Next Article in Journal
Formation of Iron (Hydr)Oxide Nanoparticles with a pH-Clock
Next Article in Special Issue
In-Depth Insight into the Effect of Hydrophilic-Hydrophobic Group Designing in Amidinium Salts for Perovskite Precursor Solution on Their Photovoltaic Performance
Previous Article in Journal
Enhanced Diabetic Wound Healing Using Electrospun Biocompatible PLGA-Based Saxagliptin Fibrous Membranes
Previous Article in Special Issue
Quantum Mechanical Analysis Based on Perturbation Theory of CdSe/ZnS Quantum-Dot Light-Emission Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimizing the Synthetic Conditions of “Green” Colloidal AgBiS2 Nanocrystals Using a Low-Cost Sulfur Source

1
School of Physics, Northwest University, Xi’an 710127, China
2
Qingdao Innovation and Development Base of Harbin Engineering University, Harbin Engineering University, Harbin 150001, China
3
Shaanxi Key Laboratory for Carbon Neutral Technology, Xi’an 710127, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(21), 3742; https://0-doi-org.brum.beds.ac.uk/10.3390/nano12213742
Submission received: 23 September 2022 / Revised: 21 October 2022 / Accepted: 21 October 2022 / Published: 25 October 2022
(This article belongs to the Special Issue Quantum Dot Materials and Optoelectronic Devices)

Abstract

:
Colloidal AgBiS2 nanocrystals (NCs) have attracted increasing attention as a near–infrared absorbent materials with non–toxic elements and a high absorption coefficient. In recent years, colloidal AgBiS2 NCs have typically been synthesized via the hot injection method using hexamethyldisilathiane (TMS) as the sulfur source. However, the cost of TMS is one of the biggest obstacles to large–scale synthesis of colloidal AgBiS2 NCs. Herein, we synthesized colloidal AgBiS2 NCs using oleylamine@sulfur (OLA–S) solution as the sulfur source instead of TMS and optimized the synthesis conditions of colloidal AgBiS2 NCs. By controlling the reaction injection temperature and the dosage of OLA–S, colloidal AgBiS2 NCs with adjustable size can be synthesized. Compared with TMS–based colloidal AgBiS2 NCs, the colloidal AgBiS2 NCs based on OLA–S has good crystallinity and fewer defects.

1. Introduction

Colloidal AgBiS2 nanocrystal (NC), a member of the I–V–VI2 ternary semiconductor materials [1,2,3,4,5,6,7], is a promising eco–friendly material for solar cells owing to colloidal AgBiS2 NC having a wide absorption spectrum in the visible to near–infrared region (300–1600 nm) [8,9,10], high absorption coefficient (~105 cm−1) [1,11,12,13], and good air stability [8]. The hot injection method is widely used to prepare high quality NCs such as PbS [14], PbSe [15,16], and halide perovskite NCs [17,18,19,20,21,22]. In 2016, Bernechea et al. synthesized colloidal AgBiS2 NCs by the hot injection method using hexamethyldisilathiane (TMS) as the sulfur source and first reported colloidal AgBiS2 NC–based solar cells with a certified power conversion efficiency (PCE) of 6.3% [1]. Subsequently, colloidal AgBiS2 NC–based solar cells have gradually become one of the mainstays of new generation solar cells. Subsequently, Hu et al. synthesized colloidal AgBiS2 NCs using an improved amine–based synthesis route, improving the PCE of colloidal AgBiS2 NC–based solar cells by 30% using colloidal AgBiS2 NCs synthesized using a modified route [23]. Oh et al. improved the size distribution of colloidal AgBiS2 NCs by an optimized synthetic route, resulting in enhanced photovoltaic performance [24]. Akgul et al. synthesized colloidal AgBiS2 NCs at room temperature using an improved sulfur amine–based synthesis route, reducing the production price by 60% and achieving a PCE of 5.5% [25]. These results demonstrate that the quality of colloidal AgBiS2 NCs plays an important role in the performance of solar cells. Therefore, it is of extreme importance to synthesize monodispersed colloidal AgBiS2 NCs in the quantum confinement regime and with a high phase purity [26]. Yet, it should be noted that although synthesis of colloidal AgBiS2 NCs has been developed by several researchers [1,27,28,29,30], it remains challenging to synthesize colloidal AgBiS2 NCs with a perfect crystal structure and complete elimination of defects [31,32,33,34,35,36,37,38]. Furthermore, using expensive and unstable TMS as the sulfur source in colloidal AgBiS2 NCs synthesis process results in high cost of the synthesis reaction, which is not conducive to large–scale preparation of colloidal AgBiS2 NC. It is urgent to explore a method to synthesize high quality colloidal AgBiS2 NC with low cost.
As a cheap sulfur source, oleylamine@sulfur (OLA–S) solution has been widely used to synthesize high quality CdS [39], ZnS [40], and PbS NCs [41,42]. Moreover, as a ligand stabilizer, oleylamine can passivate the surface defects of NCs and regulate their growth [23,43]. Recently, Nakazawa et al. successfully synthesized colloidal AgBiS2 NCs with OLA–S as sulfur source [44]. However, the effect of OLA–S on the growth and quality of colloidal AgBiS2 NCs has not been studied in detail. In this work, in order to improve the quality of colloidal AgBiS2 NCs and reduce the preparation cost, we synthesized colloidal AgBiS2 NCs by hot injection method and using OLA–S as the sulfur source to replace TMS, then investigated the effect of OLA–S on the growth of colloidal AgBiS2 NCs. By regulating the dosage of OLA–S, high quality colloidal AgBiS2 NCs with low defect density and uniform size were prepared. It was found that the size of colloidal AgBiS2 NCs could be controlled by changing the reaction injection temperature. Compared with TMS–based colloidal AgBiS2 NCs, the colloidal AgBiS2 NCs based on OLA–S have good crystallinity and fewer defects. In addition, the synthesis cost is significantly reduced.

