Next Article in Journal
Michael Addition Reaction Catalyzed by Imidazolium Chloride to Protect Amino Groups and Construct Medium Ring Heterocycles
Previous Article in Journal
Bixa orellana L. (Bixaceae) and Dysphania ambrosioides (L.) Mosyakin & Clemants (Amaranthaceae) Essential Oils Formulated in Nanocochleates against Leishmania amazonensis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Frontiers of Nanomaterials (SnS, PbS and CuS) for Dye-Sensitized Solar Cell Applications: An Exciting New Infrared Material

by
Edson L. Meyer
1,
Johannes Z. Mbese
2,* and
Mojeed A. Agoro
1,2,*
1
Department of Chemistry, University of Fort Hare, Private Bag X1314, Alice 5700, South Africa
2
Fort Hare Institute of Technology, University of Fort Hare, Private Bag X1314, Alice 5700, South Africa
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(23), 4223; https://doi.org/10.3390/molecules24234223
Submission received: 7 October 2019 / Revised: 11 November 2019 / Accepted: 12 November 2019 / Published: 20 November 2019
(This article belongs to the Section Nanochemistry)

Abstract

:
To date, extensive studies have been done on solar cells on how to harness the unpleasant climatic condition for the binary benefits of renewable energy sources and potential energy solutions. Photovoltaic (PV) is considered as, not only as the future of humanity’s source of green energy, but also as a reliable solution to the energy crisis due to its sustainability, abundance, easy fabrication, cost-friendly and environmentally hazard-free nature. PV is grouped into first, second and third-generation cells. Dye-sensitized solar cells (DSSCs), classified as third-generation PV, have gained more ground in recent times. This is linked to their transparency, high efficiency, shape, being cost-friendly and flexibility of colour. However, further improvement of DSSCs by quantum dot sensitized solar cells (QDSSCs) has increased their efficiency through the use of semiconducting materials, such as quantum dots (QDs), as sensitizers. This has paved way for the fabrication of semiconducting QDs to replace the ideal DSSCs with quantum dot sensitized solar cells (QDSSCs). Moreover, there are no absolute photosensitizers that can cover all the infrared spectrum, the infusion of QD metal sulphides with better absorption could serve as a breakthrough. Metal sulphides, such as PbS, SnS and CuS QDs could be used as photosensitizers due to their strong near infrared (NIR) absorption properties. A few great dependable and reproducible routes to synthesize better QD size have attained much ground in the past and of late. The injection of these QD materials, which display (NIR) absorption with localized surface plasmon resonances (SPR), due to self-doped p-type carriers and photocatalytic activity could enhance the performance of the solar cell. This review will be focused on QDs in solar cell applications, the recent advances in the synthesis method, their stability, and long term prospects of QDSSCs efficiency.

1. Introduction

Energy production, distribution and sustainability have become one of the pressing challenges in underdeveloped countries. This is because of the low power output from the generating dams due to changes in climatic conditions. The political and geometric challenges of oil and gas power plants, nuclear power stations and coal-fired power plants are experienced globally. This has resulted in a binary advantage for photovoltaic application as a solution to the energy crisis and climate change [1,2]. Solar energy is one of the most promising and sustainable photovoltaic sources of power due to their abundance. Photovoltaic (PV) is classified as a green energy source, which is environmentally hazard-free, cost-friendly, and can easily be fabricated. Their market growth of 40% over two decades has made them the fastest-growing energy innovation [3].
Solar photovoltaic uses the photoelectric effect as direct sunlight conversion into electricity, such as silicon/thin-film solar cells or through photochemical effect, such as quantum dot sensitized solar cells (QDSSCs) [4,5]. The use of QDSSCs over other types of solar cells is linked to their tunable bandgap, size-dependence, high extinction coefficients, ease of fabrication and good energy conversion efficiency with multiple excitations [6,7,8,9,10]. Research on QDSSCs, as third-generation solar cells with promising prospects, has increased because of the increase in the number of research outputs (as seen in Figure 1) on QDSSCs year by year.
According to Shockley–Queisser [11], the first-generation solar cell with 33% efficiency and the second-generation solar cell with 54% efficiency has paved the way for third-generation solar cell thin-film technology to a better stage of commercialization in PV technology [12,13,14]. Third-generation solar cells using near-infrared (NIR) material quantum dots, such as PbS, CuS and SnS, could play a vital role in increasing conversion efficiency. This will replace the low spectral response, which is one of the major shortfalls of the ideal DSSCs, by possessing a better-oxidized electron at the LUMO level into the conduction band, as well as better regeneration by HUMO to their oxidized form [15,16]. In a typical build-up of DSSCs, QDSSCs portrait a similar architecture with distinctive characteristics like several microns thick semiconductor photoanode (ZnO or TiO2), photosensitizers, such as QDs (CuS, SnS, PbS), CE and electrolyte with redox couple as displayed in Figure 2. The working principle of QDSSCs is as follows: Upon the absorption of light, QDs photosensitizer absorbed on the TiO2 to generate photo-excited electrons at the interface of the conduction band (CB) and QDs. Followed by the oxidation reaction of the oxidize electron from the QDs at the (CB) and the holes synchronously oxidize the electrolyte by a reduction reaction, which is quickly migrated to the external circuit through the conductive substrate and the electrolyte interfaces thereby creating an electric current. The oxidized electrolyte is recycled back by donating oxidized electrons through the cycling circuit in the cell. Semiconductor materials, such as CuS, SnS, PbS, are commonly used to fabricate quantum dots (QDs) [6,17]. The QD photosensitizer has various features in the working mechanism and structure of photovoltaic cells; the robustness of the cell is increased, through the control of its size due to their tunable bandgap, easy fabrication and near-infrared properties [17]. One of the major challenges to good size-dependence of QDs is the pathways to synthesized size QDs [18,19,20,21,22,23]. This study will highlight the recent fabrication route, the advances in QDs and their potential in QDSSCs applications, as shown in Figure 2.

2. Recent Advances in SnS, CuS and PbS QDs Synthesis

The fabrication of commercialized film in optoelectronic and solar cell applications using the quantum dot near-infrared is an in-demand area of research. Semiconductors IV–VI series like PbS, CuS and SnS with low narrow bandgaps have been highlighted as a current hotspot in energy conversion research (see Table 1). Their easy reproducibility, cost-friendliness, stability and promising performances have made researchers attracted to these semiconductors [24]. The adoption of PbS, CuS and SnS in solar cell applications is due to their novel surface properties, shape, size, composition and mono-disparity. In addition, several overviews on the recent synthesis and application of quantum dot near-infrared have been investigated. Therefore, a great emphasis on the methods that give better uniform quantum dot materials is needed, such as single-source precursors [25], hydrothermals [26], template-directed approaches [27], hot-injection [28], etc.
Alivisatos et al. synthesized copper diethyldithiocarbamate using oleic acid and dodecanethiol via the mixed solvent. They obtained Cu2S nanocrystal from ammonium and Cu(II) acetylacetonate as their starting materials through the injection reaction [39]. Application of this fabricated nanocrystal with CdS in photovoltaics yielded 1.6% energy conversion efficiency on plastic and glass substrate with a longer stability. Zhao et al. revealed Cu2-xS of different nanocrystal compositions from (covellite) CuS to Cu1.97S through the optimization of various synthetic techniques, such as solventless thermolysis, sonoelectrochemical and hydrothermal approaches [39]. Characteristic evaluation of these materials’ properties indicated that CuS is more stable than Cu2S [40]. On the other hand, solventless thermolysis in an N2 atmosphere was utilized to produce uniform Cu2S from a copper thiolate precursor, with growth from small nanoparticles to nanodisks and orthorhombic to monoclinic Cu2S [41].
Karthik et al. adopted a single-source precursor method to fabricate CuS by solution thermolysis using [Cu(SON(CNiPr2)2)2] 1,1,5,5-tetra isopropyl-2-thiobiuret of copper(II) complex. Various morphologies were observed, such as trigonal crystallites, hexagonal disks and spheres. This points to factors like the concentration of the precursor, reaction temperature and growth time. On the other hand, the used of oleylamine in the thermolysis approach gives rise to Cu7S4 nanoparticles as a blend of anilite (orthorhombic) and roxbyite (monoclinic) phases (see Figure 3). The injection of octadecene solution precursor into hot oleylamine produces pure anilite Cu7S4 nanoparticles, while the injection of oleylamine solution precursor into hot dodecanethiol resulted in djurleite Cu1.94S nanoparticles. Both anilite and djurleite phases produce an indirect bandgap in the optical properties of the materials [42].
Various synthetic routes are employed to fabricate SnS quantum dots. Hickey et al. adopted thioacetamide through the synthetic route of hot injection with oleylamine and oleic acid, octadecene, Sn[N(SiMe3)2] and trioctylphoshine to fabricate SnS nonocrystals [39]. To control the shape of nanoparticles, the ratio of oleylamine/oleic acid was altered. A similar approach was used to fabricate SnS by Liu et al. through the injection of S(SiMe3)2 solutions with octadecene into SnCl2 solution in oleylamine [43]. The control of the size distribution was achieved through changing synthesis thermal condition; this produced the orthorhombic phase at all temperatures. Also, the colloidal approach of the turnable size of SnS is well detailed. They used an injection of the precursor, such as sulfur-oleylamine with Sn-oleylamine into hexamethyldisilazane (HMDS) at different temperatures. This produces SnS nanocrystals of a direct bandgap of 1.63 to 1.68 eV. This is directly linked to the size and morphology obtained in the range of 8–60 nm and spherical shape, while the unique crystal is due to larger particle sizes [44].
One of the most appealing semiconductors of the IV–VI series are the PbS QDs, which is due to their strong quantum confinement effects and near-infrared abilities. High-quality and quantity PbS has been produced using the relatively greener, solventless and heterogeneous method [45]. They use analytical grade oleylamine and PbCl2 which was blended and heated with the addition of sulfur-oleylamine. The process was stopped with the addition of hexane and the precipitate was centrifuged to obtain PbS as the final product. The characterization of PbS using photoluminescence revealed a 40% quantum yield from the particles and a stokes of 10 meV. Cao et al. [46] adopted the non-coordinating solvent approach by stabilized oleylamine in a reaction involving Pb stearate and S, resulting in a high-quality PbS with strong emissions and absorption in the near-infrared area. A similar route was used through a mixture of Pb oleate and bis(trimethylsilyl)sulphide according to a study by Hines and Scholes. Their concentrations were on the particle size of the final fabricated materials after the precipitate had been removed. Their finding revealed a turnable particle size and bandgap in the near infrared region. The emission spectrum further cemented the narrow size possession of the material (as seen in Figure 4) [47].

