In this Research Update, we briefly summarize some of the bismuth materials that have been investigated for their use in photovoltaic solar cells. We focus on bismuth-based perovskites and bismuth halides, as alternatives to lead-halide perovskites, and bismuth-based sulfides (Bi2S3, CuxBiySz, and AgBiS2), as alternatives to lead sulfide quantum dots. These materials fulfill the requirements of being composed of abundant and non-toxic elements. Moreover, they exhibit adequate properties for photovoltaics like high absorption coefficients and suitable bandgaps, plus additional attractive characteristics in terms of robustness and stability. However, they have not been extensively studied and therefore their efficiencies are still far from those reported for their toxic counterparts. Here we collect some of the most promising results, point at possible limiting factors, and suggest some routes to improve performance.

Perovskites and inorganic quantum dot solar cells offer the highest efficiencies among the “emerging PV” technologies.1 They also have the potential to be flexible and less expensive, thinner, and efficient over a wider range of light intensities than traditional or other emerging technologies. Furthermore, they can be composed of abundant elements, which is important both for reducing costs and assuring future production.2,3

However, the best performing devices have been fabricated using materials that have lead (Pb) in their composition: PbS or CsPbI3 in quantum dot solar cells4,5 or APbI3 (A = methylammonium or formamidinium) in perovskite solar cells.6 Although the concept and extent of toxicity can be discussed, the fact is that lead is listed among the 10 chemicals of major public health concern by the World Health Organization7 and its use is restricted under several legislations worldwide, therefore compromising the future commercialization of solar cells based on these materials.8 

In this sense, bismuth-based materials can be interesting alternatives for replacing lead-containing compounds. Bismuth is a quite abundant metal on the earth crust; moreover, it is a by-product of Pb, Cu, and Sn refining and has few significant commercial applications, resulting in the price of Bi being relatively low and stable (Fig. 1).2,3,9 Additionally, despite being a heavy metal, bismuth is considered non-toxic and is even used in common medicines such as Pepto-Bismol.10 Furthermore, Bi3+ has been suggested as an excellent candidate for defect tolerant compounds, i.e., materials with good optoelectronic properties despite the presence of defects. Supposedly, the active ns2 lone pair tends to create antibonding interactions at the valence band maximum, and as a result, defects are confined to shallow states at the band edges.11,12

FIG. 1.

Price evolution of bismuth compared to other elements relevant for photovoltaics. Main figure: comparison with indium, gallium, tellurium, or selenium. Inset: comparison with iodine, cadmium, lead, and copper. Data extracted from the 2018 Mineral Commodity Summaries (https://minerals.usgs.gov/minerals/pubs/mcs/).

FIG. 1.

Price evolution of bismuth compared to other elements relevant for photovoltaics. Main figure: comparison with indium, gallium, tellurium, or selenium. Inset: comparison with iodine, cadmium, lead, and copper. Data extracted from the 2018 Mineral Commodity Summaries (https://minerals.usgs.gov/minerals/pubs/mcs/).

Close modal

In this Research Update, we will briefly summarize some of the bismuth materials that have been investigated for their use in photovoltaic solar cells. We will focus on bismuth-based perovskites and bismuth halides, as alternatives to lead-halide perovskites, and bismuth-based sulfides (Bi2S3, CuxBiySz, and AgBiS2), as alternatives to lead sulfide quantum dots.

Lead-based perovskite solar cells have attracted a huge amount of attention due to their rapid rise in efficiency, with CH3NH3PbI3 solar cells recently achieving certified efficiencies in excess of 22%.6 These materials can be solution processed and have a variety of promising characteristics including high absorption coefficients, large charge-carrier diffusion lengths, and relatively low exciton-binding energies,13,14 closer to the binding energies reported for inorganic materials (6–15 meV), than those reported for organic solar cells (0.5 eV or larger).15 

Still, commercialization of lead perovskite solar cells is hindered by their short lifetimes and the use of toxic lead.16 Some hope that these limitations can be overcome by replacing lead with other elements like tin,17 antimony,18 or bismuth.11,16,19 Bismuth-based analogs to lead perovskite photovoltaic materials offer great promise for future development, although they have been scarcely explored.

Bismuth perovskites generally have the chemical formula A3Bi2X9, where A is a monovalent cation (i.e., Cs+ or CH3NH3+) and X is a halogen anion (i.e., Cl, Br, and/or I). (CH3NH3)3Bi2I9 and Cs3Bi2I9 are two bismuth perovskite materials that have been investigated thanks to their similarities to their high-efficiency lead counterparts CH3NH3PbI3 and CsPbI3, respectively.

The crystal structure of Cs3Bi2I9 was initially studied in the 1960s.20 Almost 50 years later, Park et al. first incorporated bismuth perovskite materials into solar cells, demonstrating power-conversion efficiencies of 1.09% for Cs3Bi2I9 and 0.12% for (CH3NH3)3Bi2I9.21 These modest efficiencies were attributed to high exciton binding energies (70-300 meV compared to 8-20 meV for lead perovskites22–24), significant non-radiative recombination due to defect states,25 non-optimal charge-extraction layers, and high bandgaps (2.1-2.2 eV). Despite their modest efficiencies, these materials exhibited high absorption coefficients and were much more air-stable than their lead counterparts.

The good stability of bismuth perovskites in both dry and humid air has since been repeatedly demonstrated. For instance, (CH3NH3)3Bi2I9 exhibited stable photovoltaic performance during 10 weeks in ambient air26 and 21 days in air with an average humidity of ∼50%.27 In another study, x-ray diffraction of (CH3NH3)3Bi2I9 after 25 days in ambient air demonstrated almost no change other than the formation of a thin, protective oxidation layer that likely prevents further degradation (Fig. 2).28 This is in stark contrast to the lead analog, CH3NH3PbI3, which almost fully converted to PbI2 during the same time period. Likewise, the bismuth perovskite exhibited only minimal visual changes after 26 days, whereas the lead analog changed from dark brown to light yellow after 5 days.

FIG. 2.

(a) Photographs of methylammonium bismuth iodide (MBI) and methylammonium lead iodide (MAPbI3) overtime in ambient air. [(b) and (c)] Normalized XRD patterns of MBI over time with air exposure. (d) The relative change in the normalized intensity of the diffraction peaks of MBI (day 25 vs. day 1). Reproduced with permission from Hoye et al., “Methylammonium bismuth iodide as a lead-free, stable hybrid organic-inorganic solar absorber,” Chem. - Eur. J. 22, 2605–2610 (2016). Copyright 2016 Wiley-VCH Verlag GmbH and Co. KGaA.

FIG. 2.

