Next Article in Journal
Hybrid Fiber Influence on the Crack Permeability of Cracked Concrete Exposed to Freeze–Thaw Cycles
Previous Article in Journal
Novel Probability Density Function of Pad Asperity by Wear Effect over Time in Chemical Mechanical Planarization
Previous Article in Special Issue
Organic and Biogenic Nanocarriers as Bio-Friendly Systems for Bioactive Compounds’ Delivery: State-of-the Art and Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Low-Temperature Synthesis of Bi2S3 Hierarchical Microstructures via Co-Precipitation and Digestive Process in Aqueous Medium

by
José Alfonso Carrasco-González
1,2,
Rebeca Ortega-Amaya
3,
Esteban Díaz-Torres
1,
Manuel A. Pérez-Guzmán
4,* and
Mauricio Ortega-López
2,*
1
Sección de Electrónica del Estado Sólido, Departamento de Ingeniería Eléctrica, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional, Av. IPN No. 2508, Ciudad de México 07360, Mexico
2
Programa de Doctorado Transdisciplinario en Desarrollo Científico y Tecnológico para la Sociedad, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional, Av. IPN No. 2508, Ciudad de México 07360, Mexico
3
CICFIM-Facultad de Ciencias Físico Matemáticas, Universidad Autónoma de Nuevo León, Av. Universidad S/N, Cuidad Universitaria, San Nicolás de los Garza, Nuevo León 66451, Mexico
4
Departamento de Física, Universidad Autónoma Metropolitana-Iztapalapa, Av. San Rafael Atlixco No. 186, Ciudad de México 09340, Mexico
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(8), 1818; https://doi.org/10.3390/ma17081818
Submission received: 4 March 2024 / Revised: 4 April 2024 / Accepted: 10 April 2024 / Published: 16 April 2024
(This article belongs to the Special Issue Advanced Nanomaterials: Synthesis, Characterization and Applications)

Abstract

:
Bismuth sulfide (Bi2S3) nanostructures have gained significant attention in the fields of catalysis, optoelectronics, and biomedicine due to their unique physicochemical properties. This paper introduces a simple and cost-effective method for producing Bi2S3 microstructures at low temperatures (25 and 70 °C). These microstructures are formed by the hierarchical self-assembly of Bi2S3 nanoparticles, which are typically 15–40 nm in size. The nanoparticles are synthesized by the co-precipitation of thioglycolic acid, thioacetamide, and bismuth nitrate in water. The study delves into the phase composition and morphological evolution of the microstructures, concerning the chemical composition of the solution and the synthesis temperature. X-ray analysis has confirmed the formation of single-phase bismuthinite Bi2S3. The synthesis process generates primary building blocks in the form of 15–40 nm Bi2S3 nanocrystals, which then go through a hierarchical self-assembly process to produce a range of micrometer-sized structures. A scanning electron microscopy examination revealed that the primary nanoparticles self-assemble into quasi-1D worm-like nanostructures, which then self-assemble to create sponge-shaped microstructures. These structures subsequently self-organize and refine into either flower- or dandelion-like microstructures, mostly depending on the synthesis temperature and the chemistry of the digestion medium.

1. Introduction

The development of advanced materials with unique properties holds immense potential for various technological and scientific applications. Amongst these, 3D hierarchical nanostructures of binary V-VI compounds have gained significant attention due to their exceptional structural architecture and enhanced performance in energy conversion-related and environmental remediation applications [1]. Therefore, it is crucial to establish cost-effective and straightforward processes for synthesizing these materials. In this regard, colloidal chemistry in an aqueous medium can play a pivotal role in achieving this objective.
The semiconductor material bismuth sulfide (Bi2S3) has raised special interest due to its optical and electrical properties, which can be widely tailored by controlling the size and shape of its morphological features during synthesis [2,3,4,5,6]. Crystalline Bi2S3 exhibits n-type conductivity and has a direct band with a band gap energy of Eg = 1.4 eV. However, various synthesis methods for Bi2S3 nanocrystals reported Eg values ranging from 1.3 to 1.7 eV [6,7]. In comparison to other highly toxic mineral sulfides such as Greenockite (CdS), Orpiment (As2S3), or Cinnabar (HgS), Bismutinite (Bi2S3) can be considered a compound with low or no toxicity [8,9,10,11].
Bismuth sulfide exhibits a unique crystal structure, characterized by a unit cell that belongs to the orthorhombic crystalline system Pnma. The lattice parameters for Bi2S3 are a = 11.15 Å, b = 11.30 Å, and c = 3.981 Å. Within its unit cell, there are 20 atoms, which consist of four Bi2S3 units. These atoms arrange themselves in slats along the c-axis and are stacked together by intermolecular bonds along the a-axis, resulting in the formation of three-dimensional lamella-like structures [12]. These structural properties allow Bi2S3 to develop crystalline habits such as nanorods, nanosheets, and acicular structures [13,14,15,16,17,18,19,20,21].
Recently, nanostructured Bi2S3 has attracted significant attention due to its ease of preparation using various techniques. Chemical methods where a powdered material [22], thin films [23], and solutions [24] were obtained and physical methods that resulted in films and bulk materials [25,26,27] were proven. Consequently, this has allowed extensive bismuth sulfide study of its potential applications in polymer/Bi2S3 hybrid solar cells, hydrogen storage [28,29], and thin films for electronic devices [30,31,32].
The primary aim of this study is to synthesize nanostructures of Bi2S3 utilizing a cost-effective and straightforward process, such as co-precipitation in an aqueous medium.

