (pH = 7), MO (20 mg/L) and BPA (20 mg/L) can be completely photodegraded in 8, 15 and 15 min, respectively. Compared with the reported Bi2O2CO3–BiOI heterostructures, S2 sample exhibits excellent photocatalytic activity [39,40]. From Figure S1, the reaction rate constant (k) of S2 (0.4661 min−1) is much higher than that of Bi2O2CO3, S1, S3 and S4 (0.0049, 0.0731, 0.1363 and 0.0594 min−1) in degrading Cr (VI), exhibiting S2 sample superior photocatalyst. The reaction rate constant (k) of S2 is also higher than that of Bi2O2CO3, S1, S3 and S4 in degrading MO (20 mg/L) and BPA (20 mg/L) (Figures S2 and S3). The unique 2D vertical on 1D structure endows Bi2O2CO3–BiOI photocatalyst with distinctive photocatalytic activity. Firstly, 2D BiOI nanosheets vertically grew on the 1D Bi2O2CO3 nanorods which can provide almost exposure entire active sites; Secondly, 1D Bi2O2CO3 structures provide quickly charge carriers transfer path along their axis; In addition, the high-quality interface between Bi2O2CO3 and BiOI promotes the transfer rate of photo-generated charge carriers at junction interface, enhancing photocatalytic activity.
Free radical capture experiment is carried out to to explore active species in photocatalytical process. tert-butanol (TBA), 1, 4-benzoquinone (BQ) and ammonium oxalate (AO) were used to trap hydroxyl radical (OH·), superoxide radical (O2) and hole (H+) for Cr(VI) (30 mg/L) degradation. As can be seen from Figure 7a, the addition of AO and BQ significantly inhibited the photocatalytic reduction of Cr(VI) (30 mg/L). However, with the addition of TBA, the photoreduction efficiency of Cr(VI) (30 mg/L) only partly changes. These results prove that the main active species are·O2 and H+ during photocatalytic process.
In order to evaluate the reusability and photostability of Bi2O2CO3–BiOI heterostructure, catalytic cycle experiment was done using S2 as photocatalyst to degrade the Cr(VI) (30 mg/L) (Figure 7b). We can see that the S2 sample had good reusability, and its photocatalytic efficiency almost remained stable after five cycles. In addition, the S2 sample after 5 cycles was characterized by using XRD and SEM, and the results are shown in Figure 7c,d, respectively. The results demonstrate S2 sample retains the original structure and morphology after five cycles, implying a good photostability of S2 sample under solar light irradiation.
The room temperature PL emission spectra of the pure Bi2O2CO3, S2 and S4 (pure BiOI) are shown in Figure 8a. The PL emission intensity of S2 is the lowest one among the three samples, which implies 1D/2D heterostructure effectively suppresses the recombination of photogenerated e–h+, and thus enhancing the photocatalytic performance [51].
The photocurrent of pure Bi2O2CO3, S2 and S4 (pure BiOI) samples are shown in Figure 8b. The photocurrent density generated by the S2 sample is obviously higher than that of pure Bi2O2CO3 and BiOI. Therefore, the PL and photocurrent measurements all demonstrate that the 1D/2D Bi2O2CO3–BiOI heterostructure can significantly promote the separation and transfer of photogenerated electron–hole pairs.
In order to obtain the relative positions of the conduction band (CB) and valence band (VB) edges, VB-XPS of Bi2O2CO3 and S4 (pure BiOI) were characterized (Figure 9). The EVB top of Bi2O2CO3 and S4 locate at 1.78 and 1.15 eV, respectively. On the basis of the VB position and their band gaps, the CB edge potentials of Bi2O2CO3 and BiOI are estimated to be −1.18 and −0.6 eV, respectively through the equation ECB = EVB.
Schematic diagram for energy band of Bi2O2CO3–BiOI, the formation of the p–n junction and the possible charge separation is displayed in Figure 10. We know that p-n junctions can be formed between (Figure 10b), The internal electric field of Bi2O2CO3–BiOI p-n heterojunction promotes the migration rate of photogenerated electrons and holes, which greatly improves the photocatalytic activity. The unique vertical 2D materials on 1D structure makes BiOI nanosheets expose almost entire surface, increasing the separation and transfer rate of photo-generated electron–holes pairs. Furthermore, high-quality interface between Bi2O2CO3 and BiOI decreases the energy barrier for the photogenerated charge carriers transfer at the junction, enhancing the photocatalytic activity.