(a) Photographs of methylammonium bismuth iodide (MBI) and methylammonium lead iodide (MAPbI3) overtime in ambient air. [(b) and (c)] Normalized XRD patterns of MBI over time with air exposure. (d) The relative change in the normalized intensity of the diffraction peaks of MBI (day 25 vs. day 1). Reproduced with permission from Hoye et al., “Methylammonium bismuth iodide as a lead-free, stable hybrid organic-inorganic solar absorber,” Chem. - Eur. J. 22, 2605–2610 (2016). Copyright 2016 Wiley-VCH Verlag GmbH and Co. KGaA.

Close modal

Further investigation of these and other bismuth perovskites has led to improvements in the material properties. In particular, many studies have attempted to tune the bandgaps of bismuth perovskites since unaltered bismuth perovskites tend to have bandgaps above 2 eV, which is higher than desirable for optimal solar-cell performance. For example, one study demonstrated the ability to tune the bandgap of Cs2AgBiBr6 through pressure-induced changes in its crystal structure.29 Others have shown that sulfur doping can decrease the bandgap of Cs3Bi2I9 to a much more desirable value of 1.45 eV.30,31

Several groups have also improved the film quality and consequently the solar-cell performance of (CH3NH3)3Bi2I9 by replacing standard solution-processing methods with alternative deposition methods that result in smoother, more compact films with fewer pinholes. For instance, Ran et al. used a two-step method that combines evaporation and spin coating to push the power conversion efficiency of (CH3NH3)3Bi2I9 solar cells to 0.39%.32 Zhang et al. later employed a two-step vacuum deposition method to fabricate (CH3NH3)3Bi2I9 solar cells with a power conversion efficiency of 1.64% (0.83 V open-circuit voltage, 3.0 mA/cm2 short-circuit current, and 0.79 fill factor).33 The vacuum-processed solar cells exhibited charge-carrier diffusion lengths, trap densities, and absorption coefficients on par with many lead perovskite materials.

Likewise, new deposition techniques have resulted in improved Cs2AgBiBr6 film quality and consequently higher efficiencies. Cs2AgBiBr6 and its precursors exhibit low solubility in most common solvents, resulting in porous films full of cracks and pinholes. Dimethylsulfoxide (DMSO), on the other hand, has proven to be a good solvent for Cs2AgBiBr6 and its precursors: AgBr, CsBr, and BiBr3. Gruel et al. therefore dissolved the precursors in DMSO, heated the solution, and spin coated it onto a heated substrate.34 A subsequent annealing step at 250 °C was required to complete the formation of Cs2AgBiBr6 and maximize solar-cell performance. The best devices exhibited power-conversion efficiencies approaching 2.5% and an open-circuit voltage of 1.06 V, the highest value reported thus far for a bismuth-based perovskite. In a subsequent experiment, Gao et al. demonstrated the ability to deposit smooth films by dissolving Cs2AgBiBr6 in DMSO and spin coating the solution using the anti-solvent dropping method with isopropanol (IPA) as the anti-solvent.35 Films deposited without the anti-solvent were rough and frosted in appearance, whereas Cs2AgBiBr6 films deposited using the anti-solvent dropping method were very smooth and achieved efficiencies up to 2.2% and open-circuit voltages in excess of one volt. Again, a post-annealing treatment at 250 °C was required to produce high-quality, crystalline films.

Several related bismuth halides have also been explored as promising photovoltaic materials. For example, Kim et al. created dense, pinhole-free AgBi2I7 films by spin coating silver and bismuth precursors and subsequent annealing. The resulting air-stable material exhibited a bandgap of 1.87 eV and solar-cell efficiencies up to 1.22%.36 Bismuth triiodide (BiI3) has also been shown to be air-stable, to have a bandgap of ∼1.8 eV, and to exhibit efficiencies up to 1.0%.37–39 

One important distinction between bismuth perovskite and halide absorbers and their lead counterparts is the dimensionality of their octahedral networks. Figure 3 shows illustrations of 0D, 1D, 2D, and 3D octahedral networks and a few examples of bismuth perovskites and halides with each type of network. The octahedral network dimensionality can affect solar-cell performance by altering relevant material properties. Although there are notable exceptions, lower dimensional perovskites are usually associated with larger bandgaps, higher exciton binding energies, lower carrier mobilities, and better moisture stability due to more spatial confinement.40–45 

FIG. 3.

Illustrations and examples of some bismuth perovskite and halide absorbers with 0D, 1D, 2D, and 3D octahedral networks.

FIG. 3.

Illustrations and examples of some bismuth perovskite and halide absorbers with 0D, 1D, 2D, and 3D octahedral networks.

Close modal

The CH3NH3PbI3 perovskite structure contains a 3D network of PbI6 octahedra that share corners in all three octahedral directions.46 Bi3+ cannot directly replace Pb2+ in this 3D-perovskite structure due to its higher charge.47 Charge neutrality forces the bismuth counterpart (CH3NH3)3Bi2I9 into a 0D structure with face-sharing octahedra of (Bi2I9)3 dimers that are separated by (CH3NH3)+ ions.48 The lack of a network between the (Bi2I9)3 dimers has been blamed for lower carrier mobilities and a larger bandgap of (CH3NH3)3Bi2I9 and Cs3Bi2I9.42 Double perovskites, which contain two different cations, offer one route to forming higher dimensional bismuth perovskites.34,49 For example, Cs2AgBiBr6 has a 3D octahedral network and has demonstrated power-conversion efficiencies up to 2.43% as discussed previously.

Although the efficiencies of bismuth perovskites and halides are currently modest, their demonstrated stability in humid air and their avoidance of toxic lead encourage further investigation. As many of these materials have yet to be extensively investigated, there remains considerable hope that the efficiencies can be significantly improved.

Bi2S3 is an n-type semicondcutor,50 with a high absorption coefficient (in the range of 105 cm−1), and an absorption onset in the infrared. The reported bandgaps lie in the adequate region for photovoltaics and vary from 1.3 eV, for example in the bulk, to 1.7 eV, thanks to quantum confinement effects and/or stoichiometry variations.51,52

The first reports demonstrating the use of colloidal Bi2S3 nanocrystals (NCs) as an absorbing material in bilayer heterojunction solar cells employed as a p-type material either PbS quantum dots that had been exposed to ambient conditions (QDs)53 or a conductive organic polymer like poly(3-hexylthiophene) (P3HT).54 Both studies succeeded at demonstrating the Bi2S3 contribution to photocurrent by fabricating solar cells with ultra-small PbS QDs, with an absorption onset at 800 nm, or P3HT, with an absorption onset at 700 nm. The maximum efficiency obtained, 1.61% with PbS and 0.46% with P3HT, was modest but promising for a new material. Moreover, these studies already pointed to some factors limiting efficiency like high doping of the Bi2S3 films, poor electron mobility, and high recombination, most likely due to surface electron traps.