2. Materials and Methods

Chemicals and materials. Bismuth (III) acetate (Bi(OAc)3, 99.9%, Aladdin, China), silver acetate (AgOAc, 99.5%, Aladdin, China), oleic acid (OA, 90%, Sigma–Aldrich, USA), hexamethyldisilathiane (TMS, synthesis grade, Sigma–Aldrich, Switzerland), 1-Octadecene (ODE, 90%, Aladdin, China), oleylamine (OLA, 80–90%, Aladdin, China), sulfur (S, 99.5%, Sinopharm, China), acetone (99.5%, Sinopharm, China), toluene (99.5%, Sinopharm, China), hexane(97.0%, Sinopharm, China).
Synthesis of TMS–based colloidal AgBiS2 NCs. The TMS–based colloidal AgBiS2 NCs were synthesized following a hot injection method reported previously in the literature [29].
Synthesis of OLA–S–based colloidal AgBiS2 NCs. The synthesis of OLA–S–based colloidal AgBiS2 NCs was carried out similarly to that of TMS–based colloidal AgBiS2 NCs. The OLA–S (1 M) solution was prepared by dissolving sulfur powder in OLA solution at room temperature. To investigate the influence of the reaction injection temperature (80 °C, 100 °C, 120 °C) on synthesis of OLA–S–based colloidal AgBiS2 NCs, the dosage of OLA–S solution (1 M) was kept at 3 mL. In contrast, to investigate the influence of the dosage of OLA–S solution (1 mL, 2 mL, 3 mL and 4 mL) on synthesis of OLA–S–based colloidal AgBiS2 NCs, the reaction injection temperature was kept at 100 °C. After rapidly injected the OLA–S solution into the cationic solution, heating was stopped while stirring was maintained until the reaction solution cooled to room temperature. The colloidal AgBiS2 NCs were collected using an acetone/toluene solvent system in air, and the obtained colloidal AgBiS2 NCs precipitate was then dried in a vacuum drying oven at room temperature. Finally, the colloidal AgBiS2 NCs were dispersed in hexane solution.

Characterization

The phase identification of colloidal AgBiS2 NCs was carried out using powder X–ray diffraction (XRD, Bruker, D8 ADVANCED, Germany). The morphology and crystal structure of the prepared colloidal AgBiS2 NCs were characterized using high–resolution transmission electron microscopy (HR–TEM, Thermo Fisher Scientific, Talos F200x, USA). The TEM samples were prepared as follows: the obtained colloidal AgBiS2 NCs were dispersed in hexane with a concentration of 2 mg/mL, three drops of the mixture solution were dropped onto the TEM support film via pipette gun, and the TEM support films were fully dried in the vacuum drying oven before measurement. The acceleration voltage of electron beam was 200 KV during the TEM test. Elemental mapping of the colloidal AgBiS2 NCs was carried out using a TEM equipped with an energy dispersive spectrometer (EDS). UV–vis–NIR absorption spectra of the colloidal AgBiS2 NCs were recorded with a spectrophotometer (Shimadzu, UV–3600 PLUS, Japan). Fourier transform–infrared spectra (FT–IR) of colloidal AgBiS2 NCs were recorded using an infrared spectrometer (Bruker, INVENIO, Germany). The photoluminescent (PL) spectra of the colloidal AgBiS2 NCs were measured by using a fluorescence spectrophotometer (Zolix instruments Co., Ltd, FST1–MPL303–OP1, Beijing, China).