2. Experimental Procedures

2.1. Reagents and Material Synthesis

The Bi2S3 nanostructures were prepared by co-precipitation using aqueous solutions of bismuth nitrate pentahydrate (Bi(NO3)3•5H2O, ≥98% Sigma-Aldric, Toluca, Mexico), thioglycolic acid (TGA, HSCH2CO2H, ≥98% Sigma-Aldrich), and thioacetamide (TAA, H2C=CSNH2, Analytyka, Nuevo Leon, Mexico). All reagents are used without any additional purification process.
In all experiments, the precursor solutions of sulfur and bismuth ions were separately prepared as follows:
Solution A. The TAA solution was prepared at room temperature by dissolving 451 mg of TAA in 60 mL of deionized water for 3 min, with the concentration being 0.1 M in TAA.
Solution B: The bismuth nitrate solution involved the direct dissolution of the bismuth salt in TGA within a ball flask with constant magnetic stirring. After complete dissolution (~7 min), 5 mL of deionized water was added, and it was either heated up to 70 °C or maintained at 25 °C for 15 min. Subsequently, 30 mL of solution A was mixed with 30 mL of deionized water. The resulting solution was completely mixed with solution B. This mixing triggered a rapid reaction, causing the solution color to change from yellow to orange, then red, and finally a dark brown. The colloidal reaction was allowed to proceed for 60 min at the chosen synthesis temperature. The minimum reported reaction time is that measured just after the mixing process was performed, being around 3 min.
Following the synthesis process, the heating source was turned off, and the resulting solution was left to undergo a 24 h digestion period. This digestion process facilitates the separation of the solid phase, forming a sedimented floc, from the liquid phase. The liquid phase, containing dispersed Bi2S3 nanoparticles and byproducts, was carefully poured off. To ensure a neutral pH, the floc was subjected to multiple washes. Finally, the floc was effectively dried by heating it at 70 °C. Figure 1 details the entire Bi2S3 synthesis process described in this section.
Four series of experiments were prepared and labeled as BiS10X, BiS20X, and BiS30X (X = 1–3), where the TAA concentration, the TGA one, and the Bi(NO3)3•5H2O one were varied, respectively, and the BiSTAX series, where the reaction was performed at 25 °C, varying the TGA concentration. However, while some experiments produced samples with similar structural and morphological features, the most notable distinction was observed in samples prepared at varying temperatures. For the purpose of our discussion, we have selected samples prepared at 25 and 70 °C as representative examples.

2.2. Characterization Techniques

The phase composition of Bi2S3 nanostructures was assessed by X-ray diffraction (XRD), and a PANalytical X’Pert-PRO diffractometer (Malvern Panalytical, Malvern, UK) with Cu-Kα emission was used for this purpose. The scan range was set from 15° to 70° with a step size of 0.04°. The crystal size was determined by analyzing the position of the highest intensity peaks and their corresponding Full Width at Half Maximum (FWHM) using the Sherrer equation, with a value of k = 0.9 [33]. The morphological characterization of the powders was carried out using scanning electron microscopy (SEM), where a FE-SEM Zeiss Auriga (Carl Zeiss Microscopy GmbH, Jena, Germany) operating at 5 and 10 kV was utilized for this analysis.