Efficiency was improved close to 5% by blending Bi2S3 NCs and PbS QDs to form a bulk nano-heterojunction structure.55 The intimate mixing of the n- and p-type nanocrystalline semiconductors favors carrier transfer between materials, separating them more efficiently and reducing recombination. All these result in a prolonged lifetime and a threefold improvement in the short-circuit current. Similarly, when Bi2S3 NCs were blended with P3HT, or thiol-functionalized P3HT, an improvement in the short-circuit current provided an efficiency up to 1%.56,57 Exploiting this same idea, performance was further improved when P3HT was blended with Bi2S3/Au Schottky diodes to fabricate hybrid bulk-heterojunction solar cells. The semiconductor-metal nanostructure augmented separation of charge carriers upon photogeneration and yielded an efficiency of 2%.58 

Modifications in the solar cell structure is not the only strategy possible to improve efficiency. Changes in the synthesis enabled the production of Bi2S3 nanocrystals with different shapes and sizes, from 14 × 19 nm nanorods to spherical particles with a diameter of 2.6 nm; their performance in P3HT–Bi2S3 bilayer solar cells was studied.52 The efficiency of the devices fabricated with the smaller nanocrystals was similar to analogous bilayer structures (0.43%),54 but they showed an improved open-circuit voltage (Voc), indicating reduced recombination, most likely due to a better surface passivation and a reduction of traps.

The ns2 lone pair of Bi3+ tends to be stereochemically active and produce highly anisotropic crystalline structures.59,60 Indeed, Bi2S3 crystallizes in the orthorhombic system where tightly connected chains create layers that are connected through weak van der Waals forces that stack to form the final structure. This asymmetric structure produces preferential directions for electronic transport, enhanced along the c axis.61 

In the examples mentioned above, nanocrystals are randomly oriented in the films (Fig. 4). Therefore, some of the NCs will be properly placed for conducting charges, while others will not, limiting charge transport and device efficiency. Considering this, strategies to create aligned Bi2S3 1D structures like nanowires (NWs) or nanoribbons have been developed.62,63 Bi2S3 NWs were vertically grown on a TiO2-coated substrate, and spiro-OMeTAD was used as a hole transport layer. This structure suffered from high interfacial charge recombination, so a Bi2S3/Ag2S core-shell structure was developed and an efficiency of 2.5% was achieved.62 In a different approach, Bi2S3 NWs were engineered to form percolated networks, where the interconnected nanowires provide a continuous path for electron transport. These porous networks were infiltrated with P3HT, forming a bulk-heterojunction and achieving a maximum efficiency of 3.3%.63 

FIG. 4.

Illustrations and examples of some bismuth-based sulfides used in different solar cell structures.

FIG. 4.

Illustrations and examples of some bismuth-based sulfides used in different solar cell structures.

Close modal

Despite their potential application in many different areas, not only photovoltaics, the group of copper-bismuth sulfides has been scarcely explored. This is quite surprising, considering that all the elements are abundant and non-toxic and that analogous materials like CuInS2, the copper-antimony family, and the copper-tin family have been the focus of several studies.64–66 

Ternary Cu–Bi sulfides can show different stoichiometries and different properties. The three most studied compounds—CuBiS2, Cu3BiS3, and Cu4Bi4S9—crystallize in an orthorhombic structure and exhibit high absorption coefficients (104–105 cm−1) and p-type character. These materials differ in their absorption onsets, and the reported bandgaps lie between 1.5 and 2.62 eV for CuBiS2, 1.2 and 1.84 eV for Cu3BiS3, and 0.88–1.14 eV for Cu4Bi4S9.64,67

CuBiS2 (emplectite) has been mostly studied from the theoretical point of view;68–70 there are very few reports demonstrating the use of this material in photovoltaic solar cells (Fig. 4). In one experimental study, CuBiS2 nanoparticles were deposited on TiO2 using a chemical bath deposition method and used as sensitizers in dye sensitized solar cells (DSSCs) with a maximum efficiency of 0.62%.71 More recently, CuBiS2 was used a p-type material in planar heterojunction solar cells, sandwiched between TiO2 and P3HT acting as hole and electron blocking layers, respectively. A maximum efficiency of 0.68% was achieved after doping the CuBiS2 active layer with indium chloride.72 

According to theoretical calculations, CuBiS2 presents an indirect bandgap at 1.4–1.7 eV, however, the difference between indirect and direct bandgap energies is only 0.1–0.3 eV due to the rather flat lowest conduction band (CB); therefore, the direct transition is still accessible, and indeed the material exhibits high absorption coefficients (0.93–1.5 × 105 cm−1).70,73 Interestingly, the CuBiS2 density of states indicates that the formation of hole carriers will involve oxidation of Cu(i)–Cu(ii). This may have implications for hole transport in CuBiS2 and suggests that some Cu deficiencies may be beneficial, like in analogous photovoltaic materials such as CZTS.69 

Cu3BiS3 (wittichenite) is of interest because it contains abundant non-toxic elements and presents p-type conductivity, a high absorption coefficient (1 × 105 cm−1), and reported experimental bandgaps between 1.2 and 1.84 eV.64,74 Again, theoretical studies have calculated that the material exhibits a fundamental indirect bandgap at 1.5–1.7 eV, but the direct gap is close in energy (1.6–1.8 eV), showing promise as a solar absorber.75 Cu3BiS3 nanocrystals were synthesized, and the corresponding nanocrystal-film showed a clear photoresponse in I–V measurements;76 however, there are few studies reporting its incorporation in photovoltaic devices.

The first report on the efficiency of Cu3BiS3 employed nanosheets, obtained by a solvothermal route, that were used as a sensitizer on TiO2 nanorods. The reported energy conversion efficiency of the Cu3BiS3/TiO2 thin-film solar cell was a promising 1.281%.77 

More recently, a Cu3BiS3 thin film was deposited by a dimethyl sulfoxide (DMSO)-based solution coating process and subsequent annealing in a nitrogen atmosphere. A solar cell combining wittichenite as a p-type material and CdS as an n-type semiconductor was fabricated, achieving a power conversion efficiency of 0.17%.78 The authors attribute the quite low short-circuit current to an inadequate band alignment between the two semiconductors, indicating that improved efficiencies could be expected with a more appropriate n-type material.