3. Results and Discussion

Following the schematic diagram of the synthesis of colloidal AgBiS2 NCs by hot injection method, which is shown in Figure 1, we adopted two methods to synthesize colloidal AgBiS2 NCs using TMS (path 1) and OLA–S (path 2) as the sulfur source. The variation of the crystal structure and phase of TMS–based and OLA–S–based colloidal AgBiS2 NCs were studied using X–ray diffraction (XRD). Figure 2a shows the XRD patterns of the TMS–based and OLA–S–based colloidal AgBiS2 NCs (dosage of OLA–S: 3 mL, injection temperature: 100 °C). For both samples, six XRD peaks are found at 13.75°, 15.85°, 22.89°, 27.12°, 28.20°, and 33.11°, which can be assigned to the facets (111), (200), (220), (311), (222), and (400) of AgBiS2 cubic rock salt structure (JCPDS No. 21–1178), respectively [8,12,37,45,46,47]. This indicates that colloidal AgBiS2 NCs can be successfully prepared by using OLA–S as the sulfur source. For NC materials, the broadening of XRD peaks can be roughly attributed to the size effect of NCs and lattice strain, which is closely related to defects in the NCs [48]. The lattice strains in TMS–based and OLA–S–based colloidal AgBiS2 NCs were calculated using the Williams–Hall plot method [49], with the results shown in Figure S1 and the corresponding Table S1. The obtained values of the lattice strains for TMS–based and OLA–S–based colloidal AgBiS2 NCs are 0.33 and 0.27, respectively. It can be seen that the lattice strain for TMS–based colloidal AgBiS2 NCs is larger than that of the OLA–S–based NCs, with the larger lattice strain leading to more defects in the NCs. Thus, the results demonstrate that OLA–S–based colloidal AgBiS2 NCs have lower defect density than TMS–based NCs. In other words, as a new sulfur source, OLA–S can be used to synthesize high quality colloidal AgBiS2 NCs.
Furthermore, in order to reveal the influence of the sulfur source on the morphology and size distribution of colloidal AgBiS2 NCs, transmission electron microscopy (TEM) measurement was carried out. Figure S2 shows that colloidal AgBiS2 NCs prepared using TMS and OLA–S have a uniform spherical shape. The average diameters of TMS–based and OLA–S–based colloidal AgBiS2 NCs are about 4.3 ± 1.6 nm (Figure S2a) and 7.8 ± 1.9 nm (Figure S2b), respectively. High–resolution TEM (HR–TEM) images of spherical colloidal AgBiS2 NCs based on TMS and OLA–S are shown in Figure 2b,c [50,51,52]. Lattice spacing of about 0.32 nm, corresponding to (111) crystal plane spacing of cubic rock salt structure AgBiS2, can be clearly seen, which is consistent with the XRD results above. Elemental mapping of TMS–based colloidal AgBiS2 NCs was carried out using a TEM equipped with EDS. Elemental mapping images are provided in Figure S3a–c, where the Ag, Bi, and S elements are distributed throughout all NCs. Moreover, the atomic percentages of Ag, Bi, and S elements are 41.3, 22.3, and 36.4, respectively. These results indicate that the surface of TMS–based colloidal AgBiS2 NCs should be silver–rich and sulfur–poor. Figure S3d–f shows the element mapping images of OLA–S–based colloidal AgBiS2 NCs. The atomic percentages of Ag, Bi, and S elements are 42.6, 12.3, and 45.1, respectively. This means that the surface of OLA–S–based colloidal AgBiS2 NCs is both silver–rich and sulfur–rich. Combined with the XRD result of the OLA–S–based colloidal AgBiS2 NCs, the signal of Ag2S is not observed in Figure 2a. Therefore, we speculate that the excess Ag and S are coating on the surface of colloidal AgBiS2 NCs rather than forming Ag2S impurities.
To further reveal the effect of the sulfur source on the quality of the obtained colloidal AgBiS2 NCs, the photoluminescent (PL) spectra of the TMS–based and OLA–S–based colloidal AgBiS2 NCs solutions were measured at room temperature [53]. The PL spectra of colloidal AgBiS2 NCs based on TMS and OLA–S are shown in Figure 3a. We detected PL signals in the emission range of 400–1000 nm using a 330 nm excitation wavelength. It can be seen that the PL intensity of OLA–S–based colloidal AgBiS2 NCs is stronger than that of TMS–based colloidal AgBiS2 NCs under the same concentration. This indicates that there are fewer nonradiative charge recombination centers, which is closely related to lower defect density, in the OLA–S–based colloidal AgBiS2 NCs compared to the TMS–based colloidal AgBiS2 NCs.
The UV–vis–NIR absorption spectra of TMS–based and OLA–S–based colloidal AgBiS2 NCs are shown in Figure 3b. It can be seen that the absorption spectra of both of TMS–based and OLA–S–based colloidal AgBiS2 NCs do not exhibit any exciton absorption peak or sharp absorption edge in the range of 300–1600 nm, which is similar to the previously reported colloidal AgBiS2 NCs and other ternary semiconductor NCs such as colloidal CuInS2, AgInS2 and AgBiSe2 NCs [6,54,55,56]. The absorption spectrum of OLA–S–based colloidal AgBiS2 NCs extends to a larger IR region than the TMS–based colloidal AgBiS2 NCs, which may be owing to the size effect of larger grain size of OLA–S–based colloidal AgBiS2 NCs. This result is consistent with the result of the TEM images shown above.
FT–IR spectroscopy was employed to investigate the functional groups on the obtained colloidal AgBiS2 NCs [44,57]. Figure 3c shows the FT–IR spectra of colloidal AgBiS2 NCs. For TMS–based colloidal AgBiS2 NCs, the =C–H stretching peak at 3006 cm−1, C–H absorption stretching peaks at 2961, 2929, and 2854 cm−1, and C=O stretching mode at 1678 cm−1 are assigned to OA ligands on the surface of colloidal AgBiS2 NCs [55,58]. For OLA–S–based colloidal AgBiS2 NCs, the peaks at 903 cm−1, 721 cm−1, and 641 cm−1 are attributed to N–H vibrations of OLA. Additionally, a C–N stretching mode peak was observed at 1145 cm−1. The peak at 1678 cm−1 can be assigned to C=O stretching of OA ligands, indicating that the OLA–S–based colloidal AgBiS2 NCs are coated by OLA and OA ligands and that there are no amide–type ligands on the surface of the colloidal AgBiS2 NCs.