3. Results and Discussion

The phase composition and morphology of the Bi2S3 powder were analyzed by XRD and SEM, respectively. As stated above, various experimental conditions led to samples with similar compositions and morphological developments. The following discussion is based on the representative samples synthesized at room temperature (~25 °C, BiSTAX) and 70 °C (BiS10X).

3.1. XRD Analysis

The diffractograms shown in Figure 2 correspond to the Bi2S3 samples corresponding to the BiSTAX series (Figure 2a) and BiS10X series (Figure 2b). It is worth noting that well-crystallized samples were obtained, despite being prepared at low temperature (25 °C). All the observed diffraction lines are attributed to the orthorhombic phase of Bi2S3 bismuthinite, according to the ICCD No. 00-006-0333 reference card. No diffraction peaks corresponding to solid phases other than bismuthinite Bi2S3 were detected. As previously mentioned, the Bi salt was directly mixed with the TGA to minimize Bi-ion hydrolysis and the subsequent formation of complex hydrated bismuth oxide species [29]. This is because both thioglycolic acid and thioacetamide serve a dual role as Bi-ion ligands and as a source of sulfur (S) [30,31]. This successful process of preparing the precursor solution effectively prevents the formation of impurity phases other than Bi2S3.
The lattice parameters were estimated using the XRD data from the most intense peaks, resulting around a = 11.18 Å, b = 11.43 Å, and c = 3.98 Å. These values slightly differ from those reported in the ICCD reference card. That is, a and b are larger than the reference values in about 0.27 and 1.15%, respectively, indicating a tensile stress; on the contrary, the lattice parameter c resulted in a minimum compression stress of 0.03%, which represents a practically negligible difference. The crystallite size, on the other hand, was found to vary in the 15–40 nm range. Nevertheless, as discussed below, in some cases, certain experimental parameters slightly affected the final morphology of the self-assembled microstructures.