To date the highest efficiencies using copper-bismuth sulfides have been obtained employing Cu4Bi4S9 (Fig. 4). As commented in Sec. III, the anisotropic orthorhombic structure can be responsible for the limited efficiency. Indeed, most of the studies on Cu4Bi4S9 focus on the preparation of 1D structures (nanoribbons or nanobelts) and their incorporation in solar cells.

High conversion efficiencies, over 6%, have been observed for Cu4Bi4S9 when nanoribbons obtained through a solvothermal method were used as p-type materials combined with n-type oxides, like In2O3 or TiO2, and using In2S3 as a buffer layer.79 Later the efficiency was increased to 6.4% by using n-type ZnO nanowires coated with an In2O3 buffer layer and Cu4Bi4S9 as the absorber and hole transport layer.80 More recently, efficiencies have approached 7% for bulk heterojunctions of Cu4Bi4S9 and graphene oxide deposited on a α-Fe2O3 layer.81 

In contrast to Bi2S3 and the copper-bismuth sulfide family, AgBiS2 crystallizes in symmetrical crystalline structures: hexagonal at low temperatures, and a pseudo-cubic rock salt structure at high temperatures and as nanocrystals (Fig. 4). This fact enables possible omnidirectional transport of carriers.82 Initial reports described AgBiS2 as an n-type material with a favorable direct bandgap between 1.0 and 1.4 eV.83,84

As a first approach, AgBiS2 nanocrystals were used as a sensitizer or counter-electrode in dye sensitized solar cells (DSSCs). The first AgBiS2 nanoparticles were synthesized on TiO2 using a sequential ionic layer adsorption reaction (SILAR) process. The liquid-junction semiconductor-sensitized solar cells fabricated with the photoanode showed a power conversion efficiency of 0.53% under 1 sun, and of 0.76% at a reduced light intensity of 0.148 sun.85 More recently, a TiO2 nanorod array was decorated with AgBiS2 nanocrystals electrochemically deposited, reaching a maximum efficiency of 0.95%.86 Using a related approach, AgBiS2 nanocrystals were synthesized by a solvothermal process and used as a contra-electrode in a polysulfide electrolyte DSSC. The overall power conversion efficiency was 2.09%, higher than the reference device using platinum as a counter-electrode (1.73%).87 

Finally, AgBiS2 nanocrystals were used as an active layer in solution-processed solar cells. The effect of treating the nanocrystals with different ligands was studied, observing that treatment with tetramethylammonium iodide generates films with intrinsic behavior. In the optimized device (6.3% certified efficiency), the AgBiS2 layer was sandwiched between an electron transport/hole blocking substrate made from a thin layer of ZnO and an ultrathin electron blocking/hole transport PTB7 polymer layer. The promise of AgBiS2 as a lead-free photovoltaic material is highlighted by a high short-circuit current of 22 mA/cm2, with an active layer of only 35-nm thick. Furthermore, all the materials in the solar cell are non-toxic, and excluding the ITO layer and the metal contacts, all the layers are solution processed under ambient conditions and low temperatures.88 

This study already identified some of the factor limiting performance: inefficient extraction of carriers at higher light intensities (especially for devices thicker than 50 nm), trap-assisted recombination processes, poor carrier transport, and incomplete extraction before recombination. Some possible solutions could be improvements in the synthesis, better nanocrystal passivation (ligand/surface treatments), introduction of light-trapping schemes, or nanostructuring of the active layer.

In another study, AgBiS2 nanocrystals were obtained using an improved amine-based synthesis route.89 Films fabricated with these nanocrystals showed improved carrier mobility, reduced carrier concentration, and improved photosensitivity, as compared to nanocrystals synthesized with oleic acid. This allowed us to fabricate solar cells with thicker active layers leading to improved efficiencies. Indeed, the optimal thickness of a champion solar cell was increased from 35 nm to 65 nm yielding a maximum efficiency of 4.3% using P3HT as a hole transport layer. This value compares well with previously reported values of 3.9% for a 35 nm thick device88 and 2.6% for a 65 nm thick device fabricated with the oleic acid synthesis.

Recently, AgBiS2 films were fabricated by spray pyrolysis.90 In the optimized device, a 60–70 nm AgBiS2 layer was integrated between a ZnO electron transport layer and a MoO3-modified spiro-OMeTAD hole transport layer. Devices exhibited a maximum power-conversion efficiency of 1.5% with an aperture of 0.16 cm2 and 1.2% for a larger area of 1 cm2. Again, significant short-circuit currents were observed (18 mA/cm2). When fabricating devices with active layers thicker than 70 nm, an enhancement in open-circuit voltage and field factor was observed but short-circuit current dramatically decreased. Interestingly, solar cells showed no degradation when stored for 1 month under diffuse light under ambient conditions at approximately 50% relative humidity and 24 ± 2 °C.

These studies point to the fact that with further work on AgBiS2 efficiencies approaching those of the toxic counterparts could be achieved.

In this Research perspective, we have summarized some of the bismuth-based materials that could serve as alternatives to the best performing, but highly toxic, materials used in perovskite or quantum dot solar cells. These materials fulfill the requirements of being composed of abundant and non-toxic elements. Moreover, they exhibit adequate properties for photovoltaics like high absorption coefficients and suitable bandgaps, plus additional attractive characteristics in terms of robustness and stability.

Apparently, one of the main limiting factors of these bismuth-based materials seems to be the anisotropy of the crystalline structures, forced by the presence of the ns2 lone pair. This points to two routes to improve efficiency. On the one hand, take advantage of the anisotropic, low-dimensional crystalline structures by optimizing carrier transport through the growth of properly aligned and connected 1D structures (nanorods, nanowires, nanoribbons, etc.). On the other hand, identify more ternary compounds like AgBiS2 which crystallize in symmetrical crystalline structures or perovskites that form higher dimensional octahedral networks. In this sense, a study explaining the difference in crystalline structures for materials with alternate cations would be extremely interesting, especially if it could predict the crystalline structure of these compositions.

Even more, these materials have surprisingly received little attention. As a result, strategies that have proven successful in analogous systems—including alloying, doping, surface treatments with ligands, adequate alignment of semiconductors, or formation of bulk heterojunctions—remain largely unexplored.91–94 These areas would be of interest in future studies and could further improve the efficiency of devices based on these abundant, non-toxic materials.

Finally, bismuth-based materials find applications in other optoelectronic devices like photodetectors,95–97 photocatalytic processes like solar-driven hydrogen production or light degradation of pollutants,98–100 other clean-energy devices like batteries or thermoelectric devices,82,101–104 and as theragnostic agents (bioimaging + photothermal therapy).105–107 All these areas will benefit from a better understanding and improvement of the properties of bismuth-based materials.