In order to reveal the influence of reaction temperature on the synthesis of OLA–S–based colloidal AgBiS2 NCs, a synthetic experiment was carried out at 80 °C, 100 °C, and 120 °C. The crystal structure and phase transition of colloidal AgBiS2 NCs based on OLA–S were investigated using XRD. Figure 4a shows the XRD patterns of OLA–S–based colloidal AgBiS2 NCs which were synthesized under different injection temperatures. The results show that colloidal AgBiS2 NCs can be successfully prepared at 100 °C and 120 °C. For sample prepared at 80 °C, the weak diffraction peak intensity indicates poor crystallinity of the obtained product and there are peaks belonging to monoclinic structure Ag2S (JCPDS No. 14–0072) at which impurities can be found. Therefore, pure colloidal AgBiS2 NCs cannot be synthesized at low temperature when using OLA–S as the sulfur source. The lattice strains of OLA–S–based colloidal AgBiS2 NCs obtained at 100 °C and 120 °C were calculated using the Williams–Hall plot from Figure S4, the corresponding results are shown in Table S2. The values of the lattice strains for OLA–S–based colloidal AgBiS2 NCs prepared at 100 °C and 120 °C are 0.27 and 0.38, respectively. It can be seen that the lattice strain of OLA–S–based colloidal AgBiS2 NCs produced at 120 °C is greater than the lattice strain OLA–S–based colloidal AgBiS2 NCs prepared at 100 °C, with larger lattice strain leading to more defects in the NCs. These results show that the defect density OLA–S–based colloidal AgBiS2 NCs prepared at 100 °C was lower than that of OLA–S–based NCs produced at 120 °C, meaning that 100 °C can be used to synthesize high quality colloidal AgBiS2 NC.
TEM measurement was used to reveal the influence of reaction temperature on the morphology and size distribution of OLA–S–based colloidal AgBiS2 NCs. Figure S5 shows that the colloidal AgBiS2 NCs prepared at 120 °C have a uniform spherical shape and that the average diameter of the obtained NCs is 12.1 ± 3.5 nm (Figure S5a). An HR–TEM image of colloidal AgBiS2 NCs prepared at 120 °C is shown in Figure S5b. The cubic rock salt structure of AgBiS2 with a lattice spacing of about 0.32 nm corresponding to the (111) crystal plane spacing can be seen, and is consistent with the XRD results. Figure 4b–d provides elemental diagram images of OLA–S–based colloidal AgBiS2 NCs synthesized at 120 °C. The atomic percentages of Ag, Bi, and S are 48.4, 17.1, and 34.6, respectively. These results show that the surface of the obtained colloidal AgBiS2 NCs should be silver–rich and sulfur–rich. Because the lattice strain of the OLA–S–based colloidal AgBiS2 NCs prepared at 120 °C is larger than that prepared at 100 °C, high reaction temperature may be detrimental to the preparation of high quality OLA–S–based colloidal AgBiS2 NCs.
Figure S6a shows the UV–vis–NIR absorption spectra of OLA–S–based colloidal AgBiS2 NCs prepared at 100 °C and 120 °C. The absorption spectrum of the sample synthesized at 120 °C extends to more IR regions than that of the NCs prepared at 100 °C, which is mainly caused by the size effect of the colloidal AgBiS2 NCs. To further reveal the effect of reaction injection temperature on the quality of the obtained colloidal AgBiS2 NCs, the PL spectra of OLA–S–based colloidal AgBiS2 NCs prepared at 100 °C and 120 °C were measured using 330 nm excitation wavelength at room temperature. The PL intensity of OLA–S–based colloidal AgBiS2 NCs synthesized at 100 °C is stronger than that of NCs synthesized at 120 °C (Figure S6b). This reflects that the OLA–S–based colloidal AgBiS2 NCs prepared at 100 °C has fewer defects than the colloidal AgBiS2 NCs synthesized at 120 °C.
In addition, the influence of OLA–S dosage in the reaction when synthesizing colloidal AgBiS2 NCs was investigated. We carried out the synthesis of colloidal AgBiS2 NCs using different amount of OLA–S (1 M). The crystallinity and phase purity of OLA–S–based colloidal AgBiS2 NCs synthesized under different conditions were examined by XRD. (Figure 5). It can be seen that only Ag2S was produced in the reaction when using 1 mL OLA–S, while both Ag2S and AgBiS2 were formed when the dosage of OLA–S was increased to 2 mL. Thus, pure phase colloidal AgBiS2 NCs can be successfully prepared using 3 mL and 4 mL OLA–S. Based on these results, we speculate that the reaction process of OLA–S–based colloidal AgBiS2 NC synthesis should be that: H2S being released in situ from OLA–S after OLA-S was injected into high temperature cationic solution [59], and then reacting with Ag+ to form Ag2S, subsequently, the produced Ag2S reacts with Bi3+ to form AgBiS2 NCs. The reaction equations are as follows:
2 Ag + + H 2 S   Ag 2 S + 2 H + ,
2 Ag 2 S + Bi 3 + AgBiS 2 + 3 Ag +
The lattice strain values calculated for the colloidal AgBiS2 NCs prepared by using 3 mL and 4 mL OLA–S are 0.27 and 0.40, respectively, as shown in Figure S7 and Table S3. The smaller lattice strain in colloidal AgBiS2 NCs prepared using 3 mL OLA–S indicates a lower defect density in the NCs. To clarify the effect of the dosage of OLA–S on the morphology and size distribution of colloidal AgBiS2 NCs, we carried out HR–TEM measurements of the obtained AgBiS2 NCs. Figure S8a shows that colloidal AgBiS2 NCs prepared using 4 mL OLA–S have a uniform spherical shape and the average diameters of NCs are about 8.3 ± 2.7 nm, which is a little larger than that based on 3 mL OLA–S (7.8 ± 1.9 nm; see Figure S2b). Figure S6c shows the UV–vis–NIR absorption spectra of 3 mL and 4 mL OLA–S–based colloidal AgBiS2 NCs. The absorption spectrum of the 4 mL OLA–S–based colloidal AgBiS2 NCs extends to more IR regions than that of the 3 mL OLA–S–based colloidal AgBiS2 NCs, which is mainly caused by the size effect of the colloidal AgBiS2 NCs.
Finally, we calculated the cost and yield of the synthesis of colloidal AgBiS2 NCs using TMS and OLA–S as sulfur sources. Compared with TMS, the cost of synthesis using OLA–S was reduced about 57% (see Table S4 for details). We found that the yield of colloidal AgBiS2 NCs based on TMS and OLA–S were 86.9% and 90.5%, respectively. This indicates that OLA–S can replace TMS for synthesizing colloidal AgBiS2 NCs as a low–cost sulfur source.