3. Overview of Single-Source Molecular Precursor Synthesis of QDs

The main constraint in the fabrication of QDs that have a better size-dependence and are more mono-disperse is the synthetic route that is adopted. This has led to innovative approaches to fabricating quantum dots of control shape and size [49]. Fabrication approaches including the single-source approach, hot injection, colloidal route, hydrothermal, microwave, solvothermal, electrodeposition, aqueous solution, pyrolysis, precipitation [50,51,52,53,54,55,56,57,58,59], etc., has been adopted for fabrication QDs. Single-source precursors have been distinguished as an effective approach that produces high-quality QDs. This is owing to their shape and controlled size through the use of capping agents in the decomposition of single-source precursors at low temperatures by one solvothermal route [60]. The advantages of this technique are numerous, such as ease of handling, low-temperature growth, air and moisture stability and low toxicity. This has placed single-source precursors over conventional precursors; using one precursor reduces the possibility of contamination on the film, and this allows for intrinsic control of the film, which occurs during the pre-reaction. Also, they used ligands to reduce impurities, such as carbon in the alkyl group which is incorporated in the film, as well as offering a degree of control. However, some precursors allow the control of the phase deposited through the use of molecular cluster compounds and core structure precursor molecules [60]. Overview of the approach on the following semiconducting materials, such as PbS, CuS and SnS quantum dots will be discussed.

3.1. PbS Quantum Dots

The study by Bo Hou et al. adopted a synthetic path to fabricate reliable, high performance and effective QDSSCs using PbS QDs. These were achieved by altering the precursor concentration at constant temperature and time. This resulted in many various sizes with high reproducibility and narrow size distribution of the fabricated PbS QDs (Figure 5) [61]. Boadi et al. reported the decomposition of Pb(S2CNiBu2)2 to form PbS with various morphologies based on the temperature utilized in the fabrication processes [62]. The material thermalized at 100 °C and displaced a sphere in diameter of 6.3 nm. While cubic crystallites at a diameter of 60 nm in a mixture of 150 °C and small spherical quantum dots were produced. However, the decomposition of Pb(S2CNEt2)2 at 230 °C in hot dodecylamine and phenyl ether produces large cubes in a similar study by Lee et al. [63]. Boadi et al. reported that the injection of hot phenyl ether with Pb(S2CNEt2)2 precursor, resulted in well-defined structures and star-shaped PbS QDs [62].

3.2. CuS Quantum Dots

Copper chalcogenides have been well reported for their localized surface plasma resonances and their displaced strong NIR absorption. CuS is intermediate with strong potential in the photocatalytic application and self-doped p-type carriers due to NIR absorption [64]. In addition, different capping agents can be utilized to fabricate various morphologies of CuS QDs [65,66,67]. It is very important to look into major factors, such as the optical properties, microstructure defects, crystal growth procedures and cubic-hexagonals of CuS QDs due to ultrafast nonlinear optical properties. Looking at photovoltaic semiconductor materials, it is of great importance to understand the evolution of the optical properties from nanoparticles to QDs [68]. A recent study has shown that fabricated CuS QDs using capping agents, such as polyvinylpyrrolidone, produces particles with sizes of 2–4 nm and 5–7 nm, according to Luther et al. (as seen in Figure 6), with hexagonal and cubic phases being observed for the QDs. Furthermore, planar defects like stacking, twins and dislocations are due to a reasonable amount of CuS QDs.
Crystallographic phases suitable for crystal growth of CuS QDs are dominated by oriented attachment, while optical absorption of CuS QDs is evaluated in linear and nonlinear regimes. The presence of electron/hole traps indicated that CuS QDs displaced the lower bandgap energy [69].

3.3. SnS Quantum Dots

SnS are semiconductors in group IV–VI. They have a direct and indirect bandgap of 1 eV to 1.3eV and high coefficient absorption [70]. These absorptions were within the visible and near- infrared areas of the electromagnetic spectrum. These are the area where optimum efficiency in QDSSCs can be attained. This further underlining SnS QDs as potentials materials for photovoltaic technology. Capping agents and temperature affect the shape and size resulting in SnS within the range of 10–28 nm. The mixture of octadecylamine capped and oleic acid produce cubic and small nanocrystals using single molecular precursors on fabricated SnS QDs [71]. Deng et al. prepared SnS QDs turnable size using tinoleylamine solution into hexamethyldisilazane through the injection of the sulfur-oleylamine precursor at different heats. The smaller optical bandgap indicated larger nanocrystals [72].

4. Molecular Precursor Complexes

Molecular precursor has paved way for the synthesis of monodispersed, better crystalline materials, semiconducting quantum dots, containing metal and chalcogenide [73]. Hiroki and Mitsunobu reported the use of two molecular Cu(II) precursor solutions of EDTA and propylamine obtained from both amine and formic mixtures. The adjustment of Cu concentrations in the precursor solution of ethanolics gives rise to novel Na-free FTO films [74]. The major draw-back of molecular precursors are pinpoints to the fabrication route of the ligand by producing small yields, which constitutes a shortfall for large scale applications in various fields of research. Chemical properties of the complexes represent the second draw-back as a result of their valence orbital having been shielded [75]. Coordination compounds often resulted in different structural geometry based on properties of metal ion and ligands [76].
Therefore the complex precursor should be of good solubility in solvents and have quite good thermal and chemical stabilities. The use of the dithiocarbamate ligand has given rise to the synthesis of various novel molecular complexes. They have the ability to produce complexes with transition metals, in addition to stabilizing metals ion at high oxidation states. These have expanded their application in various fields including medicine, polymer photo stabilizers, biology, antioxidants, chemistry, photovoltaic, and so many more. Dithiocarbamate complexes of Sn(II), Cu(II) and Pb(II) have been utilized as molecular precursors for thin film growth [77].

4.1. Dithiocarbamate Cu(II) Complexes

Geraldo et al. used a bidentate complex of [Cu{S2CNR(CH2CH2OH)}2] in the dx2–y2 orbital state to form square planar geometry. This resulted in a distorted square planar molecule crystallised complex and dimer intermolecular Cu bond in the form of independent centrosymmetric monomers and similar solid-states [78]. Some dithiocarbamate complexes with Cu(II), such as pyrrolidinedithiocarbamate have also been reported for their strong antiviral, antioxidant and anti-inflammatory properties and they are able to transport metal through membranes [79]. Four coordinate bidentate metal ion structures within ligand and two sulphur of square planar geometries were reported by De Lima et al. [80]. Victor et al. in their study used a solvent molecule of Cu(Et2dtc)2 complex, which pointed to the axial of copper ion in the photochemical activity. The dimer coordinated with the Et2dtc• radical produced involves the reaction of initial complex and intermediate; accompanied by microseconds the dithiocarbamate radicals, which reunited to produce a tetranuclear cluster. These turned out to be challenging on stationary photolysis when looking at the mechanism of photochemical transformations of the Cu(Et2dtc)2 complex [81,82].