M.B. thanks the Energy and Environmental Theme at the Cardiff School of Engineering for financial support to attend the EMRS conference on May 2017.

1.
See https://www.nrel.gov/pv/assets/images/efficiency-chart.png for efficiency-chart.png (2219 × 1229) (2017); accessed 19 January 2018.
2.
C.
Wadia
,
A. P.
Alivisatos
, and
D. M.
Kammen
, “
Materials availability expands the opportunity for large-scale photovoltaics deployment
,”
Environ. Sci. Technol.
43
,
2072
2077
(
2009
).
3.
P. C. K.
Vesborg
and
T. F.
Jaramillo
, “
Addressing the terawatt challenge: Scalability in the supply of chemical elements for renewable energy
,”
RSC Adv.
2
,
7933
(
2012
).
4.
M.
Liu
 et al, “
Hybrid organic-inorganic inks flatten the energy landscape in colloidal quantum dot solids
,”
Nat. Mater.
16
,
258
263
(
2017
).
5.
E. M.
Sanehira
 et al, “
Enhanced mobility CsPbI3 quantum dot arrays for record-efficiency, high-voltage photovoltaic cells
,”
Sci. Adv.
3
,
eaao4204
(
2017
).
6.
W. S.
Yang
 et al, “
Iodide management in formamidinium-lead-halide–based perovskite layers for efficient solar cells
,”
Science
356
,
1376
1379
(
2017
).
7.
WHO
,
Ten Chemicals of Major Public Health Concern
(
WHO
,
2016
).
8.
A.
Babayigit
,
H.-G.
Boyen
, and
B.
Conings
, “
Environment versus sustainable energy: The case of lead halide perovskite-based solar cells
,”
MRS Energy Sustainability
5
,
E1
(
2018
).
9.
J. L.
DiMeglio
and
J.
Rosenthal
, “
Selective conversion of CO2 to CO with high efficiency using an inexpensive bismuth-based electrocatalyst
,”
J. Am. Chem. Soc.
135
,
8798
8801
(
2013
).
10.
R.
Mohan
, “
Green bismuth
,”
Nat. Chem.
2
,
336
(
2010
).
11.
A. M.
Ganose
,
C. N.
Savory
, and
D. O.
Scanlon
, “
Beyond methylammonium lead iodide: Prospects for the emergent field of ns2 containing solar absorbers
,”
Chem. Commun.
53
,
20
44
(
2017
).
12.
R. E.
Brandt
,
V.
Stevanovi
,
D. S.
Ginley
, and
T.
Buonassisi
, “
Identifying defect-tolerant semiconductors with high minority-carrier lifetimes: Beyond hybrid lead halide perovskites
,”
MRS Commun.
5
,
265
275
(
2015
).
13.
M. A.
Green
,
Y.
Jiang
,
A. M.
Soufiani
, and
A.
Ho-Baillie
, “
Optical properties of photovoltaic organic-inorganic lead halide perovskites
,”
J. Phys. Chem. Lett.
6
,
4774
4785
(
2015
).
14.
S. D.
Stranks
 et al, “
Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber
,”
Science
342
,
341
344
(
2013
).
15.
G.
Han
 et al, “
Towards high efficiency thin film solar cells
,”
Prog. Mater. Sci.
87
,
246
291
(
2017
).
16.
M.
Lyu
,
J. H.
Yun
,
P.
Chen
,
M.
Hao
, and
L.
Wang
, “
Addressing toxicity of Lead: Progress and applications of low-toxic metal halide perovskites and their derivatives
,”
Adv. Energy Mater.
7
,
1602512
(
2017
).
17.
M.
Konstantakou
and
T.
Stergiopoulos
, “
A critical review on tin halide perovskite solar cells
,”
J. Mater. Chem. A
5
,
11518
11549
(
2017
).
18.
K. M.
Boopathi
 et al, “
Solution-processable antimony-based light-absorbing materials beyond lead halide perovskites
,”
J. Mater. Chem. A
5
,
20843
20850
(
2017
).
19.
C.
Zhang
,
L.
Gao
,
S.
Hayase
, and
T.
Ma
, “
Current advancements in material research and techniques focusing on lead-free perovskite solar cells
,”
Chem. Lett.
46
,
1276
1284
(
2017
).
20.
O.
Lindqvist
,
G.
Johansson
,
F.
Sandberg
, and
T.
Norin
, “
The crystal structure of cesium bismuth iodide, Cs3Bi2I9
,”
Acta Chem. Scand.
22
,
2943
2952
(
1968
).
21.
B. W.
Park
 et al, “
Bismuth based hybrid perovskites A3Bi2I9 (A: Methylammonium or cesium) for solar cell application
,”
Adv. Mater.
27
,
6806
6813
(
2015
).
22.
Z.
Yang
 et al, “
Unraveling the exciton binding energy and the dielectric constant in single-crystal methylammonium lead triiodide perovskite
,”
J. Phys. Chem. Lett.
8
,
1851
1855
(
2017
).
23.
A.
Miyata
 et al, “
Direct measurement of the exciton binding energy and effective masses for charge carriers in organic-inorganic tri-halide perovskites
,”
Nat. Phys.
11
,
582
587
(
2015
).
24.
H.-H.
Fang
 et al, “
Photoexcitation dynamics in solution-processed formamidinium lead iodide perovskite thin films for solar cell applications
,”
Light: Sci. Appl.
5
,
e16056
(
2015
).
25.
B.
Ghosh
 et al, “
Poor photovoltaic performance of Cs3Bi2I9: An insight through first-principles calculations
,”
J. Phys. Chem. C
121
,
17062
17067
(
2017
).
26.
T.
Singh
,
A.
Kulkarni
,
M.
Ikegami
, and
T.
Miyasaka
, “
Effect of electron transporting layer on bismuth-based lead-free perovskite (CH3NH3)3Bi2I9 for photovoltaic applications
,”
ACS Appl. Mater. Interfaces
8
,
14542
14547
(
2016
).
27.
M.
Lyu
 et al, “
Organic–inorganic bismuth (III)-based material: A lead-free, air-stable and solution-processable light-absorber beyond organolead perovskites
,”
Nano Res.
9
,
692
702
(
2016
).
28.
R. L. Z.
Hoye
 et al, “
Methylammonium bismuth iodide as a lead-free, stable hybrid organic-inorganic solar absorber
,”
Chem. - Eur. J.