4. Conclusions

In summary, we optimized the synthesis conditions of colloidal AgBiS2 NCs using OLA–S as the sulfur source. Colloidal AgBiS2 NCs with high quality and pure phase can be synthesized by rationally optimizing the dosage and reaction injection temperature of OLA–S. Compared with TMS as the sulfur source, OLA–S–based colloidal AgBiS2 NCs have lower defect density and a wider absorption spectrum. More importantly, the cost of synthesis is greatly reduced, as the yield of colloidal AgBiS2 NCs is increased when using OLA–S to replace the more expensive TMS. In addition, the pungent odor and air sensitivity of TMS can be avoided. As a low–cost sulfur source, OLA–S has the potential to for use in synthesizing other sulfide NCs in the future.

Supplementary Materials

The supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/nano12213742/s1. Figure S1. The William–Hall plots for (a) TMS–based and (b) OLA–S–based colloidal AgBiS2 NCs. Figure S2. TEM images of (a) TMS–based and (b) OLA–S–based colloidal AgBiS2 NCs. Insets are the statistical histograms of the NCs size distribution. SAED images of (c) TMS–based and (d) OLA–S–based colloidal AgBiS2 NCs. Figure S3. EDS–Mapping images of (a), (b) and (c)TMS–based. EDS–Mapping images of (d), (e) and (f) OLA–S–based colloidal AgBiS2 NCs. Figure S4. The William–Hall plots for (a) and (b) OLA–S–based colloidal AgBiS2 NCs prepare at 100 °C and 120 °C. Figure S5. TEM images of (a) OLA–S–based colloidal AgBiS2 NCs prepare at 120 °C. Insets are the statistical histograms of the NCs size distribution. (b) HR–TEM images of OLA–S–based colloidal AgBiS2 NCs prepare at 120 °C. (c) SAED image of OLA–S–based colloidal AgBiS2 NCs prepare at 120 °C. Figure S6. (a) UV–vis–NIR absorption spectra of the colloidal AgBiS2 NCs in hexane. (b) PL spectrum and fitting curves of OLA–S–based colloidal AgBiS2 NCs prepare at 100 °C and 120 °C. (b) UV–vis–NIR absorption spectra of the 3 mL and 4 mL colloidal AgBiS2 NCs in hexane. Figure S7. The William–Hall plots for (a) 3 mL and (b) 4 mL OLA–S–based colloidal AgBiS2 NCs. Figure S8. TEM images of (a) 4 mL OLA–S–based colloidal AgBiS2 NCs. Insets are the statistical histograms of the NCs size distribution. (b) HR-TEM images of 4 mL OLA–S–based colloidal AgBiS2 NCs. (c) SAED image of 4 mL OLA–S–based colloidal AgBiS2 NCs. Table S1. Half–peak widths and lattice strains of OLA–S–based and TMS–based colloidal AgBiS2 NCs. Table S2. Half–peak widths and lattice strains of OLA–S–based colloidal AgBiS2 NCs prepare at 100 °C and 120 °C. Table S3. Half–peak widths and lattice strains of 3 mL and 4 mL OLA–S–based colloidal AgBiS2 NCs. Table S4. Raw material usage, price and yield of AgBiS2 NCs based on TMS and OLA–S. Reference [60] is cited in the Supplementary Materials.