4.2. Dithiocarbamate Pb(II) Complexes

The use of Pb(S2CNRR)2 formulated as lead(II) dithiocarbamate complexes to fabricate PbS QDs through single-source precursors approach has been reported by [83]. The advantage of these over traditional organometallics is the absence of toxic metals like lead alkyls or H2S. They are easy to fabricate and stable for months and give high yields with thermolysis. The decomposition of Pb (S2CNEt2)2 complex in tri-n-octyl phosphine oxide produces cube-shaped quantum dots [82]. Similarly, Cheon et al. produced rod-based star-shapes, cubes and highly faceted and truncated octahedrons of different structures with the same precursor (Figure 7) [63]. On the other hand, spherical to cubic shapes can be produced by altering the dodecanethiol ligand ratio through single- source precursors to form PbS. Hydrolytic processes were adopted to fabricate star-shaped PbS QDs using bis(thiosemicarbazide)lead(II) [Pb(TSC)2Cl2] [84]. Cubes shapes of PbS were synthesized using ethylene diamine with Pb xanthate complex in room temperature decomposition through the greener synthetic route [85]. Pb(II) complexes of aromatic carboxylate with thiosemicarbazide or thiourea were adopted as a molecular precursor to fabricate a stable PbS QDs by aqueous or non-aqueous solvent decomposition [86].

4.3. Dithiocarbamate Sn(II) Complexes

Tin(II) and Tin(IV) species are one of the semiconductors with interesting redox chemistry and are stable in both metal–organic complexes and organometallics. Fabrication of PbS thin-films with different single-source precursors is well documented in the literature, such as Sn(S2CNEt2)4, Sn(SPh)4 and Sn(SCH2CH2S)2, Sn(S-cyclohexane)4, and Sn(SCH2CF3)4 [87,88]. Punarja et al. use Tin(II) of [Sn(S2CNRR′)2] to fabricate unsymmetrical dithiocarbamates. The thermolysis of the complexes produces absorbance in the near-infrared region with an optical bandgap of 1.2eV, good morphology and composition [89].

5. Recent Advances in PbS, SnS and CuS Quantum Dot Application for Dye-Sensitized Solar Cells

To replace the low efficiency in ideal DSSCs and promote their commercialization as potential photovoltaic devices, semiconductor QD materials with better light absorbers could be used as a replacement in QDSSCs. Semiconductor materials with unique optoelectronic properties, multiple exciton generation, high extinction coefficients and bandgaps with a size control less than 10 nm are classified as special QDs [90,91,92,93,94,95]. Ideal QD dyes should have many properties including producing a photoexcited electron upon the absorption of light. In 2008, Zhao et al. used recombination control to produce 8.21% efficiency in their study [96]. Ren et al. overcome the instability of QDSSCs in their study through metal oxyhydroxide coatings of photoanode to produce a 9.73% efficiency [97]. Du et al. reported efficiency above 11.39% by using carbon CE-based QDSSCs [98]. Furthermore, Jiao et al. adopted nitrogen-doped mesoporous carbons with an efficiency exceeding 12% as unique potential QD-based cells [99]. Recently, Luther’s group obtained an efficiency of 13.43%, which is currently the optimum efficiency in QDSSCs [100].

5.1. Recent Advances in PbS Quantum Dots

The direct bandgap of 0.37eV with a vast exciton Bohr radius of 18 nm at room temperature has made PbS QDs one of the potential p-type semiconductors [101]. Owning to their 400–2500 nm near-infrared area and visible region through adjusting the dot size [102], Sargent et al. utilized the high-mobility hybrid perovskite by inserting PbS QDs into their matrix, which resulted in 4.9% electroluminescence power conversion efficiency [103]. Chuang et al. synthesized electron blocking/hole-extraction layer and light-absorbing layer thin-film by using PbS-EDT CQD and PbS-TBAI to construct a long-life air stable solar cell with 8.55% efficiency [104]. While Lee et al. adopted the solid-state PbS DQ-SSCs on substrate using SILAR/spiro-OMeTAD (2,2,7,7-tetrakis (N,N-di-p-methoxyphenylamine)-9,9-spirobi-fluorene) HTL/Au to achieve 1.46% [105]. Furthermore, Seok et al. constructed a spin-SILAR/PEDOT:PSS (poly(3,4-ethylene- dioxythiophene) polystyrene sulfonate)/Au/P3HT (Poly-3-hexylthiophene) with 2.0% efficiency using (bl-TiO2)/mp-TiO2/PbS QDs FTO/blocking. This later improved by using the same approach with a CH3NH3PbI3 (MAPbI3) shell resulting in 3.2% efficiency [106].
Recently, Im et al. used mesoscopic PbS QD-SSCs to achieve 4.96% efficiency and later improved the efficiency to 9.2% with the injection of an MAPbI3 interlayer [107]. Karthik et al. used the same mesoscopic PbS QDSSCs by inserting CuS in a rapid rate to the cationic exchange reaction. With the aid of Cu-oleylamine and oleic acid as the capping agents to produce 8.07% efficiency under 1 sun test condition (AM1.5G 100 mW/cm2) [108]. Seok et al. adopted the colloidal route to fabricate PbS QDSSCs merged with FTO/bl-TiO2/m-TiO2/multiple-layered PbS QDs/P3HT/PEDOT:PSS/Au to produce 2.9% power conversion efficiency [109]. The efficiency was later improved to 3.9% using TiO2 nanorod ETL/PbS QDs/P3HT HTL interface by radial directional charge transport to fabricating PbS QDSSCs [110]. Similarly, Zhang et al. achieved 3.57% efficiency by TiO2 nanorod array/spiro-OMeTAD/Au in FTO/compact PbS QDSSCs (as seen in Figure 8) [111].

5.2. Recent Advances in SnS Quantum Dots

SnS QDs displayed better ways of absorbing the whole solar spectrum due to blue shift absorption properties. Deepa and Nagaraju, in their study, fabricated SnS QDs with sizes of 2.4 to 14.4 nm through the chemical precipitation approach by controlling the growth time. The SnS QDs with different structures were fabricated with and without the buffer layer and apply in solar cell test (see Table 2). The efficiency varies when the QD dyes were loaded on the cell with the buffer layer, which reduces the efficiency and improved the fill factor (FF) [112,113]. Tang et al. gives a detailed photocatalytic activity of rhodamine B under the halogen lamp of fabricated SnS QDs using Sn(II), oleic acid, octadecene, thioacetamide, oleyamine and trioctylphosphine as the starting materials [114]. Das and Dutta reported efficient degradation of trypan blue dye photocatalyst in sunlight with the aids of mercaptoacetic acid capping agent to fabricate SnS nanorods [115]. To further improve SnS QDSSC efficiency, the limitations that hinder their large scale commercialization such as secondary phase formation and bulk defects, diffusion length, short minority carrier lifetime, band alignment, solar cell configuration, and back contact should be given important consideration to enhance the efficiency of QDSSCs

5.3. Recent Advances in CuS Quantum Dots

Metal sulphides, such as copper sulfides (Cu2S and CuS) [67], are well-known in QDSSCs as efficient materials with redox couple of polysulfide electrolytes. They have displayed good electrocatalytic activity, low resistance and strong stability with redox reaction polysulfides [67]. In addition, CuS and Cu2S are generally adopted in QDSSCs. The near-infrared absorption of CuS due to self-doped p-type carriers and photocatalytic activity has made them potential semiconducting materials for various applications [39]. The J–V curve of the QDSSCs was evaluated in 1 sunlight and the finding revealed a display of high efficiency for CuS 2 h-based QDSSCs with (η) of 4.27%, (FF) of 0.49, (Voc) of 0.603 V and (Jsc) of 14.31 mA cm−2, but are much better when compared with conventional Pt CEs ( η = 1.11%, Voc = 0.523 V, FF = 0.24 and Jsc = 8.83 mA cm−2). CuS 2 h-based efficiency is linked to the catalytic properties and the aid of CuS counter electrolyte low charge transfer resistance with polysulfide electrolytes (as seen in Figure 9).

6. Conclusions, Further Improvement and Future Studies

Semiconductors with near-infrared regions, quantum confinement and high absorption coefficients are highly desired in photovoltaic cell research. New scientific advances and the potential of greater power conversion efficiency in these semiconductors has contributed to the noteworthy strides in the third-generation solar cells. To be a pace-setter in third generation photovoltaic technology, further advancements should be made on improving the efficiency of QDSSCs. Additionally, the emphasis on the stability of efficient electrolytes in fabricated solar cells is highly recommendable. The synthesis of absorbing QD near infra-red region materials that could absorb the whole visible light harvesting regions should be also considered. Research on the reduction of charge recombination and the fabrication of QDs that will absorb photoelectrodes with better porosity should be investigated. Looking at the numbers of research articles on semiconductor QDs and the huge challenges encountered in the improvement of QDSSCs, new breakthroughs in improving the efficiency of QDSSCs is inevitable.