3. Experimental

3.1. Photocatalyst Preparation

Bismuth nitrate pentahydrate [(Bi(NO3)3·5H2O], Potassium iodide (KI), sodium sulfate (Na2SO4), potassium dichromate (K2Cr2O7), benzoquinone (BQ) and Butyl rhodamine B (RhB) were obtained from Sinopharm Chemical Reagent Co., Ltd. Methyl orange (MO) comes from Tianjin Guangfu Fine Chemical Research Institute. Phenol, Bisphenol A (BPA) were purchased from Shanghai Lingfeng Chemical Reagent Co., Ltd. Ammonium oxalate was obtained from Shanghai Sansihewei Chemical Co., Ltd. Tert-butyl alcohol was purchased from Sinopharm Shanghai Chemical Reagent Company. Ethylene glycol was purchased from Tianjin Fengchuan Chemical Reagent Technology Co., Ltd.
Bi2O2CO3 nanorods: Bi2O2CO3 nanorods were synthesized according to our early report [52].
Bi2O2CO3–BiOI heterostructures: In a typical synthesis, the 0.2 mmol Bi2O2CO3 nanorods and Bi(NO3)3·5H2O were added into 15 mL ethylene glycol. Then, 15 mL KI solution was slowly added into above solution under stirring at 30 °C for 3 h. Deionized water and anhydrous ethanol were used to wash the obtained products and then dry them at 60 °C for 6 h. The as-make products were named as S1, S2, S3 and S4 when the molar ratios of Bi2O2CO3 and BiOI are 10:1, 2:1, 1:2 and 1:20, respectively.

3.2. Photocatalytic Activity Measurements

The photocatalytic performance of Bi2O2CO3–BiOI heterojunction was evaluated by degrading Cr (VI), methyl orange (MO) and bisphenol A (BPA) under solar light irradiation. The 300W Xe lamp (CER-HXF300F, Beijing China Education Au-Light Co. Ltd. Bei Jing, China) was used as light source. 30 mg photocatalyst was dispersed in 30 mL K2Cr2O7 solution (30 mg/L), MO solution (20 mg/L) and BPA solution (20 mg/L), respectively. The suspension kept in dark place while it was stirred. Asan adsorption/desorption equilibrium was achieved, the suspension was illuminated under the solar light. Within a given time, 4 mL was collected and centrifugated. Finally, the supernatant is monitored by UV-vis spectrophotometer.
The trapping experiments of active species were done. During photocatalytic process, P-Benzoquinone (BQ) (0.001 mol/L), t-butanol (0.01 mol·L−1) and ammonium oxalate (AO) (0.01 mol·L−1) were added into reaction system, which can act as the scavengers to trap superoxide radicals (O2), hydroxyl radicals (OH) and hole (h+).

3.3. Electrochemical Impedance Spectroscopy (EIS) Measurements

Electrochemical impedance spectroscopy (EIS) measurements were tested at a frequency between 0.1 Hz and 100 kHz using the CHI760E instrument (Shanghai Chen Hua Company, Shanghai, China) Na2SO4 (0.2 M) was used as detecting electrolyte. The electrode system used a three-electrode system, which platinum wire was used as the counter electrode and the saturated ampho-mercury electrode is used as a reference electrode. Bi2O2CO3, S2 sample and BiOI film electrodes served as the working electrodes, separately.