3.2. Morphological Analysis

The morphology of the Bi2S3 powder was analyzed using SEM. Figure 3 presents SEM images of representative samples prepared at room temperature (25 °C) (Figure 3a) and 70 °C (Figure 3b). The synthesis procedure, which combines co-precipitation and digestion, resulted in Bi2S3 microstructures with a wide variety of forms, all seemingly derived from a self-assembly process. The XRD analysis confirmed that these microstructures are composed of 15–40 nm-in-size Bi2S3 nanoparticles, whereas the SEM images revealed larger Bi2S3 structures in the micrometer range. Therefore, the SEM characterization suggests that the synthesis procedure produces nanosized Bi2S3 as the primary building blocks, from which microstructures develop through sequential self-assembly steps. The dominant morphologies that emerged at different temperatures are highlighted in Figure 3. At room temperature, Figure 3a,c show the formation of sponge-, urchin-, and dandelion-like bismuth sulfide microstructures, which were influenced by the TGA content. On the other hand, at 70 °C (Figure 3b,d), sponge-, urchin, flower- and coral-like microstructures are predominantly observed under all the tested experimental conditions. Detailed close-up views of each morphology can be seen in Figure 3c,d. The sponge-like and dandelion structures have average sizes of 1.16 µm and 0.90 µm, respectively, while the flower-like structure is 1.30 µm in size, and the maximal visible cross-sectional length of their leaves is 122.73 nm.
To gain a better understanding of the sequential self-assembly process, aliquots from representative precipitation reactions were taken out ~1 min after starting the precipitation reaction. Figure 4a illustrates the morphological evolution of Bi2S3 microstructures after ~1 min of initiating phase solid precipitation at 70 °C. It is evident that upon the onset of precipitation, nanoscale Bi2S3 particles undergo self-assembly to form worm-like nanostructures, which then further assemble into intricate sponge-like microstructures, indicating that the hierarchical self-assembly of Bi2S3 nanoparticles begins accompanying the solid phase precipitation and continues to progress during the digestive process, as shown in Figure 3b. Figure 4a reveals that worm-like structures are commonly observed in samples synthesized at both 25 °C and 70 °C and that they self-assemble into a woven network sponge-like microstructure, as shown in Figure 4b.
We have proposed a potential pathway that leads to the formation of microstructures by closely examining the SEM images shown in Figure 3a,b. We have assumed that these figures serve to illustrate the critical steps of the self-assembly process. That is, we propose that the SEM images include micron-sized Bi2S3 with different degrees of development. The different stages of self-assembly and formation are highlighted in Figure 3a,b, denoted by numbers, illustrating (1) the early stage of worm formation, (2) the sponge-like formation stage, and (3) the final stage of acicular microstructure formation. Based on Figure 3a, we propose that the morphology progresses sequentially as follows:
Initially, primary Bi2S3 nanoparticles self-assemble into worm-like 1D nanostructures, which subsequently group and coalesce into waved worm-like structures resembling sponge-like microstructures. The subsequent shape modification of self-assembled Bi2S3 microstructures develops during the digestion process. Figure 4a illustrates the morphological details of the Bi2S3 microstructures. It is evident that at the early stage of the self-assembly process, worm- and sponge-like microstructures emerged as the dominant ones. These develop into porous Bi2S3 microstructures resembling a sponge-like structure, which are prone to further shape transformations during the digestion process, depending on the temperature. The worm-like and sponge-shaped structures appear to be a common occurrence across all the samples, regardless of the experimental conditions used for their preparation. This assertion is illustrated by Figure 4b for synthesis at room temperature.
Following this starting formation, the sponge-shaped microstructures are subjected to Ostwald ripening, redissolution, and reshaping during digestion to achieve their final shape. At 25 °C, the mechanisms responsible for consolidating and refining the crystalline microstructure typically result in mostly porous, dandelion-like Bi2S3 structures, as seen in Figure 3a,c. Conversely, the samples prepared at 70 °C exhibit a transformation from nanoworms into acicular structures, which then self-assemble to form flower-like microstructures, as shown in Figure 3b,d.
Figure 5 provides a summary of the self-assembly and digestion processes described above.
To compare our results with other related works, it is worth mentioning that the synthesis of bismuth sulfide microstructures has extensively been studied, and Bi2S3 microstructures like those reported here have been observed using both chemical and physical synthesis. For instance, urchin-like Bi2S3 nanocrystals were obtained by Ma, Li, Chen, Sasikala, and Sang [21,22,34,35,36] by chemical techniques, and Song, Ten Haaf, and Li [25,37,38] obtained needle-, block-, bar-, and rod-shaped ones by vacuum techniques. Salavati-Niasari and Zhang et al. [20,39] synthesized urchin- and dandelion-like microstructures by using aqueous chemical methods containing TGA and/or thioacetamide. These researchers conducted a comprehensive analysis to propose that the TGA plays a crucial role in the development of such Bi2S3 microstructures.
The observations above indicate that the morphological development of nanostructured Bi2S3 appears to be unaffected by the synthesis method and is primarily determined by the crystal structure of Bi2S3.
Our results significantly differ from others previously reported. In our study, nanosized Bi2S3 particles self-assembled into worm-like nanostructures almost at the onset of precipitation. Subsequently, the self-assembly steps of these 1D nanostructures produced sponge-, urchin-, dandelion-, or flower-like Bi2S3 microstructures. The dandelion formation was favored when the synthesis was carried out at room temperature, regardless of the TGA content. Whereas, for the synthesis at 70 °C, all the prepared samples exhibited a similar morphology as that developed by the BiS10X series, resembling a flower-like structure. On the basis of these observations, the hierarchical self-assembly pathway leading to the formation of bismuth sulfide microstructures is primarily dictated by the crystalline structure of Bi2S3 and by the chemistry of the solution in which the digestion process takes place. Notice that digestion was carried out at room temperature.
In our experiments, the role of TGA as an assembly director agent was not clear enough. In this sense, our interpretation differs from that of Salavati et al. [16], who proposed that TGA directs the final microstructure shape of Bi2S3. Instead, our results suggest that the crystal habit of Bi2S3, the synthesis temperature, and the solution digestion are the key factors influencing the microstructure formation. By understanding these factors, we can gain valuable insights into the synthesis of bismuth sulfide microstructures and potentially control their shape and properties.
Overall, this research demonstrates the fascinating hierarchical self-assembly process that leads to the formation of a great variety of Bi2S3 microstructures using simple and cheap synthesis techniques. The study also highlights the impact of digestion on the refinement of these structures, providing valuable insights into their formation and development. The SEM analysis has provided valuable information related to the potential applications of Bi2S3 in various fields such as optoelectronics, catalysis, and energy storage. The thorough investigation of the morphology of Bi2S3 powders by SEM has enhanced our understanding of their physical properties and paved the way for their utilization in advanced device research.