Author Contributions

Conceptualization, J.Z.M., M.A.A., and E.L.M.; investigation, J.Z.M., M.A.A. and E.L.M.; resources, J.Z.M., M.A.A., and E.L.M.; data curation, M.A.A.; writing—original draft preparation, M.A.A.; writing—review and editing, J.Z.M., M.A.A., and E.L.M.; supervision, J.Z.M., and E.L.M.; funding acquisition, J.Z.M. and E.L.M.

Funding

The research was funded by the National Department of Science and Technology through its Energy Research Programme and the National Research Foundation (GUN 109769). The Govan Mbeki Research and Development Centre at the University of Fort Hare also contributed financially to the research.

Acknowledgments

The authors appreciate the support from Mr. Julian C. Nwodo for the reproduction of data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yuksel, I. Renewable energy status of electricity generation and future prospect hydropower in Turkey. Renew. Energy 2013, 50, 1037–1043. [Google Scholar] [CrossRef]
  2. El Chaar, L.; El Zein, N. Review of photovoltaic technologies. Renew. Sustain. Energy Rev. 2011, 15, 2165–2175. [Google Scholar] [CrossRef]
  3. Shah, A.; Torres, P.; Tscharner, R.; Wyrsch, N.; Keppner, H. Photovoltaic technology: The case for thin-film solar cells. Science 1999, 285, 692–698. [Google Scholar] [CrossRef] [PubMed]
  4. Cook, T.R.; Dogutan, D.K.; Reece, S.Y.; Surendranath, Y.; Teets, T.S.; Nocera, D.G. Solar energy supply and storage for the legacy and non-legacy worlds. Chem. Rev. 2010, 110, 6474–6502. [Google Scholar] [CrossRef] [PubMed]
  5. Lewis, N.S. Toward cost-effective solar energy use. Science 2007, 315, 798–801. [Google Scholar] [CrossRef]
  6. Chebrolu, V.T.; Kim, H.J. Recent progress in quantum dot sensitized solar cells: An inclusive review of photoanode, sensitizer, electrolyte, and the counter electrode. J. Mater. Chem. C 2019, 7, 4911–4933. [Google Scholar] [CrossRef]
  7. Hossain, M.A.; Jennings, J.R.; Shen, C.; Pan, J.H.; Koh, Z.Y.; Mathews, N.; Wang, Q. CdSe-sensitized mesoscopic TiO2 solar cells exhibiting >5% efficiency: Redundancy of CdS buffer layer. J. Mater. Chem. 2012, 22, 16235–16242. [Google Scholar] [CrossRef]
  8. Kamat, P.V. Quantum dot solar cells. Semiconductor nanocrystals as light harvesters. J. Phys. Chem. C 2008, 112, 18737–18753. [Google Scholar] [CrossRef]
  9. Ma, W.; Swisher, S.L.; Ewers, T.; Engel, J.; Ferry, V.E.; Atwater, H.A.; Alivisatos, A.P. Photovoltaic performance of ultrasmall PbSe quantum dots. ACS Nano 2011, 5, 8140–8147. [Google Scholar] [CrossRef]
  10. Segets, D.; Lucas, J.M.; Klupp Taylor, R.N.; Scheele, M.; Zheng, H.; Alivisatos, A.P.; Peukert, W. Determination of the quantum dot band gap dependence on particle size from optical absorbance and transmission electron microscopy measurements. ACS Nano 2012, 6, 9021–9032. [Google Scholar] [CrossRef]
  11. Shockley, W.; Queisser, H.J. Detailed balance limit of efficiency of p-n junction solar cells. J. Appl. Phys. 1961, 32, 510–519. [Google Scholar] [CrossRef]
  12. Jørgensen, M.; Norrman, K.; Gevorgyan, S.A.; Tromholt, T.; Andreasen, B.; Krebs, F.C. Stability of polymer solar cells. Adv. Mater. 2012, 24, 580–612. [Google Scholar] [CrossRef] [PubMed]
  13. Zhu, J.; Yu, Z.; Burkhard, G.F.; Hsu, C.M.; Connor, S.T.; Xu, Y.; Wang, Q.; McGehee, M.; Fan, S.; Cui, Y. Optical absorption enhancement in amorphous silicon nanowire and nanocone arrays. Nano Lett. 2008, 9, 279–282. [Google Scholar] [CrossRef] [PubMed]
  14. Kim, M.R.; Ma, D. Quantum-dot-based solar cells: Recent advances, strategies, and challenges. J. Phys. Chem. Lett. 2014, 6, 85–99. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, C.H.; Chen, P.Y.; Chen, W.J.; Wang, T.L.; Shieh, Y.T. Spectroscopic evidences of synergistic co-sensitization in dye-sensitized solar cells via experimentation of mixture design. Electrochim. Acta 2013, 107, 170–177. [Google Scholar] [CrossRef]
  16. Yang, L.; Leung, W.W.F.; Wang, J. Improvement in light harvesting in a dye sensitized solar cell based on cascade charge transfer. Nanoscale 2013, 5, 7493–7498. [Google Scholar] [CrossRef]
  17. Ye, M.; Gao, X.; Hong, X.; Liu, Q.; He, C.; Liu, X.; Lin, C. Recent advances in quantum dot-sensitized solar cells: Insights into photoanodes, sensitizers, electrolytes and counter electrodes. Sustain. Energy Fuels 2017, 1, 1217–1231. [Google Scholar] [CrossRef]
  18. Shrestha, A.; Batmunkh, M.; Tricoli, A.; Qiao, S.Z.; Dai, S. Near-infrared active lead chalcogenide quantum dots: preparation, post-synthesis ligand exchange, and applications in solar cells. Angew. Chem. Int. Ed. 2019, 58, 5202–5224. [Google Scholar] [CrossRef]
  19. Lim, J.; Bae, W.K.; Kwak, J.; Lee, S.; Lee, C.; Char, K. Perspective on synthesis, device structures, and printing processes for quantum dot displays. Opt. Mater. Express 2012, 2, 594–628. [Google Scholar] [CrossRef]
  20. Kershaw, S.V.; Susha, A.S.; Rogach, A.L. Narrow bandgap colloidal metal chalcogenide quantum dots: Synthetic methods, heterostructures, assemblies, electronic and infrared optical properties. Chem. Soc. Rev. 2013, 42, 3033–3087. [Google Scholar] [CrossRef]
  21. Owen, J.; Brus, L. Chemical synthesis and luminescence applications of colloidal semiconductor quantum dots. J. Am. Chem. Soc. 2017, 139, 10939–10943. [Google Scholar] [CrossRef] [PubMed]
  22. Van Embden, J.; Chesman, A.S.; Jasieniak, J.J. The heat-up synthesis of colloidal nanocrystals. Chem. Mater. 2015, 27, 2246–2285. [Google Scholar] [CrossRef]
  23. Goubet, N.; Jagtap, A.; Livache, C.; Martinez, B.; Portalès, H.; Xu, X.Z.; Lobo, R.P.; Dubertret, B.; Lhuillier, E. Terahertz HgTe nanocrystals: Beyond confinement. J. Am. Chem. Soc. 2018, 140, 5033–5036. [Google Scholar] [CrossRef] [PubMed]
  24. Ma, W.; Luther, J.M.; Zheng, H.; Wu, Y.; Alivisatos, A.P. Photovoltaic devices employing ternary PbSx Se1−x nanocrystals. Nano Lett. 2009, 9, 1699–1703. [Google Scholar] [CrossRef] [PubMed]
  25. Jäger-Waldau, A. Progress in chalcopyrite compound semiconductor research for photovoltaic applications and transfer of results into actual solar cell production. In Practical Handbook of Photovoltaics; Academic Press: Cambridge, MA, USA, 2012; pp. 373–395. [Google Scholar]
  26. Klenk, R.; Klaer, J.; Köble, C.; Mainz, R.; Merdes, S.; Rodriguez-Alvarez, H.; Scheer, R.; Schock, H.W. Development of CuInS2-based solar cells and modules. Sol. Energy Mater. Sol. Cells 2011, 95, 1441–1445. [Google Scholar] [CrossRef]
  27. Lai, C.H.; Lu, M.Y.; Chen, L.J. Metal sulfide nanostructures: Synthesis, properties and applications in energy conversion and storage. J. Mater. Chem. 2012, 22, 19–30. [Google Scholar] [CrossRef]