3.4. Characterization

X-ray powder diffractometer (XRD) (Shimazu XRD-6000, Kyoto, Japan), scanning Angle of 10–80°. Field emission scanning electron microscope (SEM) (Hitachi S-4800, Tokyo, Japan, operating voltage: 5 Kv); Transmission electron microscope (Hitachi HT7700, Tokyo, Japan); Fluorescence spectrometer (FL, F-4500, Shimadzu, Kyoto, Japan); Infrared spectrometer (IR) (IR-21, Brock, Bochum, Germany); X-ray Photoelectron spectrometer (XPS) (Thermo EscalAB 250Xi, Waltham, MA, United States). A UV vis diffuse-reflectance spectroscopy (DRS) (UV-2450 spectrophotometer, Kyoto, Japan).

4. Conclusions

In conclusion, a novel 1D/2D Bi2O2CO3–BiOI p-n heterojunction was synthesized by hydrothermal method. BiOI nanosheets were uniformly and vertically grown from the interior of Bi2O2CO3 nanorods on the basis of a crystallography-oriented epitaxial mechanism, which provides a small barrier for the transport of photogenerated electron–hole pairs through the junctions because forming high-quality interfaces between Bi2O2CO3 and BiOI. The Bi2O2CO3–BiOI heterojunction photocatalyst exhibits a superior photocatalytic activity for the degradation of Cr (VI), MO and BPA. The outstanding photocatalytic performance is ascribed to the synergistic effects of unique 2D BiOI vertical on 1D Bi2O2CO3 multidimensional structure, the formation of the p–n junction and the high-quality interfaces between Bi2O2CO3 and BiOI. Trapping experiments show that h+ and O2 play key roles for photodegradation.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11111284/s1, Figure S1: Kinetic linear simulation curves of Cr (VI) (30 mg/L) photodegradation over the synthesized samples under visible light irradiation; Figure S2: Kinetic linear simulation curves of MO (20 mg/L) photodegradation over the synthesized samples under visible light irradiation; Figure S3: Kinetic linear simulation curves of BPA (20 mg/L) photodegradation over the synthesized samples under visible light irradiation.

Author Contributions

Methodology, investigation, N.Z.; data curation, writing—original draft, H.Q.; writing—review and editing, Y.P.; editing and polishing, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work is supported by the Natural Science Foundation of Anhui Province (2008085MB33).

Conflicts of Interest

There are no conflict to declare.