22
,
2605
2610
(
2016
).
29.
Q.
Li
 et al, “
High pressure band gap engineering in lead-free Cs2AgBiBr6 double perovskite
,”
Angew. Chem., Int. Ed.
56
,
15969
15973
(
2017
).
30.
M.
Vigneshwaran
 et al, “
Facile synthesis and characterization of sulfur doped low bandgap bismuth based perovskites by soluble precursor route
,”
Chem. Mater.
28
,
6436
6440
(
2016
).
31.
K.-H.
Hong
,
J.
Kim
,
L.
Debbichi
,
H.
Kim
, and
S. H.
Im
, “
Band gap engineering of Cs3Bi2I9 perovskites with trivalent atoms using a dual metal cation
,”
J. Phys. Chem. C
121
,
969
974
(
2017
).
32.
C.
Ran
 et al, “
Construction of compact methylammonium bismuth iodide film promoting lead-free inverted planar heterojunction organohalide solar cells with open-circuit voltage over 0.8 V
,”
J. Phys. Chem. Lett.
8
,
394
400
(
2017
).
33.
Z.
Zhang
 et al, “
High-quality (CH3NH3)3Bi2I9 film-based solar cells: Pushing efficiency up to 1.64%
,”
J. Phys. Chem. Lett.
8
,
4300
4307
(
2017
).
34.
E.
Greul
,
M. L.
Petrus
,
A.
Binek
,
P.
Docampo
, and
T.
Bein
, “
Highly stable, phase pure Cs2AgBiBr6 double perovskite thin films for optoelectronic applications
,”
J. Mater. Chem. A
5
,
19972
19981
(
2017
).
35.
W.
Gao
 et al, “
High quality Cs2AgBiBr6 double perovskite film for lead-free inverted planar heterojunction solar cells with 2.2% efficiency
,”
ChemPhysChem
(published online,
2018
).
36.
Y.
Kim
 et al, “
Pure cubic-phase hybrid iodobismuthates AgBi2I7 for thin-film photovoltaics
,”
Angew. Chem., Int. Ed.
55
,
9586
9590
(
2016
).
37.
A. J.
Lehner
 et al, “
Electronic structure and photovoltaic application of BiI3
,”
Appl. Phys. Lett.
107
,
131109
(
2015
).
38.
R. E.
Brandt
 et al, “
Investigation of bismuth triiodide (BiI3) for photovoltaic applications
,”
J. Phys. Chem. Lett.
6
,
4297
4302
(
2015
).
39.
U. H.
Hamdeh
 et al, “
Solution-processed BiI3 thin films for photovoltaic applications: Improved carrier collection via solvent annealing
,”
Chem. Mater.
28
,
6567
6574
(
2016
).
40.
X.
Huang
,
S.
Huang
,
P.
Biswas
, and
R.
Mishra
, “
Band gap insensitivity to large chemical pressures in ternary bismuth iodides for photovoltaic applications
,”
J. Phys. Chem. C
120
,
28924
28932
(
2016
).
41.
T.
Li
 et al, “
Lead-free pseudo-three-dimensional organic–inorganic iodobismuthates for photovoltaic applications
,”
Sustainable Energy Fuels
1
,
308
316
(
2017
).
42.
A. J.
Lehner
 et al, “
Crystal and electronic structures of complex bismuth iodides A3Bi2I9(A = K, Rb, Cs) related to perovskite: Aiding the rational design of photovoltaics
,”
Chem. Mater.
27
,
7137
7148
(
2015
).
43.
C. W. M.
Timmermans
,
S. O.
Cholakh
,
R. L.
van der Woude
, and
G.
Blasse
, “
Some optical and electrical measurements on Cs3Bi2Br9 single crystals
,”
Phys. Status Solidi B
115
,
267
271
(
1983
).
44.
M. I.
Saidaminov
,
O. F.
Mohammed
, and
O. M.
Bakr
, “
Low-dimensional-networked metal halide perovskites: The next big thing
,”
ACS Energy Lett.
2
,
889
896
(
2017
).
45.
I. C.
Smith
,
E. T.
Hoke
,
D.
Solis-Ibarra
,
M. D.
McGehee
, and
H. I. A.
Karunadasa
, “
Layered hybrid perovskite solar-cell absorber with enhanced moisture stability
,”
Angew. Chem., Int. Ed.
53
,
11232
11235
(
2014
).
46.
J. H.
Lee
,
J. H.
Lee
,
E. H.
Kong
, and
H. M.
Jang
, “
The nature of hydrogen-bonding interaction in the prototypic hybrid halide perovskite, tetragonal CH3NH3PbI3
,”
Sci. Rep.
6
,
1
12
(
2016
).
47.
S.
Sun
 et al, “
Synthesis, crystal structure, and properties of a perovskite-related bismuth phase, (NH4)3Bi2I9
,”
APL Mater.
4
,
031101
(
2016
).
48.
K.
Eckhardt
 et al, “
Crystallographic insights into (CH3NH3)3(Bi2I9): A new lead-free hybrid organic–inorganic material as a potential absorber for photovoltaics
,”
Chem. Commun.
52
,
3058
3060
(
2016
).
49.
E. T.
McClure
,
M. R.
Ball
,
W.
Windl
, and
P. M.
Woodward
, “
Cs2AgBiX6 (X = Br, Cl): New visible light absorbing, lead-free halide perovskite semiconductors
,”
Chem. Mater.
28
,
1348
1354
(
2016
).
50.
G.
Konstantatos
,
L.
Levina
,
J.
Tang
, and
E. H.
Sargent
, “
Sensitive solution-processed Bi2S3 nanocrystalline photodetectors
,”
Nano Lett.
8
,
4002
4006
(
2008
).
51.
L.
Cademartiri
 et al, “
Large-scale synthesis of ultrathin Bi2S3 necklace nanowires
,”
Angew. Chem., Int. Ed.
47
,
3814
3817
(
2008
).
52.
M.
Bernechea
,
Y.
Cao
, and
G.
Konstantatos
, “
Size and bandgap tunability in Bi2S3 colloidal nanocrystals and its effect in solution processed solar cells
,”
J. Mater. Chem. A
3
,
20642
20648
(
2015
).
53.
A. K.
Rath
,
M.
Bernechea
,
L.
Martinez
, and
G.
Konstantatos
, “
Solution-processed heterojunction solar cells based on p-type PbS quantum dots and n-type Bi2S3 nanocrystals
,”
Adv. Mater.
23
,
3712
3717
(
2011
).
54.
L.
Martinez
,
M.
Bernechea
,
F. P. G.
de Arquer
, and
G.
Konstantatos
, “
Near IR-sensitive, non-toxic, polymer/nanocrystal solar cells employing Bi2S3 as the electron acceptor
,”
Adv. Energy Mater.