  28. Zhao, Y.; Burda, C. Development of plasmonic semiconductor nanomaterials with copper chalcogenides for a future with sustainable energy materials. Energy Environ. Sci. 2012, 5, 5564–5576. [Google Scholar] [CrossRef]
  29. Kim, H.J.; Myung-Sik, L.; Gopi, C.V.; Venkata-Haritha, M.; Rao, S.S.; Kim, S.K. Cost-effective and morphology controllable PVP based highly efficient CuS counter electrodes for high-efficiency quantum dot-sensitized solar cells. Dalton Trans. 2015, 44, 11340–11351. [Google Scholar] [CrossRef]
  30. Mousavi-Kamazani, M.; Zarghami, Z.; Salavati-Niasari, M. Facile and novel chemical synthesis, characterization, and formation mechanism of copper sulfide (Cu2S, Cu2S/CuS, CuS) nanostructures for increasing the efficiency of solar cells. J. Phys. Chem. C 2016, 120, 2096–2108. [Google Scholar] [CrossRef]
  31. Tian, J.; Shen, T.; Liu, X.; Fei, C.; Lv, L.; Cao, G. Enhanced performance of PbS-quantum-dot-sensitized solar cells via optimizing precursor solution and electrolytes. Sci. Rep. 2016, 6, 23094. [Google Scholar] [CrossRef]
  32. An, J.; Yang, X.; Wang, W.; Li, J.; Wang, H.; Yu, Z.; Gong, C.; Wang, X.; Sun, L. Stable and efficient PbS colloidal quantum dot solar cells incorporating low-temperature processed carbon paste counter electrodes. Sol. Energy 2017, 158, 28–33. [Google Scholar] [CrossRef]
  33. Li, H.; Ji, J.; Zheng, X.; Ma, Y.; Jin, Z.; Ji, H. Preparation of SnS quantum dots for solar cells application by an in-situ solution chemical reaction process. Mater. Sci. Semicond. Process. 2015, 36, 65–70. [Google Scholar] [CrossRef]
  34. Ngoi, K.K.; Jun, H.K. Study of fabrication of fully aqueous solution processed SnS quantum dot-sensitized solar cell. Green Process. Synth. 2019, 8, 443–450. [Google Scholar] [CrossRef]
  35. Wu, Y.; Wadia, C.; Ma, W.; Sadtler, B.; Alivisatos, A.P. Synthesis and photovoltaic application of copper (I) sulfide nanocrystals. Nano Lett. 2008, 8, 2551–2555. [Google Scholar] [CrossRef] [PubMed]
  36. Zhao, Y.; Pan, H.; Lou, Y.; Qiu, X.; Zhu, J.; Burda, C. Plasmonic Cu2−x S nanocrystals: Optical and structural properties of copper-deficient copper (I) sulfides. J. Am. Chem. Soc. 2009, 131, 4253–4261. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, Y.B.; Chen, L.; Wu, L.M. The structure-controlling solventless synthesis and optical properties of uniform Cu2s nanodisks. Chem. Eur. J. 2008, 14, 11069–11075. [Google Scholar] [CrossRef]
  38. Spalatu, N.; Hiie, J.; Kaupmees, R.; Volobujeva, O.; Krustok, J.; Oja Acik, I.; Krunks, M. Post-deposition processing of SnS thin films and solar cells: prospective strategy to obtain large, sintered and doped sns grains by recrystallization in the presence of a metal halide Flux. ACS Appl. Mater. 2019, 11, 17539–17554. [Google Scholar] [CrossRef]
  39. Abdelhady, A.L.; Ramasamy, K.; Malik, M.A.; O’Brien, P.; Haigh, S.J.; Raftery, J. New routes to copper sulfide nanostructures and thin films. J. Mater. Chem. 2011, 21, 17888–17895. [Google Scholar] [CrossRef]
  40. Hickey, S.G.; Waurisch, C.; Rellinghaus, B.; Eychmüller, A. Size and shape control of colloidally synthesized IV−VI nanoparticulate tin (II) sulfide. J. Am. Chem. Soc. 2008, 130, 14978–14980. [Google Scholar] [CrossRef]
  41. Liu, H.; Liu, Y.; Wang, Z.; He, P. Facile synthesis of monodisperse, size-tunable SnS nanoparticles potentially for solar cell energy conversion. Nanotechnology 2010, 21, 105707. [Google Scholar] [CrossRef]
  42. Ramasamy, K.; Malik, M.A.; Revaprasadu, N.; O’Brien, P. Routes to nanostructured inorganic materials with potential for solar energy applications. J. Mater. Chem. 2013, 25, 3551–3569. [Google Scholar] [CrossRef]
  43. Oda, Y.; Shen, H.; Zhao, L.; Li, J.; Iwamoto, M.; Lin, H. Energetic alignment in nontoxic SnS quantum dot-sensitized solar cell employing spiro-OMeTAD as the solid-state electrolyte. Sci. Technol. Adv. Mater. 2014, 15, 035006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Bai, B.; Kou, D.; Zhou, W.; Zhou, Z.; Tian, Q.; Meng, Y.; Wu, S. Quaternary Cu2ZnSnS4 quantum dot-sensitized solar cells: Synthesis, passivation and ligand exchange. J. Power Sources 2016, 318, 35–40. [Google Scholar] [CrossRef]
  45. Jiao, S.; Wang, J.; Shen, Q.; Li, Y.; Zhong, X. Surface engineering of PbS quantum dot sensitized solar cells with a conversion efficiency exceeding 7%. J. Mater. Chem. A 2016, 4, 7214–7221. [Google Scholar] [CrossRef]
  46. Cademartiri, L.; Bertolotti, J.; Sapienza, R.; Wiersma, D.S.; Von Freymann, G.; Ozin, G.A. Multigram scale, solventless, and diffusion-controlled route to highly monodisperse PbS nanocrystals. J. Phys. Chem. B 2006, 110, 671–673. [Google Scholar] [CrossRef]
  47. Liu, J.; Yu, H.; Wu, Z.; Wang, W.; Peng, J.; 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]
  48. Hines, M.A.; Scholes, G.D. Colloidal PbS nanocrystals with size-tunable near-infrared emission: Observation of post-synthesis self-narrowing of the particle size distribution. Adv. Mater. 2003, 15, 1844–1849. [Google Scholar] [CrossRef]
  49. Carey, G.H.; Abdelhady, A.L.; Ning, Z.; Thon, S.M.; Bakr, O.M.; Sargent, E.H. Colloidal quantum dot solar cells. Chem. Rev. 2015, 115, 12732–12763. [Google Scholar] [CrossRef]
  50. Manukumar, K.N.; Nagaraju, G.; Kishore, B.; Madhu, C.; Munichandraiah, N. Ionic liquid-assisted hydrothermal synthesis of SnS nanoparticles: Electrode materials for lithium batteries, photoluminescence and photocatalytic activities. J. Energy Chem. 2018, 27, 806–812. [Google Scholar] [CrossRef] [Green Version]
  51. Chauhan, H.; Singh, M.K.; Hashmi, S.A.; Deka, S. Synthesis of surfactant-free SnS nanorods by a solvothermal route with better electrochemical properties towards supercapacitor applications. RSC Adv. 2015, 5, 17228–17235. [Google Scholar] [CrossRef]
  52. Nørby, P.; Johnsen, S.; Iversen, B.B. Fine tunable aqueous solution synthesis of textured flexible SnS2 thin films and nanosheets. Phys. Chem. Chem. Phys. 2015, 17, 9282–9287. [Google Scholar] [CrossRef] [PubMed]
  53. Dai, X.; Shi, C.; Zhang, Y.; Liu, F.; Fang, X.; Zhu, J. SnS thin film prepared by pyrolytic synthesis as an efficient counter electrode in quantum dot sensitized solar cells. J. Nanosci. Nanotechnol. 2015, 15, 6813–6817. [Google Scholar] [CrossRef] [PubMed]
  54. Yin, D.; Liu, Y.; Dun, C.; Carroll, D.L.; Swihart, M.T. Controllable colloidal synthesis of anisotropic tin dichalcogenide nanocrystals for thin film thermoelectrics. Nanoscale 2018, 10, 2533–2541. [Google Scholar] [CrossRef] [PubMed]
  55. Thompson, J.R.; Ahmet, I.Y.; Johnson, A.L.; Kociok-Köhn, G. Tin (IV) chalcogenide complexes: Single source precursors for SnS, SnSe and SnTe nanoparticle synthesis. Eur. J. Inorg. Chem. 2016, 2016, 4711–4720. [Google Scholar] [CrossRef] [Green Version]