4. Conclusions

In conclusion, this study introduces a simple and cost-effective method for producing Bi2S3 microstructures in a variety of sizes and shapes. The synthesis process involves co-precipitation at low temperatures (25 and 70 °C), followed by a 24 h digestion period. This method results in a highly pure Bi2S3 powder. The unique morphology of the Bi2S3 microstructures is achieved through a hierarchical self-assembly process and a digestion mechanism.
Starting with Bi2S3 particles ranging from 15 to 40 nm in size, it was observed that these nanoparticles self-assemble into worm-like 1D nanostructures during the precipitation reaction. These nanoworms then further assemble into sponge-, urchin-, and dandelion or flower-like Bi2S3 microstructures. What sets our results apart from previous studies is the role of worm-like nanostructures as building blocks for more complex forms. These structures undergo shape transformations due to Ostwald ripening, nanocrystal adhesion, and reshaping during the digestion process.
The final shape of the Bi2S3 microstructures, particularly the crystal habit of Bi2S3, is influenced by the chemical composition of the solution and temperature. Dandelion-like and flower-like microstructures were observed at 25 and 70 °C, respectively. The resulting Bi2S3 microstructures morphology have a high surface area, making them ideal for applications in sensing and catalysis.

Author Contributions

J.A.C.-G. participated in the investigation, methodology, conceptualization, and writing original draft of this manuscript, R.O.-A. and E.D.-T. participated in the methodology, validation, writing original draft, review, and editing, M.A.P.-G. and M.O.-L. participated in the conceptualization, supervision, writing original draft, review, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors wish to thank Álvaro Guzmán Campuzano for assistance in the material synthesis, Adolfo Tavira Fuentes for the XRD measurements, and LANE CINVESTAV-IPN for the SEM images, and CONAHCYT-México for the postdoctoral fellowships provided to R.O.-A., E.D.-T. and M.A.P.-G.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Boles, M.A.; Engel, M.; Talapin, D.V. Self-assembly of colloidal nanocrystals: From intricate structures to functional materials. Chem. Rev. 2016, 116, 11220–11289. [Google Scholar] [CrossRef] [PubMed]
  2. Miller, N.C.; Bernechea, M. Research Update: Bismuth based materials for photovoltaics. APL Mater. 2018, 6, 084503. [Google Scholar] [CrossRef]
  3. Aresti, M.; Saba, M.; Piras, R.; Marongiu, D.; Mula, G.; Quochi, F.; Mura, A.; Cannas, C.; Mureddu, M.; Ardu, A.; et al. Colloidal Bi2S3 nanocrystals: Quantum size effects and midgap states. Adv. Funct. Mater. 2014, 24, 3341–3350. [Google Scholar] [CrossRef]
  4. Ma, J.; Yang, J.; Jiao, L.; Wang, T.; Lian, J.; Duan, X.; Zheng, W. Bi2S3 nanomaterials: Morphology manipulation and related properties. Dalton Trans. 2011, 40, 10100–10109. [Google Scholar] [CrossRef] [PubMed]
  5. Bernechea, M.; Cao, Y.; Konstantatos, G. Size and bandgap tunability in Bi2S3 colloidal nanocrystals and its effect in solution processed solar cells. J. Mater. Chem. A 2015, 3, 20642–20648. [Google Scholar] [CrossRef]
  6. Calzia, V.; Malloci, G.; Bongiovanni, G.; Mattoni, A. Electronic Properties and Quantum Confinement in Bi2S3 Ribbon-like Nanostructures. J. Phys. Chem. C 2013, 117, 21923–21929. [Google Scholar] [CrossRef]
  7. Rincón, M.E.; Sánchez, M.; George, P.J.; Sánchez, A.; Nair, P.K. Comparison of the Properties of Bismuth Sulfide Thin Films Prepared by Thermal Evaporation and Chemical Bath Deposition. J. Solid State Chem. 1998, 136, 167–174. [Google Scholar] [CrossRef]