References

  1. Feng, X.; Wang, P.; Hou, J.; Qian, J.; Wang, C.; Ao, Y. Oxygen vacancies and phosphorus codoped black titania coated carbon nanotube composite photocatalyst with efficient photocatalytic performance for the degradation of acetaminophen under visible light irradiation. Chem. Eng. 2018, 352, 947–956. [Google Scholar] [CrossRef]
  2. Lashgari, M.; Zeinalkhani, P. Ammonia photosynthesis under ambient conditions using an efficient nanostructured FeS2/CNT solar-energy-material with water feedstock and nitrogen gas. Nano Energy 2018, 48, 361–368. [Google Scholar] [CrossRef]
  3. Xu, H.-M.; Wang, H.-C.; Shen, Y.; Lin, Y.-H.; Nan, C.-W. Low-dimensional nanostructured photocatalysts. J. Adv. Ceram. 2015, 4, 159–182. [Google Scholar] [CrossRef] [Green Version]
  4. Azami, M.; Haghighi, M.; Allahyari, S. Sono-precipitation of Ag2CrO4-C composite enhanced by carbon-based materials (AC, GO, CNT and C3N4) and its activity in photocatalytic degradation of acid orange 7 in water. Ultrason. Sonochem 2018, 40, 505–516. [Google Scholar] [PubMed]
  5. Chaudhary, D.; Vankar, V.D.; Khare, N. Noble metal-free g-C3N4/TiO2/CNT ternary nanocomposite with enhanced photocatalytic performance under visible-light irradiation via multi-step charge transfer process. Sol. Energy 2017, 158, 132–139. [Google Scholar] [CrossRef]
  6. He, Z.; Byun, J.-H.; Zhou, G.; Park, B.-J.; Kim, T.-H.; Lee, S.-B.; Yi, J.-W.; Um, M.-K.; Chou, T.-W. Effect of MWCNT content on the mechanical and strain-sensing performance of Thermoplastic Polyurethane composite fibers. Carbon 2019, 146, 701–708. [Google Scholar] [CrossRef]
  7. Zhong, Y.; Peng, C.; He, Z.; Chen, D.; Jia, H.; Zhang, J.; Ding, H.; Wu, X. Interface engineering of heterojunction photocatalysts based on 1D nanomaterials. Catal. Sci. Technol. 2021, 11, 27–42. [Google Scholar] [CrossRef]
  8. Chen, X.; Zhang, J.; Zeng, J.H.; Shi, Y.X.; Lin, S.Y.; Huang, G.Z.; Wang, H.B.; Kong, Z.; Xi, J.H.; Ji, Z.G. MnS coupled with ultrathin MoS2 nanolayers as heterojunction photocatalyst for high photocatalytic and photoelectrochemical activities. J. Alloy. Compd. 2019, 771, 364–372. [Google Scholar] [CrossRef]
  9. Fan, X.B.; Zhang, G.L.; Zhang, F.B. Multiple roles of graphene in heterogeneous catalysis. Chem. Soc. Rev. 2015, 44, 3023–3035. [Google Scholar] [CrossRef] [PubMed]
  10. Aradi, E.; Erasmus, R.M.; Derry, T.E. Formation of c-BN nanoparticles by helium, lithium and boron ion implantation. Nucl. Instrum. Meth. B 2012, 272, 57–60. [Google Scholar] [CrossRef]
  11. Cao, X.; Jiang, C.; Tan, D.; Li, Q.; Bi, S.; Song, J. Recent mechanical processing techniques of two-dimensional layered materials: A review. J. Sci. Adv. Mater. Devices 2021, 6, 135–152. [Google Scholar] [CrossRef]
  12. Hou, H.; Zhang, X. Rational design of 1D/2D heterostructured photocatalyst for energy and environmental applications. Chem. Eng. J. 2020, 395, 125030. [Google Scholar] [CrossRef]
  13. Dubale, A.A.; Su, W.-N.; Tamirat, A.G.; Pan, C.-J.; Aragaw, B.A.; Chen, H.-M.; Chen, C.-H.; Hwang, B.-J. The synergetic effect of graphene on Cu2O nanowire arrays as a highly efficient hydrogen evolution photocathode in water splitting. J. Mater. Chem. A 2014, 2, 18383–18397. [Google Scholar] [CrossRef]