Author Contributions

Q.L. and Y.Z. conceived and designed the experiments; Q.L., X.Z., X.S. and S.D. performed the experiments; Q.L., X.Z. and Y.Z. analyzed the data; H.F., G.W. and Y.Z. contributed reagents/materials/analysis tools; Q.L. wrote the paper; Y.Z. and G.W. corrected the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China under grant No. 52103223.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bernechea, M.; Miller, N.C.; Xercavins, G.; So, D.; Stavrinadis, A.; Konstantatos, G. Solution-processed solar cells based on environmentally friendly AgBiS2 nanocrystals. Nat. Photon. 2016, 10, 521–525. [Google Scholar] [CrossRef]
  2. Liu, Y.; Cadavid, D.; Ibanez, M.; De Roo, J.; Ortega, S.; Dobrozhan, O.; Kovalenko, M.V.; Cabot, A. Colloidal AgSbSe2 nanocrystals: Surface analysis, electronic doping and processing into thermoelectric nanomaterials. J. Mater. Chem. C 2016, 4, 4756–4762. [Google Scholar] [CrossRef] [Green Version]
  3. Liu, Y.; Garcia, G.; Ortega, S.; Cadavid, D.; Palacios, P.; Lu, J.; Ibanez, M.; Xi, L.; De Roo, J.; Lopez, A.M.; et al. Solution-based synthesis and processing of Sn- and Bi-doped Cu3SbSe4 nanocrystals, nanomaterials and ring-shaped thermoelectric generators. J. Mater. Chem. A 2017, 5, 2592–2602. [Google Scholar] [CrossRef] [Green Version]
  4. Zhou, B.; Li, M.; Wu, Y.; Yang, C.; Zhang, W.H.; Li, C. Monodisperse AgSbS2 nanocrystals: Size-control strategy, large-scale synthesis, and photoelectrochemistry. Chem. Eur. J. 2015, 21, 11143–11151. [Google Scholar] [CrossRef] [PubMed]
  5. Guria, A.K.; Prusty, G.; Chacrabarty, S.; Pradhan, N. Fixed aspect ratio rod-to-rod conversion and localized surface plasmon resonance in semiconducting I-V-VI nanorods. Adv. Mater. 2016, 28, 447–453. [Google Scholar] [CrossRef] [PubMed]
  6. Kuznetsova, V.; Tkach, A.; Cherevkov, S.; Sokolova, A.; Gromova, Y.; Osipova, V.; Baranov, M.; Ugolkov, V.; Fedorov, A.; Baranov, A. Spectral-time multiplexing in FRET complexes of AgInS2/ZnS quantum dot and organic dyes. Nanomaterials 2020, 10, 1569. [Google Scholar] [CrossRef]
  7. Bai, X.; Purcell-Milton, F.; Gun’ko, Y.K. Optical properties, Synthesis, and potential applications of Cu-Based ternary or quaternary anisotropic quantum dots, polytypic nanocrystals, and core/shell heterostructures. Nanomaterials 2019, 9, 85. [Google Scholar] [CrossRef] [Green Version]
  8. Pejova, B.; Nesheva, D.; Aneva, Z.; Petrova, A. Photoconductivity and relaxation dynamics in sonochemically synthesized assemblies of AgBiS2 quantum dots. J. Mater. Chem. C 2011, 115, 37–46. [Google Scholar] [CrossRef]
  9. Smith, A.M.; Duan, H.; Rhyner, M.N.; Ruan, G.; Nie, S. A systematic examination of surface coatings on the optical and chemical properties of semiconductor quantum dots. Phys. Chem. Chem. Phys. 2006, 8, 3895–3903. [Google Scholar] [CrossRef]
  10. Do, S.; Kwon, W.; Rhee, S.W. Soft-template synthesis of nitrogen-doped carbon nanodots: Tunable visible-light photoluminescence and phosphor-based light-emitting diodes. J. Mater. Chem. C 2014, 2, 4221–4226. [Google Scholar] [CrossRef]
  11. Vines, F.; Bernechea, M.; Konstantatos, G.; Illas, F. Matildite versus schapbachite: First-principles investigation of the origin of photoactivity in AgBiS2. Phys. Rev. B 2016, 94, 235203. [Google Scholar] [CrossRef] [Green Version]
  12. Hameed, T.A.; Wassel, A.R.; El Radaf, I.M. Investigating the effect of thickness on the structural, morphological, optical and electrical properties of AgBiSe2 thin films. J. Alloys Compd. 2019, 805, 1–11. [Google Scholar] [CrossRef]
  13. Zhou, S.; Yang, J.; Li, W.; Jiang, Q.; Luo, Y.; Zhang, D.; Zhou, Z.; Li, X. Preparation and photovoltaic properties of ternary AgBiS2 quantum dots sensitized TiO2 nanorods photoanodes by electrochemical atomic layer deposition. J. Electrochem. Soc. 2016, 163, D63–D67. [Google Scholar] [CrossRef]
  14. Zhang, Y.H.; Wu, G.H.; Ding, C.; Liu, F.; Liu, D.; Masuda, T.; Yoshino, K.; Hayase, S.; Wang, R.X.; Shen, Q. Surface-modified graphene oxide/lead sulfide hybrid film-forming ink for high-efficiency bulk nano-heterojunction colloidal quantum dot solar cells. Nano Micro Lett. 2020, 12, 111. [Google Scholar] [CrossRef]
  15. Zhang, Y.H.; Wu, G.H.; Ding, C.; Liu, F.; Yao, Y.F.; Zhou, Y.; Wu, C.P.; Nakazawa, N.; Huang, Q.X.; Toyoda, T.; et al. Lead selenide colloidal quantum dot solar cells achieving high open-circuit voltage with one-step deposition strategy. J. Phys. Chem. Lett. 2018, 9, 3598–3603. [Google Scholar] [CrossRef]
  16. Zhang, Y.H.; Ding, C.; Wu, G.H.; Nakazawa, N.; Chang, J.; Ogomi, Y.; Toyoda, T.; Hayase, S.; Katayama, K.; Shen, Q. Air stable PbSe colloidal quantum dot heterojunction solar cells: Ligand-dependent exciton dissociation, recombination, photovoltaic property, and stability. J. Phys. Chem. C 2016, 120, 28509–28518. [Google Scholar] [CrossRef]
  17. Wang, X.C.; Bai, T.X.; Yang, B.; Zhang, R.L.; Zheng, D.Y.; Jiang, J.K.; Tao, S.X.; Liu, F.; Han, K.L. Germanium halides serving as ideal precursors: Designing a more effective and less toxic route to high-optoelectronic-quality metal halide perovskite nanocrystals. Nano Lett. 2022, 22, 636–643. [Google Scholar] [CrossRef]
  18. Wang, X.C.; Bai, T.X.; Meng, X.; Ji, S.J.; Zhang, R.L.; Zheng, D.Y.; Yang, B.; Jiang, J.K.; Han, K.L.; Liu, F. Filling chlorine vacancy with bromine: A two-step hot-injection approach achieving defect-free hybrid halogen perovskite nanocrystals. ACS Appl. Mater. Interfaces 2022, 14, 18353–18359. [Google Scholar] [CrossRef]
  19. Zeng, Y.T.; Li, Z.R.; Chang, S.P.; Ansay, A.; Wang, Z.H.; Huang, C.Y. Bright CsPbBr3 perovskite nanocrystals with improved stability by in-situ Zn-doping. Nanomaterials 2022, 12, 759. [Google Scholar] [CrossRef]
  20. Zhang, W.Z.; Li, X.; Peng, C.C.; Yang, F.; Lian, L.Y.; Guo, R.D.; Zhang, J.B.; Wang, L. CsPb(Br/Cl)(3) perovskite nanocrystals with bright blue emission synergistically modified by calcium halide and ammonium ion. Nanomaterials 2022, 12, 2026. [Google Scholar] [CrossRef]
  21. Wang, J.; Xiong, L.; Bai, Y.; Chen, Z.J.; Zheng, Q.; Shi, Y.Y.; Zhang, C.; Jiang, G.C.; Li, Z.Q. Mn-doped perovskite nanocrystals for photocatalytic CO2 reduction: Insight into the role of the charge carriers with prolonged lifetime. Solar RRL 2022, 6, 2200294. [Google Scholar] [CrossRef]