1
,
1029
1035
(
2011
).
55.
A. K.
Rath
 et al, “
Solution-processed inorganic bulk nano-heterojunctions and their application to solar cells
,”
Nat. Photonics
6
,
529
534
(
2012
).
56.
L.
Martinez
 et al, “
Hybrid solution-processed bulk heterojunction solar cells based on bismuth sulfide nanocrystals
,”
Phys. Chem. Chem. Phys.
15
,
5482
5487
(
2013
).
57.
L.
Martinez
 et al, “
Improved electronic coupling in hybrid organic–inorganic nanocomposites employing thiol-functionalized P3HT and bismuth sulfide
,”
Nanoscale
6
,
10018
(
2014
).
58.
S. K.
Saha
and
A. J.
Pal
, “
Schottky diodes between Bi2S3 nanorods and metal nanoparticles in a polymer matrix as hybrid bulk-heterojunction solar cells
,”
J. Appl. Phys.
118
,
014503
(
2015
).
59.
A.
Walsh
,
D. J.
Payne
,
R. G.
Egdell
, and
G. W.
Watson
, “
Stereochemistry of post-transition metal oxides: Revision of the classical lone pair model
,”
Chem. Soc. Rev.
40
,
4455
(
2011
).
60.
L. A.
Olsen
,
J.
Lpez-Solano
,
A.
Garca
,
T.
Balić-Unić
, and
E.
Makovicky
, “
Dependence of the lone pair of bismuth on coordination environment and pressure: An ab initio study on Cu4Bi5S10 and Bi2S3
,”
J. Solid State Chem.
183
,
2133
2143
(
2010
).
61.
H.
Bao
 et al, “
Synthesis of a highly ordered single-crystalline Bi2S3 nanowire array and its metal/semiconductor/metal back-to-back Schottky diode
,”
Small
4
,
1125
1129
(
2008
).
62.
Y.
Cao
 et al, “
Solution processed bismuth sulfide nanowire array core/silver sulfide shell solar cells
,”
Chem. Mater.
27
,
3700
3706
(
2015
).
63.
L.
Whittaker-Brooks
 et al, “
Bi2S3 nanowire networks as electron acceptor layers in solution-processed hybrid solar cells
,”
J. Mater. Chem. C
3
,
2686
2692
(
2015
).
64.
C.
Coughlan
 et al, “
Compound copper chalcogenide nanocrystals
,”
Chem. Rev.
117
,
5865
6109
(
2017
).
65.
H.
Fu
, “
Environment-friendly and earth-abundant colloidal chalcogenide nanocrystals for photovoltaic applications
,”
J. Mater. Chem. C
6
,
414
445
(
2018
).
66.
A.
Zakutayev
, “
Brief review of emerging photovoltaic absorbers
,”
Curr. Opin. Green Sustainable Chem.
4
,
8
15
(
2017
).
67.
S. G.
Deshmukh
and
V.
Kheraj
, “
A comprehensive review on synthesis and characterizations of Cu3BiS3 thin films for solar photovoltaics
,”
Nanotechnol. Environ. Eng.
2
,
15
(
2017
).
68.
D. J.
Temple
,
A. B.
Kehoe
,
J. P.
Allen
,
G. W.
Watson
, and
D. O.
Scanlon
, “
Geometry, electronic structure, and bonding in CuMCh2(M = Sb, Bi; Ch = S, Se): Alternative solar cell absorber materials?
,”
J. Phys. Chem. C
116
,
7334
7340
(
2012
).
69.
J. T. R.
Dufton
 et al, “
Structural and electronic properties of CuSbS2 and CuBiS2: Potential absorber materials for thin-film solar cells
,”
Phys. Chem. Chem. Phys.
14
,
7229
(
2012
).
70.
M.
Kumar
and
C.
Persson
, “
CuSbS2 and CuBiS2 as potential absorber materials for thin-film solar cells
,”
J. Renewable Sustainable Energy
5
,
031616
(
2013
).
71.
N.
Suriyawong
,
B.
Aragaw
,
J. B.
Shi
, and
M. W.
Lee
, “
Ternary CuBiS2 nanoparticles as a sensitizer for quantum dot solar cells
,”
J. Colloid Interface Sci.
473
,
60
65
(
2016
).
72.
J.-J.
Wang
,
M. Z.
Akgul
,
Y.
Bi
,
S.
Christodoulou
, and
G.
Konstantatos
, “
Low-temperature colloidal synthesis of CuBiS2 nanocrystals for optoelectronic devices
,”
J. Mater. Chem. A
5
,
24621
24625
(
2017
).
73.
M.
Kumar
and
C.
Persson
, “
Cu(Sb,Bi)(S,Se)2 as indium-free absorber material with high optical efficiency
,”
Energy Procedia
44
,
176
183
(
2014
).
74.
N. J.
Gerein
and
J. A.
Haber
, “
One-step synthesis and optical and electrical properties of thin film Cu3BiS3 for use as a solar absorber in photovoltaic devices
,”
Chem. Mater.
18
,
6297
6302
(
2006
).
75.
M.
Kumar
and
C.
Persson
, “
Cu3BiS3 as a potential photovoltaic absorber with high optical efficiency
,”
Appl. Phys. Lett.
102
,
062109
(
2013
).
76.
C.
Yan
 et al, “
Colloidal synthesis and characterizations of wittichenite copper bismuth sulphide nanocrystals
,”
Nanoscale
5
,
1789
(
2013
).
77.
J.
Yin
and
J.
Jia
, “
Synthesis of Cu3BiS3 nanosheet films on TiO2 nanorod arrays by a solvothermal route and their photoelectrochemical characteristics
,”
CrystEngComm
16
,
2795
(
2014
).
78.
J.
Li
 et al, “
One–step synthesis of Cu3BiS3 thin films by a dimethyl sulfoxide (DMSO)–based solution coating process for solar cell application
,”
Sol. Energy Mater. Sol. Cells
174
,
593
598
(
2018
).
79.
X.
Liu
,
H.
Zheng
,
J.
Zhang
,
Y.
Xiao
, and
Z.
Wang
, “
Photoelectric properties and charge dynamics for a set of solid state solar cells with Cu4Bi4S9 as the absorber layer
,”
J. Mater. Chem. A
1
,
10703
(
2013
).
80.
X.
Liu
,
S.
Wang
,
J.
Zhang
,
J.
Zhang
, and
Y.
Gu
, “
Photoelectric properties and charge dynamics in ZnO nanowires/Cu4Bi4S9 and ZnO nanowires/In2O3/Cu4Bi4S9 heterostructures
,”
J. Appl. Phys.
116
,
245101
(
2014
).
81.
S.
Wang
,
X. Y.
Liu
, and
Y. Z.
Gu
, “