  14. Feng, W.; Wang, Y.; Huang, X.; Wang, K.; Gao, F.; Zhao, Y.; Wang, B.; Zhang, L.; Liu, P. One-pot construction of 1D/2D Zn1-Cd S/D-ZnS(en)0.5 composites with perfect heterojunctions and their superior visible-light-driven photocatalytic H2 evolution. Appl. Catal. B Environ. 2018, 220, 324–336. [Google Scholar] [CrossRef]
  15. Zhang, G.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. Preparation of ZnIn2S4 nanosheet-coated CdS nanorod heterostructures for efficient photocatalytic reduction of Cr(VI). Appl. Catal. B Environ. 2018, 232, 164–174. [Google Scholar] [CrossRef]
  16. He, Z.; Zhang, J.; Li, X.; Guan, S.; Dai, M.; Wang, S. 1D/2D Heterostructured Photocatalysts: From Design and Unique Properties to Their Environmental Applications. Small 2020, 16, e2005051. [Google Scholar] [CrossRef]
  17. Rahman, M.Z.; Kwong, C.W.; Davey, K.; Qiao, S.Z. 2D phosphorene as a water splitting photocatalyst: Fundamentals to applications. Energy Environ. Sci. 2016, 9, 709–728. [Google Scholar] [CrossRef]
  18. Li, Y.; Huang, L.; Li, B.; Wang, X.; Zhou, Z.; Li, J.; Wei, Z. Co-nucleus 1D/2D Heterostructures with Bi2S3 Nanowire and MoS2 Monolayer: One-Step Growth and Defect-Induced Formation Mechanism. ACS Nano 2016, 10, 8938–8946. [Google Scholar] [CrossRef] [PubMed]
  19. Peng, Y.; Zhang, Q.; Kan, P.-F. Synthesis of a novel one-dimensional Bi2O2CO3–BiOCl heterostructure and its enhanced photocatalytic activity. Crystengcomm 2020, 22, 6822–6830. [Google Scholar] [CrossRef]
  20. Peng, Y.; Yan, M.; Chen, Q.-G.; Fan, C.-M.; Zhou, H.-Y.; Xu, A.-W. Novel one-dimensional Bi2O3–Bi2WO6 p–n hierarchical heterojunction with enhanced photocatalytic activity. J. Mater. Chem. A 2014, 2, 8517–8524. [Google Scholar] [CrossRef]
  21. Huang, H.; Ma, C.; Zhu, Z.; Yao, X.; Liu, Y.; Liu, Z.; Li, C.; Yan, Y. Insights into enhanced visible light photocatalytic activity of t-Se nanorods/BiOCl ultrathin nanosheets 1D/2D heterojunctions. Chem. Eng. J. 2018, 338, 218–229. [Google Scholar] [CrossRef]
  22. Vattikuti, S.V.P.; Shim, J.; Byon, C. 1D Bi2S3 nanorod/2D e-WS2 nanosheet heterojunction photocatalyst for enhanced photocatalytic activity. J. Solid State Chem. 2018, 258, 526–535. [Google Scholar] [CrossRef]
  23. Guo, Y.; Zhang, G.; Gan, H.; Zhang, Y. Micro/nano-structured CaWO4/Bi2WO6 composite: Synthesis, characterization and photocatalytic properties for degradation of organic contaminants. Dalton Trans. 2012, 41, 12697–12703. [Google Scholar] [CrossRef] [PubMed]
  24. Peng, Y.; Wang, K.K.; Liu, T.; Xu, J.; Xu, B.G. Synthesis of one-dimensional Bi2O3-Bi2O2.33 heterojunctions with high interface quality for enhanced visible light photocatalysis in degradation of high-concentration phenol and MO dyes. Appl. Catal. B Environ. 2017, 203, 946–954. [Google Scholar] [CrossRef]
  25. Peng, Y.; Wang, K.K.; Yu, P.-P.; Liu, T.; Xu, A.W. Synthesis of one-dimensional Bi2O2CO3–Bi(OHC2O42H2O heterojunctions with excellent adsorptive and photocatalytic performance. RSC Adv. 2016, 6, 42452–42460. [Google Scholar] [CrossRef]
  26. Huang, Y.J.; Zheng, Y.Q.; Wang, J.J.; Zhou, L.X. A new bismuth-based coordination polymer as an efficient visible light responding photocatalyst under white LED irradiation. J. Solid State Chem. 2017, 246, 42–47. [Google Scholar] [CrossRef]