  8. Khan, Z.; Elahi, A.; Bukhari, D.A.; Rehman, A. Cadmium sources, toxicity, resistance and removal by microorganisms-A potential strategy for cadmium eradication. J. Saudi Chem. Soc. 2022, 26, 101569. [Google Scholar] [CrossRef]
  9. Min, X.; Xu, Q.; Ke, Y.; Xu, H.; Yao, L.; Wang, J.; Ren, H.; Li, T.; Lin, Z. Transformation behavior of the morphology, structure and toxicity of amorphous As2S3 during hydrothermal process. Hydrometallurgy 2021, 200, 105549. [Google Scholar] [CrossRef]
  10. Nádudvari, Á.; Cabała, J.; Marynowski, L.; Jabłońska, M.; Dziurowicz, M.; Malczewski, D.; Kozielska, B.; Siupka, P.; Piotrowska-Seget, Z.; Simoneit, B.R.T.; et al. High concentrations of HgS, MeHg and toxic gas emissions in thermally affected waste dumps from hard coal mining in Poland. J. Hazard Mater. 2022, 431, 128542. [Google Scholar] [CrossRef]
  11. Guo, J.; Lou, Q.; Qiu, Y.; Wang, Z.Y.; Ge, Z.H.; Feng, J.; He, J. Remarkably enhanced thermoelectric properties of Bi2S3 nanocomposites via modulation doping and grain boundary engineering. Appl. Surf. Sci. 2020, 520, 146341. [Google Scholar] [CrossRef]
  12. Ajiboye, T.O.; Onwudiwe, D.C. Bismuth sulfide based compounds: Properties, synthesis and applications. Results Chem. 2021, 3, 100151. [Google Scholar] [CrossRef]
  13. Dong, Y.; Hu, M.; Zhang, Z.; Zapien, J.A.; Wang, X.; Lee, J.M. Hierarchical self-assembled Bi2S3 hollow nanotubes coated with sulfur-doped amorphous carbon as advanced anode materials for lithium ion batteries. Nanoscale 2018, 10, 13343–13350. [Google Scholar] [CrossRef] [PubMed]
  14. Arabzadeh, A.; Salimi, A. Facile Synthesis of Ultra-wide Two Dimensional Bi2S3 Nanosheets: Characterizations, Properties and Applications in Hydrogen Peroxide Sensing and Hydrogen Storage. Electroanalysis 2017, 29, 2027–2035. [Google Scholar] [CrossRef]
  15. Sahu, M.; Park, C. A comprehensive review on bismuth-sulfide-based compounds. Mater. Today Sustain. 2023, 23, 10441. [Google Scholar] [CrossRef]
  16. Chen, Z.; Cao, M. Synthesis, characterization, and hydrophobic properties of Bi2S3 hierarchical nanostructures. Mater. Res. Bull. 2011, 46, 555–562. [Google Scholar] [CrossRef]
  17. Zhu, G.; Liu, P. Low-temperature urea-assisted hydrothermal synthesis of Bi2S3 nanostructures with different morphologies. Cryst. Res. Technol. 2009, 44, 713–720. [Google Scholar] [CrossRef]
  18. Miniach, E.; Zyna Gryglewicz, G. Solvent-controlled morphology of bismuth sulfide for supercapacitor applications. J. Mater. Sci. 2018, 53, 16511–16523. [Google Scholar] [CrossRef]
  19. Deshpande, M.P.; Sakariya, P.N.; Bhatt, S.V.; Garg, N.; Patel, K.; Chaki, S.H. Characterization of Bi2S3 nanorods prepared at room temperature. Mater. Sci. Semicond. Process. 2014, 21, 180–185. [Google Scholar] [CrossRef]
  20. Salavati-Niasari, M.; Ghanbari, D.; Davar, F. Synthesis of different morphologies of bismuth sulfide nanostructures via hydrothermal process in the presence of thioglycolic acid. J. Alloys Compd. 2009, 488, 442–447. [Google Scholar] [CrossRef]
  21. Ma, J.; Liu, Z.; Lian, J.; Duan, X.; Kim, T.; Peng, P.; Liu, X.; Chen, Q.; Yao, G.; Zheng, W. Ionic liquids-assisted synthesis and electrochemical properties of Bi2S3 nanostructures. CrystEngComm 2011, 13, 3072–3079. [Google Scholar] [CrossRef]
  22. Li, Y.; Wei, F.; Ma, Y.; Zhang, H.; Gao, Z.; Dai, L.; Qin, G. Selected-control hydrothermal synthesis and photoresponse properties of Bi2S3 micro/nanocrystals. CrystEngComm 2013, 15, 6611. [Google Scholar] [CrossRef]
  23. Carrillo-Castillo, A.; Rivas-Valles, B.G.; Castillo, S.J.; Ramirez, M.M.; Luque-Morales, P.A. New Formulation to Synthetize Semiconductor Bi2S3 Thin Films Using Chemical Bath Deposition for Optoelectronic Applications. Symmetry 2022, 14, 2487. [Google Scholar] [CrossRef]