Excellent photoelectric properties and charge dynamics of two types of bulk heterojunction solar cells
,”
Mater. Lett.
166
,
251
254
(
2016
).
82.
S. N.
Guin
and
K.
Biswas
, “
Cation disorder and bond anharmonicity optimize the thermoelectric properties in kinetically stabilized rocksalt AgBiS2 nanocrystals
,”
Chem. Mater.
25
,
3225
3231
(
2013
).
83.
B.
Pejova
,
D.
Nesheva
,
Z.
Aneva
, and
A.
Petrova
, “
Photoconductivity and relaxation dynamics in sonochemically synthesized assemblies of AgBiS2 quantum dots
,”
J. Phys. Chem. C
115
,
37
46
(
2011
).
84.
B.
Pejova
,
I.
Grozdanov
,
D.
Nesheva
, and
A.
Petrova
, “
Size-dependent properties of sonochemically synthesized three-dimensional arrays of close-packed semiconducting AgBiS2 quantum dots
,”
Chem. Mater.
20
,
2551
2565
(
2008
).
85.
P. C.
Huang
,
W. C.
Yang
, and
M. W.
Lee
, “
AgBiS2 semiconductor-sensitized solar cells
,”
J. Phys. Chem. C
117
,
18308
18314
(
2013
).
86.
S.
Zhou
 et al, “
Preparation and photovoltaic properties of ternary AgBiS2 quantum dots sensitized TiO2 nanorods photoanodes by electrochemical atomic layer deposition
,”
J. Electrochem. Soc.
163
,
D63
D67
(
2016
).
87.
N.
Liang
 et al, “
Homogenously hexagonal prismatic AgBiS2 nanocrystals: controlled synthesis and application in quantum dot-sensitized solar cells
,”
CrystEngComm
17
,
1902
1905
(
2015
).
88.
M.
Bernechea
 et al, “
Solution-processed solar cells based on environmentally friendly AgBiS2 nanocrystals
,”
Nat. Photonics
10
,
521
525
(
2016
).
89.
L.
Hu
 et al, “
Enhanced optoelectronic performance in AgBiS2 nanocrystals from an improved amine-based synthesis route
,”
J. Mater. Chem. C
6
,
731
737
(
2018
).
90.
N.
Pai
 et al, “
Spray deposition of AgBiS2 and Cu3BiS3 thin films for photovoltaic applications
,”
J. Mater. Chem. C
6
,
2483
2494
(
2018
).
91.
G. H.
Carey
 et al, “
Colloidal quantum dot solar cells
,”
Chem. Rev.
115
,
12732
12763
(
2015
).
92.
F.
Liu
 et al, “
Colloidal synthesis of air-stable alloyed CsSn1-xPbxI3 perovskite nanocrystals for use in solar cells
,”
J. Am. Chem. Soc.
139
,
16708
(
2017
).
93.
M.
Yuan
,
M.
Liu
, and
E. H.
Sargent
, “
Colloidal quantum dot solids for solution-processed solar cells
,”
Nat. Energy
1
,
16016
(
2016
).
94.
C. R.
Kagan
and
C. B.
Murray
, “
Charge transport in strongly coupled quantum dot solids
,”
Nat. Nanotechnol.
10
,
1013
1026
(
2015
).
95.
J.
Chao
 et al, “
Large-scale synthesis of Bi2S3 nanorods and nanoflowers for flexible near-infrared laser detectors and visible light photodetectors
,”
Mater. Res. Bull.
98
,
194
199
(
2018
).
96.
W.
Huang
 et al, “
Facile fabrication and characterization of two-dimensional bismuth(III) sulfide nanosheets for high-performance photodetector applications under ambient conditions
,”
Nanoscale
10
,
2404
2412
(
2018
).
97.
H.
Yu
 et al, “
Scalable colloidal synthesis of uniform Bi2S3 nanorods as sensitive materials for visible-light photodetectors
,”
CrystEngComm
19
,
727
733
(
2017
).
98.
A.
Sarkar
 et al, “
Enhanced photocatalytic performance of morphologically tuned Bi2S3 NPs in the degradation of organic pollutants under visible light irradiation
,”
J. Colloid Interface Sci.
483
,
49
59
(
2016
).
99.
X.
Meng
and
Z.
Zhang
, “
Bismuth-based photocatalytic semiconductors: Introduction, challenges and possible approaches
,”
J. Mol. Catal. A: Chem.
423
,
533
549
(
2016
).
100.
H.
Abdullah
and
D.-H.
Kuo
, “
Photocatalytic performance of Ag and CuBiS2 nanoparticle-coated SiO2@TiO2 composite sphere under visible and ultraviolet light irradiation for azo dye degradation with the assistance of numerous nano p–n diodes
,”
J. Phys. Chem. C
119
,
13632
13641
(
2015
).
101.
J.
Ni
 et al, “
Strongly coupled Bi2S3@CNT hybrids for robust lithium storage
,”
Adv. Energy Mater.
4
,
1400798
(
2014
).
102.
G.
Nie
,
X.
Lu
,
J.
Lei
,
L.
Yang
, and
C.
Wang
, “
Facile and controlled synthesis of bismuth sulfide nanorods-reduced graphene oxide composites with enhanced supercapacitor performance
,”
Electrochim. Acta
154
,
24
30
(
2015
).
103.
W.
Yang
,
H.
Wang
,
T.
Liu
, and
L.
Gao
, “
A Bi2S3@CNT nanocomposite as anode material for sodium ion batteries
,”
Mater. Lett.
167
,
102
105
(
2016
).
104.
K.
Biswas
,
L. D.
Zhao
, and
M. G.
Kanatzidis
, “
Tellurium-free thermoelectric: The anisotropic n-type semiconductor Bi2S3
,”
Adv. Energy Mater.
2
,
634
638
(
2012
).
105.
Y.
Cheng
 et al, “
Deep-level defect enhanced photothermal performance of bismuth sulfide–gold heterojunction nanorods for photothermal therapy of cancer guided by computed tomography imaging
,”
Angew. Chem., Int. Ed.
57
,
246
251
(
2018
).
106.
Y.
Yang
 et al, “
Hydrophilic Cu3BiS3 nanoparticles for computed tomography imaging and photothermal therapy
,”
Part. Part. Syst. Charact.
32
,
668
679
(
2015
).
107.
Z.
Li
 et al, “
Untrasmall Bi2S3 nanodots for in vivo X-ray CT imaging-guided photothermal therapy of cancer
,”
RSC Adv.
7
,
29672
29678
(
2017
).