  27. Xiong, J.; Song, P.; Di, J.; Li, H.M.; Liu, Z. Freestanding ultrathin bismuth-based materials for diversified photocatalytic applications. J. Mater. Chem. A 2019, 7, 25203–25226. [Google Scholar] [CrossRef]
  28. Shang, J.; Chen, T.Z.; Huang, G.; Zhou, F.; Wang, X.W.; Sun, L.Y. Oxygen vacancy induced bismuth basic nitrate with excellent photocatalytic activity. J. Mater. Sci. Mater. El. 2018, 29, 18067–18073. [Google Scholar] [CrossRef]
  29. Gao, P.; Yang, Y.; Yin, Z.; Kang, F.; Fan, W.; Sheng, J.; Feng, L.; Liu, Y.; Du, Z.; Zhang, L. A critical review on bismuth oxyhalide based photocatalysis for pharmaceutical active compounds degradation: Modifications, reactive sites, and challenges. J. Hazard. Mater. 2021, 412, 125186. [Google Scholar] [CrossRef]
  30. Peng, S.; Li, L.; Tan, H.; Wu, Y.; Cai, R.; Yu, H.; Huang, X.; Zhu, P.; Ramakrishna, S.; Srinivasan, M.; et al. Monodispersed Ag nanoparticles loaded on the PVP-assisted synthetic Bi2O2CO3 microspheres with enhanced photocatalytic and supercapacitive performances. J. Mater. Chem. A 2013, 1, 7630–7638. [Google Scholar] [CrossRef]
  31. Ni, Z.; Sun, Y.; Zhang, Y.; Dong, F. Fabrication, modification and application of (BiO)2CO3-based photocatalysts: A review. Appl. Surf. Sci. 2016, 365, 314–335. [Google Scholar] [CrossRef]
  32. Meng, X.; Zhang, Z. Bismuth-based photocatalytic semiconductors: Introduction, challenges and possible approaches. J. Mol. Catal. A Chem. 2016, 423, 533–549. [Google Scholar] [CrossRef]
  33. Arumugam, M.; Choi, M.Y. Recent progress on bismuth oxyiodide (BiOI) photocatalyst for environmental remediation. J. Ind. Eng. Chem. 2020, 81, 237–268. [Google Scholar] [CrossRef]
  34. Wang, X.; Zhou, C.; Yin, L.; Zhang, R.; Liu, G. Iodine-Deficient BiOI Nanosheets with Lowered Valence Band Maximum To Enable Visible Light Photocatalytic Activity. ACS Sustain. Chem. Eng. 2019, 7, 7900–7907. [Google Scholar] [CrossRef]
  35. Contreras, D.; Melin, V.; Márquez, K.; Pérez-González, G.; Mansilla, H.D.; Pecchi, G.; Henríquez, A. Selective oxidation of cyclohexane to cyclohexanol by BiOI under visible light: Role of the ratio (1 1 0)/(0 0 1) facet. Appl. Catal. B Environ. 2019, 251, 17–24. [Google Scholar] [CrossRef]
  36. Wang, S.; Wang, L.; Huang, W. Bismuth-based photocatalysts for solar energy conversion. J. Mater. Chem. A 2020, 8, 24307–24352. [Google Scholar] [CrossRef]
  37. Wang, W.; Huang, F.; Lin, X.; Yang, J. Visible-light-responsive photocatalysts xBiOBr–(1−x)BiOI. Catal. Commun. 2008, 9, 8–12. [Google Scholar] [CrossRef]
  38. Wang, Y.; Long, Y.; Zhang, D. Facile in Situ Growth of High Strong BiOI Network Films on Metal Wire Meshes with Photocatalytic Activity. ACS Sustain. Chem. Eng. 2017, 5, 2454–2462. [Google Scholar] [CrossRef]
  39. Chen, L.; Yin, S.-F.; Luo, S.-L.; Huang, R.; Zhang, Q.; Hong, T.; Au, P.C.T. Bi2O2CO3/BiOI Photocatalysts with Heterojunctions Highly Efficient for Visible-Light Treatment of Dye-Containing Wastewater. Ind. Eng. Chem. Res. 2012, 51, 6760–6768. [Google Scholar] [CrossRef]
  40. Song, P.-Y.; Xu, M.; Zhang, W.-D. Sodium citrate-assisted anion exchange strategy for construction of Bi2O2CO3/BiOI photocatalysts. Mater. Res. Bull. 2015, 62, 88–95. [Google Scholar] [CrossRef]