  22. Li, N.; Chen, X.; Wang, J.; Liang, X.; Ma, L.; Jing, X.; Chen, D.L.; Li, Z. ZnSe nanorods–CsSnCl3 perovskite heterojunction composite for photocatalytic CO2 reduction. ACS Nano 2022, 16, 3332–3340. [Google Scholar] [CrossRef] [PubMed]
  23. Hu, L.; Patterson, R.J.; Zhang, Z.; Hu, Y.; Li, D.; Chen, Z.; Yuan, L.; Teh, Z.L.; Gao, Y.; Conibeer, G.J.; et al. Enhanced optoelectronic performance in AgBiS2 nanocrystals obtained via an improved amine-based synthesis route. J. Mater. Chem. C 2018, 6, 731–737. [Google Scholar] [CrossRef]
  24. Oh, J.T.; Cho, H.; Bae, S.Y.; Lim, S.J.; Kang, J.; Jung, I.H.; Choi, H.; Kim, Y. Improved size distribution of AgBiS2 colloidal nanocrystals by optimized synthetic route enhances photovoltaic performance. Int. J. Energy Res. 2020, 44, 11006–11014. [Google Scholar] [CrossRef]
  25. Zafer Akgul, M.; Figueroba, A.; Pradhan, S.; Bi, Y.; Konstantatos, G. Low-cost RoHS compliant solution processed photovoltaics enabled by ambient condition synthesis of AgBiS2 nanocrystals. ACS Photon. 2020, 7, 588–595. [Google Scholar] [CrossRef] [Green Version]
  26. Liang, N.; Chen, W.; Dai, F.; Wu, X.; Zhang, W.; Li, Z.; Shen, J.; Huang, S.; He, Q.; Zai, J.; et al. Homogenously hexagonal prismatic AgBiS2 nanocrystals: Controlled synthesis and application in quantum dot-sensitized solar cells. Crystengcomm 2015, 17, 1902–1905. [Google Scholar] [CrossRef]
  27. Aldakov, D.; Lefrancois, A.; Reiss, P. Ternary and quaternary metal chalcogenide nanocrystals: Synthesis, properties and applications. J. Mater. Chem. C 2013, 1, 3756–3776. [Google Scholar] [CrossRef]
  28. Burgues-Ceballos, I.; Wang, Y.; Akgul, M.Z.; Konstantatos, G. Colloidal AgBiS2 nanocrystals with reduced recombination yield 6.4% power conversion efficiency in solution-processed solar cells. Nano Energy 2020, 75, 104961. [Google Scholar] [CrossRef]
  29. Wang, Y.; Kavanagh, S.R.; Burgues-Ceballos, I.; Walsh, A.; Scanlon, D.; Konstantatos, G. Cation disorder engineering yields AgBiS2 nanocrystals with enhanced optical absorption for efficient ultrathin solar cells. Nat. Photon. 2022, 16, 235–241. [Google Scholar] [CrossRef]
  30. Bae, S.Y.; Oh, J.T.; Park, J.Y.; Ha, S.R.; Choi, J.; Choi, H.; Kim, Y. Improved eco-friendly photovoltaics based on stabilized AgBiS2 nanocrystal inks. Chem. Mater. 2020, 32, 10007–10014. [Google Scholar] [CrossRef]
  31. Gao, Y.; Peng, X. Photogenerated excitons in plain core CdSe nanocrystals with unity radiative decay in single channel: The effects of surface and ligands. J. Am. Chem. Soc. 2015, 137, 4230–4235. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, F.; Huang, S.; Wang, P.; Chen, X.; Zhao, S.; Dong, Y.; Zhong, H. Colloidal synthesis of air-stable CH3NH3PbI3 quantum dots by gaining chemical insight into the solvent effects. Chem. Mater. 2017, 29, 3793–3799. [Google Scholar] [CrossRef]
  33. Oh, J.T.; Bae, S.Y.; Yang, J.; Ha, S.R.; Song, H.; Lee, C.B.; Shome, S.; Biswas, S.; Lee, H.M.; Seo, Y.H.; et al. Ultra-stable all-inorganic silver bismuth sulfide colloidal nanocrystal photovoltaics using pin type architecture. J. Power Sources 2021, 514, 230585. [Google Scholar] [CrossRef]
  34. Chen, D.; Shivarudraiah, S.B.; Geng, P.; Ng, M.; Li, C.H.A.; Tewari, N.; Zou, X.; Wong, K.S.; Guo, L.; Halpert, J.E. Solution-processed, inverted AgBiS2 nanocrystal solar cells. ACS Appl. Mater. Interfaces 2022, 14, 1634–1642. [Google Scholar] [CrossRef]
  35. Kirchartz, T.; Deledalle, F.; Tuladhar, P.S.; Durrant, J.R.; Nelson, J. On the differences between dark and light ideality factor in polymer: Fullerene solar cells. J. Phys. Chem. Lett. 2013, 4, 2371–2376. [Google Scholar] [CrossRef]
  36. Cowan, S.R.; Roy, A.; Heeger, A.J. Recombination in polymer-fullerene bulk heterojunction solar cells. Phys. Rev. B 2010, 82, 245207. [Google Scholar] [CrossRef] [Green Version]
  37. Chen, C.; Qiu, X.; Ji, S.; Jia, C.; Ye, C. The synthesis of monodispersed AgBiS2 quantum dots with a giant dielectric constant. Crystengcomm 2013, 15, 7644–7648. [Google Scholar] [CrossRef]
  38. Huang, P.C.; Yang, W.C.; Lee, M.W. AgBiS2 semiconductor-sensitized solar cells. J. Phys. Chem. C 2013, 117, 18308–18314. [Google Scholar] [CrossRef]
  39. He, J.G.; Chen, J.; Yu, Y.; Zhang, L.; Zhang, G.Z.; Jiang, S.L.; Liu, W.; Song, H.S.; Tang, J. Effect of ligand passivation on morphology, optical and photoresponse properties of CdS colloidal quantum dots thin film. J. Mater. Sci.-Mater. Electron. 2014, 25, 1499–1504. [Google Scholar] [CrossRef]
  40. Rath, T.; Novák, J.; Amenitsch, H.; Pein, A.; Maier, E.; Haas, W.; Hofer, F.; Trimmel, G. Real time X-ray scattering study of the formation of ZnS nanoparticles using synchrotron radiation. Mater. Chem. Phys. 2014, 144, 310–317. [Google Scholar] [CrossRef]
  41. Liu, J.C.; Yu, H.Z.; Wu, Z.L.; Wang, W.L.; Peng, J.B.; Cao, Y. Size-tunable near-infrared PbS nanoparticles synthesized from lead carboxylate and sulfur with oleylamine as stabilizer. Nanotechnology 2008, 19, 345602. [Google Scholar] [CrossRef] [PubMed]
  42. Yuan, B.; Tian, X.C.; Shaw, S.; Petersen, R.E.; Cademartiri, L. Sulfur in oleylamine as a powerful and versatile etchant for oxide, sulfide, and metal colloidal nanoparticles. Phys. Status Solidi A 2017, 214, 1600543. [Google Scholar] [CrossRef]
  43. Ojha, G.P.; Pant, B.; Muthurasu, A.; Chae, S.H.; Park, S.J.; Kim, T.; Kim, H.Y. Three-dimensionally assembled manganese oxide ultrathin nanowires: Prospective electrode material for asymmetric supercapacitors. Energy 2019, 188, 116066. [Google Scholar] [CrossRef]
  44. Nakazawa, T.; Kim, D.; Oshima, Y.; Sato, H.; Park, J.; Kim, H. Synthesis and application of AgBiS2 and Ag2S nanoinks for the production of IR photodetectors. ACS Omega 2021, 6, 20710–20718. [Google Scholar] [CrossRef] [PubMed]
  45. Guin, S.N.; Biswas, K. Cation disorder and bond anharmonicity optimize the thermoelectric properties in kinetically stabilized rocksalt AgBiS2 nanocrystals. Chem. Mater. 2013, 25, 3225–3231. [Google Scholar] [CrossRef]
  46. Ju, M.G.; Dai, J.; Ma, L.; Zhou, Y.; Zeng, X.C. AgBiS2 as a low-cost and eco-friendly all-inorganic photovoltaic material: Nanoscale morphology-property relationship. Nanoscale Adv. 2020, 2, 770–776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pejova, B.; Grozdanov, I.; Nesheva, D.; Petrova, A. Size-dependent properties of sonochemically synthesized three-dimensional arrays of close-packed semiconducting AgBiS2 quantum dots. Chem. Mater. 2008, 20, 2551–2565. [Google Scholar] [CrossRef]