  56. Kevin, P.; Lewis, D.J.; Raftery, J.; Malik, M.A.; O’Brien, P. Thin films of tin (II) sulphide (SnS) by aerosol-assisted chemical vapour deposition (AACVD) using tin (II) dithiocarbamates as single-source precursors. J. Cryst. Growth 2015, 415, 93–99. [Google Scholar] [CrossRef]
  57. Henry, J.; Mohanraj, K.; Kannan, S.; Barathan, S.; Sivakumar, G. Structural and optical properties of SnS nanoparticles and electron-beam-evaporated SnS thin films. J. Exp. Nanosci. 2015, 10, 78–85. [Google Scholar] [CrossRef]
  58. Park, S.; Park, J.; Selvaraj, R.; Kim, Y. Facile microwave-assisted synthesis of SnS2 nanoparticles for visible-light responsive photocatalyst. J. Ind. Eng. Chem. 2015, 31, 269–275. [Google Scholar] [CrossRef]
  59. Lin, Y.T.; Shi, J.B.; Chen, Y.C.; Chen, C.J.; Wu, P.F. Synthesis and characterization of tin disulfide (SnS2) nanowires. Nanoscale Res. Lett. 2009, 4, 694. [Google Scholar] [CrossRef] [Green Version]
  60. Oluwalana, A.E.; Ajibade, P.A. Effect of temperature and capping agents on structural and optical properties of tin sulphide nanocrystals. J. Nanotechnol. 2019, 2019, 11. [Google Scholar] [CrossRef] [Green Version]
  61. Hou, B.; Cho, Y.; Kim, B.S.; Hong, J.; Park, J.B.; Ahn, S.J.; Sohn, J.I.; Cha, S.; Kim, J.M. Highly monodispersed PbS quantum dots for outstanding cascaded-junction solar cells. ACS Energy Lett. 2016, 1, 834–839. [Google Scholar] [CrossRef]
  62. Boadi, N.O.; Malik, M.A.; O’Brien, P.; Awudza, J.A. Single source molecular precursor routes to lead chalcogenides. Dalton Trans. 2012, 41, 10497–10506. [Google Scholar] [CrossRef] [PubMed]
  63. Lee, S.M.; Jun, Y.W.; Cho, S.N.; Cheon, J. Single-crystalline star-shaped nanocrystals and their evolution: Programming the geometry of nano-building blocks. J. Am. Chem. Soc. 2002, 124, 11244–11245. [Google Scholar] [CrossRef] [PubMed]
  64. Ratanatawanate, C.; Bui, A.; Vu, K.; Balkus Jr, K.J. Low-temperature synthesis of copper (II) sulfide quantum dot decorated TiO2 nanotubes and their photocatalytic properties. J. Phys. Chem. C 2011, 115, 6175–6180. [Google Scholar] [CrossRef]
  65. Zhang, S.; Zhong, H.; Ding, C. Ultrasensitive flow injection chemiluminescence detection of DNA hybridization using signal DNA probe modified with Au and CuS nanoparticles. Anal. Chem. 2008, 80, 7206–7212. [Google Scholar] [CrossRef]
  66. Li, Y.; Lu, W.; Huang, Q.; Li, C.; Chen, W. Copper sulfide nanoparticles for photothermal ablation of tumor cells. Nanomedicine 2010, 5, 1161–1171. [Google Scholar] [CrossRef] [Green Version]
  67. Li, B.; Xie, Y.; Xue, Y. Controllable synthesis of CuS nanostructures from self-assembled precursors with biomolecule assistance. J. Phys. Chem. C 2007, 111, 12181–12187. [Google Scholar] [CrossRef]
  68. Luther, J.M.; Jain, P.K.; Ewers, T.; Alivisatos, A.P. Localized surface plasmon resonances arising from free carriers in doped quantum dots. Nat. Mater. 2011, 10, 361. [Google Scholar] [CrossRef]
  69. Mary, K.A.; Unnikrishnan, N.V.; Philip, R. Role of surface states and defects in the ultrafast nonlinear optical properties of CuS quantum dots. APL Mater. 2014, 2, 076104. [Google Scholar] [CrossRef]
  70. Rahaman, S.; Jagannatha, K.B.; Sriram, A. Synthesis and Characterization of SnS Quantum Dots materialfor Solar Cell. Mater. Today Proc. 2018, 5, 3117–3120. [Google Scholar] [CrossRef]
  71. Buenzli, J.C.G. On the design of highly luminescent lanthanide complexes. Coord. Chem. Rev. 2015, 293, 19–47. [Google Scholar] [CrossRef]
  72. Deng, Z.; Han, D.; Liu, Y. Colloidal synthesis of metastable zinc-blende IV–VI SnS nanocrystals with tunable 493 sizes. Nanoscale 2011, 3, 4346–4351. [Google Scholar] [CrossRef] [PubMed]
  73. Nagai, H.; Sato, M. Heat treatment in molecular precursor method for fabricating metal oxide thin films. In Heat Treatment—Conventional and Novel Applications; Czerwinski, F., Ed.; InTech Open: London, UK, 2012; pp. 297–322. [Google Scholar]
  74. Nagai, H.; Sato, M. Molecular precursor method for fabricating p-Type Cu2O and metallic Cu thin films. In Modern Technologys Creating Thin-Film Systems and Coatings; InTech Open: London, UK, 2017; pp. 3–20. [Google Scholar]
  75. Moreau, F.; Audebrand, N.; Poriel, C.; Moizan-Baslé, V.; Ouvry, J. A 9, 9′-spirobifluorene based Metal–Organic Framework: Synthesis, structure analysis and gas sorption properties. J. Mater. Chem. 2011, 21, 18715–18722. [Google Scholar] [CrossRef] [Green Version]
  76. Le Natur, F.; Calvez, G.; Freslon, S.; Daiguebonne, C.; Bernot, K.; Guillou, O. Extending the lanthanide–terephthalate system: Isolation of an unprecedented Tb (III)-based coordination polymer with high potential porosity and luminescence properties. J. Mol. Struct. 2015, 1086, 34–42. [Google Scholar] [CrossRef]
  77. Luo, Y.; Bernot, K.; Calvez, G.; Freslon, S.; Daiguebonne, C.; Guillou, O.; Kerbellec, N.; Roisnel, T. 1, 2, 4, 5-Benzene-tetra-carboxylic acid: A versatile ligand for high dimensional lanthanide-based coordination polymers. Cryst. Eng. Commun. 2013, 15, 1882–1896. [Google Scholar] [CrossRef]
  78. Plyusnin, V.F.; Kolomeets, A.V.; Grivin, V.P.; Larionov, S.V.; Lemmetyinen, H. Photochemistry of dithiocarbamate Cu (II) complex in CCl4. J. Phys. Chem. A 2011, 115, 1763–1773. [Google Scholar] [CrossRef]
  79. Vickers, M.S.; Cookson, J.; Beer, P.D.; Bishop, P.T.; Thiebaut, B. Dithiocarbamate ligand stabilised gold nanoparticles. J. Mater. Chem. 2006, 16, 209–215. [Google Scholar] [CrossRef]
  80. De Lima, G.M.; Menezes, D.C.; Cavalcanti, C.A.; Dos Santos, J.A.; Ferreira, I.P.; Paniago, E.B.; Wardell, J.L.; Wardell, S.M.; Krambrock, K.; Mendes, I.C.; et al. Synthesis, characterisation and biological aspects of copper (II) dithiocarbamate complexes, [Cu {S2CNR (CH2CH2OH)} 2],(R= Me, Et, Pr and CH2CH2OH). J. Mol. Struct. 2011, 988, 1–8. [Google Scholar] [CrossRef] [Green Version]
  81. Gaudernak, E.; Seipelt, J.; Triendl, A.; Grassauer, A.; Kuechler, E. Antiviral effects of pyrrolidine dithiocarbamate on human rhinoviruses. J. Virol. 2002, 76, 6004–6015. [Google Scholar] [CrossRef] [Green Version]
  82. Zamora-Moreno, J.; Montiel-Palma, V. Versatile silylphosphine ligands for transition metal complexation. In Ligand; Intech Open: London, UK, 2018. [Google Scholar]
  83. Trindade, T.; O’Brien, P.; Zhang, X.M.; Motevalli, M. Synthesis of PbS nanocrystallites using a novel single molecule precursors approach: X-ray single-crystal structure of Pb (S2CNEtPri) 2. J. Mater. Chem. 1997, 7, 1011–1016. [Google Scholar] [CrossRef] [Green Version]
  84. Mandal, T.; Piburn, G.; Stavila, V.; Rusakova, I.; Ould-Ely, T.; Colson, A.C.; Whitmire, K.H. New mixed ligand single-source precursors for PbS nanoparticles and their solvothermal decomposition to anisotropic nano-and microstructures. Chem. Mater. 2011, 23, 4158–4169. [Google Scholar] [CrossRef]