  41. Liu, C.; Chai, B. Facile ion-exchange synthesis of BiOI/Bi2O2CO3 heterostructure for efficient photocatalytic activity under visible light irradiation. J. Mater. Sci. Mater. Electron. 2015, 26, 2296–2304. [Google Scholar] [CrossRef]
  42. Peng, Y.; Yu, P.-P.; Zhou, H.-Y.; Xu, A.-W. Synthesis of BiOI/Bi4O5I2/Bi2O2CO3 p-n-p heterojunctions with superior photocatalytic activities. New J. Chem. 2015, 39, 8321–8328. [Google Scholar] [CrossRef]
  43. Liang, L.; Cao, J.; Lin, H.; Chen, S. Surface Na2CO3 etching induced activity enhancement of 2D BiOI photocatalyst working under visible light. Sci. Bull. 2017, 62, 546–553. [Google Scholar] [CrossRef] [Green Version]
  44. Liang, L.; Cao, J.; Lin, H.; Guo, X.; Zhang, M.; Chen, S. Synergetic effects of I− ions and BiOI on visible-light-activity enhancement of wide-band-gap (BiO)2CO3. Appl. Surf. Sci. 2017, 414, 365–372. [Google Scholar] [CrossRef]
  45. Liu, C.; Zhang, D. Bi5O7I Nanobelts: Synthesis, Modification, and Photocatalytic Antifouling Activity. Chem. Eur. J. 2019, 25, 16157–16165. [Google Scholar] [CrossRef] [PubMed]
  46. Zhou, T.; Zhang, H.; Zhang, X.; Yang, W.; Cao, Y.; Yang, P. BiOI/Bi2O2CO3 Two-Dimensional Heteronanostructures with Boosting Charge Carrier Separation Behavior and Enhanced Visible-Light Photocatalytic Performance. J. Phys. Chem. C 2020, 124, 20294–20308. [Google Scholar] [CrossRef]
  47. Zhao, Q.; Lu, L.; Wang, B.; Jiang, T. An efficient electrostatic self-assembly of reduced graphene oxide-BiOI/Bi2O2CO3 p-n junction nanocomposites for enhanced visible-light photocatalytic activity. React. Kinet. Mech. Catal. 2021, 132, 581–597. [Google Scholar] [CrossRef]
  48. Zhong, S.; Zhou, H.; Shen, M.; Yao, Y.; Gao, Q. Rationally designed a g-C3N4/BiOI/Bi2O2CO3 composite with promoted photocatalytic activity. J. Alloy. Compd. 2021, 853, 157307. [Google Scholar] [CrossRef]
  49. Butler, M.A. Photoelectrolysis and physical properties of the semiconducting electrode WO2. J. Appl. Phys. 1977, 48, 1914–1920. [Google Scholar] [CrossRef]
  50. Peng, Y.; Liu, T.; Xu, J.; Wang, K.K.; Mao, Y.G. Facet-selective interface design of a BiOI(110)/Br-Bi2O2CO3(110) p–n heterojunction photocatalyst. Cryst. Eng. Comm. 2017, 19, 6837–6844. [Google Scholar] [CrossRef]
  51. Zhuang, J.D.; Dai, W.X.; Tian, Q.F.; Li, Z.H.; Xie, L.Y.; Wang, J.X.; Liu, P.; Shi, X.C.; Wang, D.H. Photocatalytic Degradation of RhB over TiO2 Bilayer Films: Effect of Defects and Their Location. Langmuir 2010, 26, 9686–9694. [Google Scholar] [CrossRef] [PubMed]
  52. Peng, Y.; Liu, M.-Q.; Zhao, N.-N.; Kan, P.-F. Controlled synthesis of Bi2O2CO3 nanorods with enhanced photocatalytic performance. CrystEngComm 2021, 23, 3671–3680. [Google Scholar] [CrossRef]
Figure 1. FE-SEM images of (a) pure Bi2O2CO3 nanorods, Bi2O2CO3–BiOI heterojunctions of (b) S1, (c) S2, (d) S3, (e) pure BiOI. (f) XRD patterns of Bi2O2CO3–BiOI heterostructure (S1–S4) and Bi2O2CO3.
Figure 1. FE-SEM images of (a) pure Bi2O2CO3 nanorods, Bi2O2CO3–BiOI heterojunctions of (b) S1, (c) S2, (d) S3, (e) pure BiOI. (f) XRD patterns of Bi2O2CO3–BiOI heterostructure (S1–S4) and Bi2O2CO3.
Catalysts 11 01284 g001
Figure 2. (a) TEM and (b) HRTEM images of the sample S2.
Figure 2. (a) TEM and (b) HRTEM images of the sample S2.
Catalysts 11 01284 g002
Figure 3. Survey (a) and high-resolution Bi 4f (b), I 3d(c), and C 1s (d) XPS spectra of Bi2O2CO3, S2 and S4 samples.