  48. Dolabella, S.; Borzi, A.; Dommann, A.; Neels, A. Lattice strain and defects analysis in nanostructured semiconductor materials and devices by high-resolution X-ray diffraction: Theoretical and practical aspects. Small Methods 2022, 6, 2100932. [Google Scholar] [CrossRef]
  49. Nishimura, K.; Hirotani, D.; Kamarudin, M.A.; Shen, Q.; Toyoda, T.; Iikubo, S.; Minemoto, T.; Yoshino, K.; Hayase, S. Relationship between lattice strain and efficiency for Sn-perovskite solar cells. ACS Appl. Mater. Interfaces 2019, 11, 31105–31110. [Google Scholar] [CrossRef]
  50. Li, X.P.; Huang, R.J.; Chen, C.; Li, T.; Gao, Y.J. Simultaneous conduction and valence band regulation of indium-based quantum dots for efficient H2 photogeneration. Nanomaterials 2021, 11, 1115. [Google Scholar] [CrossRef]
  51. Wang, Q.; Wang, J.; Wang, J.C.; Hu, X.; Bai, Y.; Zhong, X.; Li, Z. Coupling CsPbBr3 quantum dots with covalent triazine frameworks for visible-light-driven CO2 reduction. ChemSusChem 2021, 14, 1131–1139. [Google Scholar] [CrossRef] [PubMed]
  52. Chang, J.; Oshima, T.; Hachiya, S.; Sato, K.; Toyoda, T.; Katayama, K.; Hayase, S.; Shen, Q. Uncovering the charge transfer and recombination mechanism in ZnS-coated PbS quantum dot sensitized solar cells. Sol. Energy 2015, 122, 307–313. [Google Scholar] [CrossRef] [Green Version]
  53. Chang, J.; Zhang, S.; Wang, N.; Sun, Y.; Wei, Y.; Li, R.; Yi, C.; Wang, J.; Huang, W. Enhanced performance of red perovskite light-emitting diodes through the dimensional tailoring of perovskite multiple quantum wells. J. Phys. Chem. Lett. 2018, 9, 881–886. [Google Scholar] [CrossRef] [PubMed]
  54. Amaral-Junior, J.C.; Mansur, A.A.P.; Carvalho, I.C.; Mansur, H.S. Tunable luminescence of Cu-In-S/ZnS quantum dots-polysaccharide nanohybrids by environmentally friendly synthesis for potential solar energy photoconversion applications. Appl. Surf. Sci. 2021, 542, 148701. [Google Scholar] [CrossRef]
  55. Ma, Y.; Zhang, Y.; Liu, M.; Han, T.; Wang, Y.; Wang, X. Improving the performance of quantum dot sensitized solar cells by employing Zn doped CuInS2 quantum dots. Adv. Compos. Hybrid Mater. 2022, 5, 402–409. [Google Scholar] [CrossRef]
  56. Akgul, M.Z.; Konstantatos, G. AgBiSe2 colloidal nanocrystals for use in solar cells. ACS Appl. Nano Mater. 2021, 4, 2887–2894. [Google Scholar] [CrossRef]
  57. Ming, S.; Liu, X.; Zhang, W.; Xie, Q.; Wu, Y.; Chen, L.; Wang, H.Q. Eco-friendly and stable silver bismuth disulphide quantum dot solar cells via methyl acetate purification and modified ligand exchange. J. Clean. Prod. 2020, 246, 118966. [Google Scholar] [CrossRef]
  58. Lee, D.H.; Condrate, R.A.; Lacourse, W.C. FTIR spectral characterization of thin film coatings of oleic acid on glasses—Part II—Coatings on glass from different media such as water, alcohol, benzene and air. J. Mater. Sci. 2000, 35, 4961–4970. [Google Scholar] [CrossRef]
  59. Thomson, J.W.; Nagashima, K.; Macdonald, P.M.; Ozin, G.A. From sulfur–amine solutions to metal sulfide nanocrystals: Peering into the oleylamine−sulfur black box. J. Am. Chem. Soc. 2011, 133, 5036–5041. [Google Scholar] [CrossRef]
  60. Williamson, G.K.; Hall, W.H. X-ray line broadening from filed aluminium and wolfram L’elargissement des raies de rayons x obtenues des limailles d’aluminium et de tungsten Die verbreiterung der roentgeninterferenzlinien von aluminium- und wolframspaenen. Acta Metall. 1953, 1, 22–31. [Google Scholar] [CrossRef]
Figure 1. Synthetic route for colloidal AgBiS2 NCs using different sulfur sources (path 1: TMS, path 2: OLA–S).
Figure 1. Synthetic route for colloidal AgBiS2 NCs using different sulfur sources (path 1: TMS, path 2: OLA–S).
Nanomaterials 12 03742 g001
Figure 2. (a) XRD patterns of colloidal AgBiS2 NCs based on TMS and OLA–S; (b,c) HR–TEM images of TMS–based and OLA–S–based colloidal AgBiS2 NCs, respectively.
Figure 2. (a) XRD patterns of colloidal AgBiS2 NCs based on TMS and OLA–S; (b,c) HR–TEM images of TMS–based and OLA–S–based colloidal AgBiS2 NCs, respectively.
Nanomaterials 12 03742 g002
Figure 3. (a) PL spectra of TMS–based and OLA–S–based colloidal AgBiS2 NCs, (b) UV–vis–NIR absorption spectra of the colloidal AgBiS2 NCs in hexane, and (c) FT–IR spectra of colloidal AgBiS2 NCs.
Figure 3. (a) PL spectra of TMS–based and OLA–S–based colloidal AgBiS2 NCs, (b) UV–vis–NIR absorption spectra of the colloidal AgBiS2 NCs in hexane, and (c) FT–IR spectra of colloidal AgBiS2 NCs.
Nanomaterials 12 03742 g003
Figure 4. (a) XRD patterns of OLA–S–based colloidal AgBiS2 NCs synthesized at 80 °C, 100 °C, and 120 °C; (bd) EDS elemental mapping (TEM) images of OLA–S–based colloidal AgBiS2 NCs prepared at 120 °C.
Figure 4. (a) XRD patterns of OLA–S–based colloidal AgBiS2 NCs synthesized at 80 °C, 100 °C, and 120 °C; (bd) EDS elemental mapping (TEM) images of OLA–S–based colloidal AgBiS2 NCs prepared at 120 °C.
Nanomaterials 12 03742 g004
Figure 5. XRD patterns of colloidal OLA–S–based AgBiS2 NCs using 1 mL, 2 mL, 3 mL, and 4 mL.
Figure 5. XRD patterns of colloidal OLA–S–based AgBiS2 NCs using 1 mL, 2 mL, 3 mL, and 4 mL.
Nanomaterials 12 03742 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Q.; Zheng, X.; Shen, X.; Ding, S.; Feng, H.; Wu, G.; Zhang, Y. Optimizing the Synthetic Conditions of “Green” Colloidal AgBiS2 Nanocrystals Using a Low-Cost Sulfur Source. Nanomaterials 2022, 12, 3742. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12213742

AMA Style

Li Q, Zheng X, Shen X, Ding S, Feng H, Wu G, Zhang Y. Optimizing the Synthetic Conditions of “Green” Colloidal AgBiS2 Nanocrystals Using a Low-Cost Sulfur Source. Nanomaterials. 2022; 12(21):3742. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12213742

Chicago/Turabian Style

Li, Qiao, Xiaosong Zheng, Xiaoyu Shen, Shuai Ding, Hongjian Feng, Guohua Wu, and Yaohong Zhang. 2022. "Optimizing the Synthetic Conditions of “Green” Colloidal AgBiS2 Nanocrystals Using a Low-Cost Sulfur Source" Nanomaterials 12, no. 21: 3742. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12213742

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