  85. Salavati-Niasari, M.; Sobhani, A.; Davar, F. Synthesis of star-shaped PbS nanocrystals using single-source precursor. J. Alloys Compd. 2010, 507, 77–83. [Google Scholar] [CrossRef]
  86. Lewis, D.J.; Kevin, P.; Bakr, O.; Muryn, C.A.; Malik, M.A.; O’Brien, P. Routes to tin chalcogenide materials as thin films or nanoparticles: A potentially important class of semiconductor for sustainable solar energy conversion. Inorg. Chem. Front. 2014, 1, 577–598. [Google Scholar] [CrossRef] [Green Version]
  87. Malik, M.A.; Afzaal, M.; O’Brien, P. Precursor chemistry for main group elements in semiconducting materials. Chem. Rev. 2010, 110, 4417–4446. [Google Scholar] [CrossRef] [PubMed]
  88. Lewis, D.J.; Tedstone, A.A.; Zhong, X.L.; Lewis, E.A.; Rooney, A.; Savjani, N.; Brent, J.R.; Haigh, S.J.; Burke, M.G.; Muryn, C.A.; et al. Thin films of molybdenum disulfide doped with chromium by aerosol-assisted chemical vapor deposition (AACVD). Chem. Mater. 2015, 27, 1367–1374. [Google Scholar] [CrossRef] [Green Version]
  89. Sun, J.K.; Jiang, Y.; Zhong, X.; Hu, J.S.; Wan, L.J. Three-dimensional nanostructured electrodes for efficient quantum-dot-sensitized solar cells. Nano Energy 2017, 32, 130–156. [Google Scholar] [CrossRef]
  90. Kumar, S.; Nehra, M.; Deep, A.; Kedia, D.; Dilbaghi, N.; Kim, K.H. Quantum-sized nanomaterials for solar cell applications. Renew. Sustain. Energy Rev. 2017, 73, 821–839. [Google Scholar] [CrossRef]
  91. Yu, W.W.; Qu, L.; Guo, W.; Peng, X. Experimental determination of the extinction coefficient of CdTe, CdSe, and CdS nanocrystals. Chem. Mater. 2003, 15, 2854–2860. [Google Scholar] [CrossRef]
  92. Vogel, R.; Hoyer, P.; Weller, H. Quantum-sized PbS, CdS, Ag2S, Sb2S3, and Bi2S3 particles as sensitizers for various nanoporous wide-bandgap semiconductors. J. Phys. Chem. 1994, 98, 3183–3188. [Google Scholar] [CrossRef]
  93. Luther, J.M.; Beard, M.C.; Song, Q.; Law, M.; Ellingson, R.J.; Nozik, A.J. Multiple exciton generation in films of electronically coupled PbSe quantum dots. Nano Lett. 2007, 7, 1779–1784. [Google Scholar] [CrossRef]
  94. Yang, Z.; Chen, C.Y.; Roy, P.; Chang, H.T. Quantum dot-sensitized solar cells incorporating nanomaterials. Chem. Commun. 2011, 47, 9561–9571. [Google Scholar] [CrossRef]
  95. Zhao, K.; Pan, Z.; Mora-Seró, I.; Cánovas, E.; Wang, H.; Song, Y.; Gong, X.; Wang, J.; Bonn, M.; Bisquert, J.; et al. Boosting power conversion efficiencies of quantum-dot-sensitized solar cells beyond 8% by recombination control. J. Am. Chem. Soc. 2015, 137, 5602–5609. [Google Scholar] [CrossRef] [PubMed]
  96. Ren, Z.; Wang, Z.; Wang, R.; Pan, Z.; Gong, X.; Zhong, X. Effects of metal oxyhydroxide coatings on photoanode in quantum dot sensitized solar cells. Chem. Mater. 2016, 28, 2323–2330. [Google Scholar] [CrossRef]
  97. Du, Z.; Pan, Z.; Fabregat-Santiago, F.; Zhao, K.; Long, D.; Zhang, H.; Zhao, Y.; Zhong, X.; Yu, J.S.; Bisquert, J. Carbon counter-electrode-based quantum-dot-sensitized solar cells with certified efficiency exceeding 11%. J. Phys. Chem. Lett. 2016, 7, 3103–3111. [Google Scholar] [CrossRef] [PubMed]
  98. Jiao, S.; Du, J.; Du, Z.; Long, D.; Jiang, W.; Pan, Z.; Li, Y.; Zhong, X. Nitrogen-doped mesoporous carbons as counter electrodes in quantum dot sensitized solar cells with a conversion efficiency exceeding 12%. J. Phys. Chem. Lett. 2017, 8, 559–564. [Google Scholar] [CrossRef]
  99. Kowshik, M.; Vogel, W.; Urban, J.; Kulkarni, S.K.; Paknikar, K.M. Microbial synthesis of semiconductor PbS nanocrystallites. Adv. Mater. 2002, 14, 815–818. [Google Scholar] [CrossRef]
  100. Díaz, S.A.; Gillanders, F.; Jares-Erijman, E.A.; Jovin, T.M. Photoswitchable semiconductor nanocrystals with self-regulating photochromic Förster resonance energy transfer acceptors. Nat. Commun. 2015, 6, 6036. [Google Scholar] [CrossRef] [Green Version]
  101. Gong, X.; Yang, Z.; Walters, G.; Comin, R.; Ning, Z.; Beauregard, E.; Adinolfi, V.; Voznyy, O.; Sargent, E.H. Highly efficient quantum dot near-infrared light-emitting diodes. Nat. Photonics 2016, 10, 253. [Google Scholar] [CrossRef]
  102. Galian, R.E.; De la Guardia, M. The use of quantum dots in organic chemistry. Trends Anal. Chem. 2009, 28, 279–291. [Google Scholar] [CrossRef]
  103. Crisp, R.W.; Kroupa, D.M.; Marshall, A.R.; Miller, E.M.; Zhang, J.; Beard, M.C.; Luther, J.M. Metal halide solid-state surface treatment for high efficiency PbS and PbSe QD solar cells. Sci. Rep. 2015, 5, 9945. [Google Scholar] [CrossRef] [Green Version]
  104. Sasaki, A.; Tsukasaki, Y.; Komatsuzaki, A.; Sakata, T.; Yasuda, H.; Jin, T. Recombinant protein (EGFP-Protein G)-coated PbS quantum dots for in vitro and in vivo dual fluorescence (visible and second-NIR) imaging of breast tumors. Nanoscale 2015, 7, 5115–5119. [Google Scholar] [CrossRef] [Green Version]
  105. Kong, Y.; Chen, J.; Fang, H.; Heath, G.; Wo, Y.; Wang, W.; Li, Y.; Guo, Y.; Evans, S.D.; Chen, S.; et al. Highly fluorescent ribonuclease-A-encapsulated lead sulfide quantum dots for ultrasensitive fluorescence in vivo imaging in the second near-infrared window. Chem. Mater. 2016, 28, 3041–3050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Chen, J.; Kong, Y.; Feng, S.; Chen, C.; Wo, Y.; Wang, W.; Dong, Y.; Wu, Z.; Li, Y.; Chen, S. Recycled synthesis of whey-protein-capped lead sulfide quantum dots as the second near-infrared reporter for bioimaging application. ACS Sustain. Chem. Eng. 2016, 4, 2932–2938. [Google Scholar] [CrossRef]
  107. Freyria, F.S.; Cordero, J.M.; Caram, J.R.; Doria, S.; Dodin, A.; Chen, Y.; Willard, A.P.; Bawendi, M.G. Near-infrared quantum dot emission enhanced by stabilized self-assembled j-aggregate antennas. Nano Lett. 2017, 17, 7665–7674. [Google Scholar] [CrossRef] [Green Version]
  108. Zhao, P.; Xu, Q.; Tao, J.; Jin, Z.; Pan, Y.; Yu, C.; Yu, Z. Near infrared quantum dots in biomedical applications: Current status and future perspective. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2018, 10, 1483. [Google Scholar] [CrossRef]
  109. Yang, T.; Tang, Y.A.; Liu, L.; Lv, X.; Wang, Q.; Ke, H.; Deng, Y.; Yang, H.; Yang, X.; Liu, G.; et al. Size-dependent Ag2S nanodots for second near-infrared fluorescence/photoacoustics imaging and simultaneous photothermal therapy. ACS Nano 2017, 11, 1848–1857. [Google Scholar] [CrossRef] [PubMed]
  110. Xu, F.; Gerlein, L.; Ma, X.; Haughn, C.; Doty, M.; Cloutier, S. Impact of different surface ligands on the optical properties of PbS quantum dot solids. Materials 2015, 8, 1858–1870. [Google Scholar] [CrossRef] [Green Version]