  24. Han, P.; Mihi, A.; Ferre-borrull, J.; Pallarés, J.; Marsal, L.F. Interplay Between Morphology, Optical Properties, and Electronic Structure of Solution-Processed Bi2S3 Colloidal Nanocrystals. J. Phys. Chem. C 2015, 119, 10693–10699. [Google Scholar] [CrossRef]
  25. Song, H.; Zhan, X.; Li, D.; Zhou, Y.; Yang, B.; Zeng, K.; Zhong, J.; Miao, X.; Tang, J. Rapid thermal evaporation of Bi2S3 layer for thin film photovoltaics. Sol. Energy Mater. Sol. C 2016, 146, 1–7. [Google Scholar] [CrossRef]
  26. Wu, Y.; Lou, Q.; Qiu, Y.; Guo, J.; Mei, Z.Y.; Xu, X.; Feng, J.; He, J.; Ge, Z.H. Highly enhanced thermoelectric properties of nanostructured Bi2S3 bulk materials: Via carrier modification and multi-scale phonon scattering. Inorg. Chem. Front. 2019, 6, 1374–1381. [Google Scholar] [CrossRef]
  27. Liu, W.; Lukas, K.C.; McEnaney, K.; Lee, S.; Zhang, Q.; Opeil, C.P.; Chen, G.; Ren, Z. Studies on the Bi2Te3–Bi2Se3–Bi2S3 system for mid-temperature thermoelectric energy conversion. Energy Environ. Sci. 2013, 6, 552–560. [Google Scholar] [CrossRef]
  28. Saha, S.K.; Pal, A.J. Schottky diodes between Bi2S3 nanorods and metal nanoparticles in a polymer matrix as hybrid bulk-heterojunction solar cells. J. Appl. Phys. 2015, 118, 014503. [Google Scholar] [CrossRef]
  29. Jin, R.; Li, G.; Xu, Y.; Liu, J.; Chen, G. Uniform Bi2S3 nanorods-assembled hollow spheres with excellent electrochemical hydrogen storage abilities. Int. J. Hydrogen. Energy 2014, 39, 356–365. [Google Scholar] [CrossRef]
  30. Shkir, M. Improved opto-electronic properties of Bi2S3 thin films through trivalent atom (Al3+) doping for photodiode applications. Surf. Interfaces 2023, 40, 103025. [Google Scholar] [CrossRef]
  31. Chitara, B.; Kolli, B.S.C.; Yan, F. Near-Infrared photodetectors based on 2D Bi2S3. Chem. Phys. Lett. 2022, 804, 139876. [Google Scholar] [CrossRef]
  32. Kim, Y.; Jeong, E.; Joe, M.; Lee, C. Synthesis of 2D semiconducting single crystalline Bi2S3 for high performance electronics. Phys. Chem. Chem. Phys. 2021, 23, 26806–26812. [Google Scholar] [CrossRef] [PubMed]
  33. Uddin, I.; Abzal, S.M.; Kalyan, K.; Janga, S.; Rath, A.; Patel, R.; Gupta, D.K.; Ravindran, T.R.; Ateeq, H.; Khan, M.S.; et al. Starch-Assisted Synthesis of Bi2S3 Nanoparticles for Enhanced Dielectric and Antibacterial Applications. ACS Omega 2022, 7, 42438–42445. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, J.; Qin, S.; Song, G.; Xiang, T.; Xin, F.; Yin, X. Shape-controlled solvothermal synthesis of Bi2S3 for photocatalytic reduction of CO2 to methyl formate in methanol. Dalton Trans. 2013, 42, 15133. [Google Scholar] [CrossRef] [PubMed]
  35. Sasikala, S.; Balakrishnan, M.; Kumar, M.; Chang, J.H.; Manivannan, M.; Thangabalu, S. Effect of sulfur source and temperature on the morphological characteristics and photocatalytic activity of Bi2S3 nanostructure synthesized by microwave irradiation technique. J. Mater. Sci. Mater. Electron. 2023, 34, 997. [Google Scholar] [CrossRef]
  36. Sang, Y.; Dai, G.; Wang, L.; Gao, X.; Fang, C. Hydrothermal Synthesis of Urchin-like Bi2S3 Nanostructures for Superior Visible-light-driven Cr(VI) Removal Capacity. ChemistrySelect 2018, 3, 7123–7128. [Google Scholar] [CrossRef]
  37. ten Haaf, S.; Sträter, H.; Brüggemann, R.; Bauer, G.H.; Felser, C.; Jakob, G. Physical vapor deposition of Bi2S3 as absorber material in thin film photovoltaics. Thin Solid Films 2013, 535, 394–397. [Google Scholar] [CrossRef]
  38. Li, Z.; Zhang, Q.; Dan, M.; Guo, Z.; Zhou, Y. A facile preparation route of Bi2S3 nanorod films for photocatalytic H2 production from H2S. Mater. Lett. 2017, 201, 118–121. [Google Scholar] [CrossRef]
  39. Zhang, Z.; Zhou, C.; Lu, H.; Jia, M.; Lai, Y.; 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]