Figure 3. Survey (a) and high-resolution Bi 4f (b), I 3d(c), and C 1s (d) XPS spectra of Bi2O2CO3, S2 and S4 samples.
Catalysts 11 01284 g003
Figure 4. FT- IR spectra of Bi2O2CO3, S2 and S4 (pure BiOI) samples.
Figure 4. FT- IR spectra of Bi2O2CO3, S2 and S4 (pure BiOI) samples.
Catalysts 11 01284 g004
Figure 5. (a) UV-Vis diffuse reflectance spectra of Bi2O2CO3, S2 heterostructures and S4 (pure BiOI) and (b) the plots of (αhν)2/n vs. hν (n = 4 for Bi2O2CO3 and BiOI).
Figure 5. (a) UV-Vis diffuse reflectance spectra of Bi2O2CO3, S2 heterostructures and S4 (pure BiOI) and (b) the plots of (αhν)2/n vs. hν (n = 4 for Bi2O2CO3 and BiOI).
Catalysts 11 01284 g005
Figure 6. The photocatalytic degradation curves of (a) Cr(VI), (c) MO and (e) BPA by different photocatalysts. UV-vis absorption spectrum of (b) Cr(VI), (d) MO and (f) BPA using S2 as a photocatalyst under solar-light irradiation.
Figure 6. The photocatalytic degradation curves of (a) Cr(VI), (c) MO and (e) BPA by different photocatalysts. UV-vis absorption spectrum of (b) Cr(VI), (d) MO and (f) BPA using S2 as a photocatalyst under solar-light irradiation.
Catalysts 11 01284 g006aCatalysts 11 01284 g006b
Figure 7. (a) The photo-degradation curves of Cr (VI) (30 mg/L) over S2 in the presence of different scavengers, (b) cycling times of the photocatalytic degradation of Cr(VI) (30 mg/L) under solar light irradiation, (c) the XRD pattern and (d) the SEM pattern of S2 after five repeated cycles.
Figure 7. (a) The photo-degradation curves of Cr (VI) (30 mg/L) over S2 in the presence of different scavengers, (b) cycling times of the photocatalytic degradation of Cr(VI) (30 mg/L) under solar light irradiation, (c) the XRD pattern and (d) the SEM pattern of S2 after five repeated cycles.
Catalysts 11 01284 g007
Figure 8. The room temperature PL spectra (λex = 320 nm) (a) and photocurrent spectra (b) of Bi2O2CO3, S2 and S4 samples.
Figure 8. The room temperature PL spectra (λex = 320 nm) (a) and photocurrent spectra (b) of Bi2O2CO3, S2 and S4 samples.
Catalysts 11 01284 g008
Figure 9. The VB-XPS spectra of Bi2O2CO3 and S4 (pure BiOI).
Figure 9. The VB-XPS spectra of Bi2O2CO3 and S4 (pure BiOI).
Catalysts 11 01284 g009
Figure 10. Schematic diagram for (a) energy band of Bi2O2CO3 and BiOI, (b) the formation of the p-n junction and the possible charge separation.
Figure 10. Schematic diagram for (a) energy band of Bi2O2CO3 and BiOI, (b) the formation of the p-n junction and the possible charge separation.
Catalysts 11 01284 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peng, Y.; Qian, H.; Zhao, N.; Li, Y. Synthesis of a Novel 1D/2D Bi2O2CO3–BiOI Heterostructure and Its Enhanced Photocatalytic Activity. Catalysts 2021, 11, 1284. https://doi.org/10.3390/catal11111284

AMA Style

Peng Y, Qian H, Zhao N, Li Y. Synthesis of a Novel 1D/2D Bi2O2CO3–BiOI Heterostructure and Its Enhanced Photocatalytic Activity. Catalysts. 2021; 11(11):1284. https://doi.org/10.3390/catal11111284

Chicago/Turabian Style

Peng, Yin, Haozhi Qian, Nannan Zhao, and Yuan Li. 2021. "Synthesis of a Novel 1D/2D Bi2O2CO3–BiOI Heterostructure and Its Enhanced Photocatalytic Activity" Catalysts 11, no. 11: 1284. https://doi.org/10.3390/catal11111284

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