Figure 1. Schematic representation of Bi2S3 colloidal synthesis process, indicating the dissolving, mixing, reaction, digestion, and powder obtention stages.
Figure 1. Schematic representation of Bi2S3 colloidal synthesis process, indicating the dissolving, mixing, reaction, digestion, and powder obtention stages.
Materials 17 01818 g001
Figure 2. XRD patterns of the Bi2S3 samples. (a) Samples corresponding to the BiSTAX series (synthesized at 25 °C) and (b) samples corresponding to the BiS10X series (synthesized at 70 °C). Notice that the vertical lines belonging to the bismuthinite mineral reference card ICCD No. 00-006-0333 were added.
Figure 2. XRD patterns of the Bi2S3 samples. (a) Samples corresponding to the BiSTAX series (synthesized at 25 °C) and (b) samples corresponding to the BiS10X series (synthesized at 70 °C). Notice that the vertical lines belonging to the bismuthinite mineral reference card ICCD No. 00-006-0333 were added.
Materials 17 01818 g002
Figure 3. (a,b) Panoramic SEM images of the two Bi2S3 crystalline acicular structures, simultaneously displaying different self-assembly stages. (1) indicates the early stage (worm formation), (2) sponge-like formation stage, and (3) formed acicular structures (dandelions at 25 °C or flowers at 70 °C). (c,d) Close-up view of dandelion and flower Bi2S3 crystalline microstructures.
Figure 3. (a,b) Panoramic SEM images of the two Bi2S3 crystalline acicular structures, simultaneously displaying different self-assembly stages. (1) indicates the early stage (worm formation), (2) sponge-like formation stage, and (3) formed acicular structures (dandelions at 25 °C or flowers at 70 °C). (c,d) Close-up view of dandelion and flower Bi2S3 crystalline microstructures.
Materials 17 01818 g003
Figure 4. SEM images of Bi2S3 crystals in some stages of the self-assembly process. (a) Rolled-up worm formation and (b) sponge-like structure formation stage. Rectangles were added to the image (a) to indicate the unrolled (blue) and rolled-up worm (yellow) crystalline structures.
Figure 4. SEM images of Bi2S3 crystals in some stages of the self-assembly process. (a) Rolled-up worm formation and (b) sponge-like structure formation stage. Rectangles were added to the image (a) to indicate the unrolled (blue) and rolled-up worm (yellow) crystalline structures.
Materials 17 01818 g004
Figure 5. Schematic representation of the hierarchical self-assembly process of Bi2S3 crystals changing into sponge-like and acicular structures.
Figure 5. Schematic representation of the hierarchical self-assembly process of Bi2S3 crystals changing into sponge-like and acicular structures.
Materials 17 01818 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Carrasco-González, J.A.; Ortega-Amaya, R.; Díaz-Torres, E.; Pérez-Guzmán, M.A.; Ortega-López, M. Low-Temperature Synthesis of Bi2S3 Hierarchical Microstructures via Co-Precipitation and Digestive Process in Aqueous Medium. Materials 2024, 17, 1818. https://doi.org/10.3390/ma17081818

AMA Style

Carrasco-González JA, Ortega-Amaya R, Díaz-Torres E, Pérez-Guzmán MA, Ortega-López M. Low-Temperature Synthesis of Bi2S3 Hierarchical Microstructures via Co-Precipitation and Digestive Process in Aqueous Medium. Materials. 2024; 17(8):1818. https://doi.org/10.3390/ma17081818

Chicago/Turabian Style

Carrasco-González, José Alfonso, Rebeca Ortega-Amaya, Esteban Díaz-Torres, Manuel A. Pérez-Guzmán, and Mauricio Ortega-López. 2024. "Low-Temperature Synthesis of Bi2S3 Hierarchical Microstructures via Co-Precipitation and Digestive Process in Aqueous Medium" Materials 17, no. 8: 1818. https://doi.org/10.3390/ma17081818

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