  111. Thulasi-Varma, C.V.; Rao, S.S.; Ikkurthi, K.D.; Kim, S.K.; Kang, T.S.; Kim, H.J. Enhanced photovoltaic performance and morphological control of the PbS counter electrode grown on functionalized self-assembled nanocrystals for quantum-dot sensitized solar cells via cost-effective chemical bath deposition. J. Mater. Chem. C 2015, 3, 10195–10206. [Google Scholar] [CrossRef]
  112. Deepa, K.G.; Nagaraju, J. Development of SnS quantum dot solar cells by SILAR method. Mater. Sci. Semicond. Process. 2014, 27, 649–653. [Google Scholar] [CrossRef]
  113. Wang, W.; Shen, H.; Yao, H.; Li, J.; Jiao, J. Influence of solution temperature on the properties of Cu2 ZnSnS4 nanoparticles by ultrasound-assisted microwave irradiation. J. Mater. Sci. Mater. Electron. 2015, 26, 1449–1454. [Google Scholar] [CrossRef]
  114. Tang, P.; Chen, H.; Cao, F.; Pan, G.; Wang, K.; Xu, M.; Tong, Y. Nanoparticulate SnS as an efficient photocatalyst under visible-light irradiation. Mater. Lett. 2011, 65, 450–452. [Google Scholar] [CrossRef]
  115. Das, D.; Dutta, R.K. A novel method of synthesis of small band gap SnS nanorods and its efficient photocatalytic dye degradation. J. Colloid Interface Sci. 2015, 457, 339–344. [Google Scholar] [CrossRef] [PubMed]
  116. Koyasu, S.; Atarashi, D.; Sakai, E.; Miyauchi, M. Copper sulfide catalyzed porous fluorine-doped tin oxide counter electrode for quantum dot-sensitized solar cells with high fill factor. Int. J. Photoenergy 2017, 2017, 9. [Google Scholar] [CrossRef]
  117. Kim, H.J.; Ko, B.; Gopi, C.V.; Venkata-Haritha, M.; Lee, Y.S. Facile synthesis of morphology dependent CuS nanoparticle thin film as a highly efficient counter electrode for quantum dot-sensitized solar cells. J. Electroanal. Chem. 2017, 791, 95–102. [Google Scholar] [CrossRef]
  118. Punnoose, D.; Kumar, C.S.P.; Rao, S.S.; Varma, C.V.T.; Naresh, B.; Reddy, A.E.; Kundarala, N.; Lee, Y.S.; Kim, M.Y.; Kim, H.J. In situ synthesis of CuS nano platelets on nano wall networks of Ni foam and its application as an efficient counter electrode for quantum dot sensitized solar cells. Org. Electron. 2017, 42, 115–122. [Google Scholar] [CrossRef]
  119. Palve, B.M.; Kadam, V.S.; Jagtap, C.V.; Jadkar, S.R.; Pathan, H.M. A simple chemical route to synthesis the CuSe and CuS counter electrodes for titanium oxide based quantum dot solar cells. J. Mater. Sci. Mater. Electron. 2017, 28, 14394–14401. [Google Scholar] [CrossRef]
  120. Wang, X.; Wang, S.; Xie, Y.; Jiao, Y.; Liu, J.; Wang, X.; Zhou, W.; Xing, Z.; Pan, K. Facet-Dependent SnS Nanocrystals as the High-Performance Counter Electrode Materials for Dye-Sensitized Solar Cells. ACS Sustain. Chem. Eng. 2019, 7, 14353–14360. [Google Scholar] [CrossRef]
  121. Gopi, C.V.; Bae, J.H.; Venkata-Haritha, M.; Kim, S.K.; Lee, Y.S.; Sarat, G.; Kim, H.J. One-step synthesis of solution processed time-dependent highly efficient and stable PbS counter electrodes for quantum dot-sensitized solar cells. RSC Adv. 2015, 5, 107522–107532. [Google Scholar] [CrossRef]
  122. Lu, K.; Wang, Y.; Yuan, J.; Cui, Z.; Shi, G.; Shi, S.; Han, L.; Chen, S.; Zhang, Y.; Ling, X.; et al. Efficient PbS quantum dot solar cells employing a conventional structure. J. Mater. Chem. A 2017, 5, 23960–23966. [Google Scholar] [CrossRef]
  123. Ahmet, I.Y.; Guc, M.; Sánchez, Y.; Neuschitzer, M.; Izquierdo-Roca, V.; Saucedo, E.; Johnson, A.L. Evaluation of AA-CVD deposited phase pure polymorphs of SnS for thin films solar cells. RSC Adv. 2019, 9, 14899–14909. [Google Scholar] [CrossRef] [Green Version]
  124. Mohammadnezhad, M.; Selopal, G.S.; Alsayyari, N.; Akilimali, R.; Navarro-Pardo, F.; Wang, Z.M.; Stansfield, B.; Zhao, H.; Rosei, F. CuS/graphene nanocomposite as a transparent conducting oxide and pt-free counter electrode for dye-sensitized solar cells. J. Electrochem. Soc. 2019, 166, 3065–3073. [Google Scholar] [CrossRef]
  125. Dou, Y.; Zhou, R.; Wan, L.; Niu, H.; Zhou, J.; Xu, J.; Cao, G. Nearly monodisperse PbS quantum dots for highly efficient solar cells: An in situ seeded ion exchange approach. Chem. Commun. 2018, 54, 12598–12601. [Google Scholar] [CrossRef] [PubMed]
  126. Kwon, O.H.; Heo, J.H.; Park, S.; Kim, S.W.; Im, S.H. High performance solid-state PbS/CuS hetero-nanostructured quantum dots-sensitized solar cells. J. Ind. Eng. Chem. 2019, 75, 164–170. [Google Scholar] [CrossRef]
  127. Cheraghizade, M.; Jamali-Sheini, F.; Yousefi, R.; Niknia, F.; Mahmoudian, M.R.; Sookhakian, M. The effect of tin sulfide quantum dots size on photocatalytic and photovoltaic performance. Mater. Chem. Phys. 2017, 195, 187–194. [Google Scholar] [CrossRef]
  128. Pei, Q.; Chen, Z.; Wang, S.; Zhang, D.; Ma, P.; Li, S.; Zhou, X.; Lin, Y. PbS decorated multi-walled carbon nanotube/Ti mesh films as efficient counter electrodes for quantum dots sensitized solar cells. Sol. Energy 2019, 178, 108–113. [Google Scholar] [CrossRef]
  129. Lu, K.; Wang, Y.; Liu, Z.; Han, L.; Shi, G.; Fang, H.; Chen, J.; Ye, X.; Chen, S.; Yang, F.; et al. High-Efficiency PbS Quantum-Dot Solar Cells with Greatly Simplified Fabrication Processing via “Solvent-Curing”. Adv. Mater. 2018, 30, 1707572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Sunesh, C.D.; Gopi, C.V.; Muthalif, M.P.A.; Kim, H.J.; Choe, Y. Improving the efficiency of quantum-dot-sensitized solar cells by optimizing the growth time of the CuS counter electrode. Appl. Surf. Sci. 2017, 416, 446–453. [Google Scholar] [CrossRef]
Figure 1. Research outputs from 2009 to 2018 on quantum dot sensitized solar cells (QDSSCs) (reproduced from ref. [6] with permission from the American Chemical Society).
Figure 1. Research outputs from 2009 to 2018 on quantum dot sensitized solar cells (QDSSCs) (reproduced from ref. [6] with permission from the American Chemical Society).
Molecules 24 04223 g001
Figure 2. Schematic illustration of the QDSSCs working procedure involving the quantum dot (QD) photosensitizer, counter electrode, photoanode and electrolyte.
Figure 2. Schematic illustration of the QDSSCs working procedure involving the quantum dot (QD) photosensitizer, counter electrode, photoanode and electrolyte.
Molecules 24 04223 g002
Figure 3. Morphology images of (a) octahedral Cu2O crystals, (b) octahedral CuxS cages, (c) TEM octahedral and (d) the SAED pattern of octahedral (reproduced from ref. [41] with permission from the American Chemical Society); (e,f) are for the sample prepared using 10 mM solution of the precursor at 200 ℃ for 1h showing flat and standing hexagonal nanodisks and (g) SAED. (h) TEM images for the sample prepared using 20 mM solution of the precursor at 280 ℃ (reproduced from ref. [42] with permission from the American Chemical Society).
Figure 3. Morphology images of (a) octahedral Cu2O crystals, (b) octahedral CuxS cages, (c) TEM octahedral and (d) the SAED pattern of octahedral (reproduced from ref. [41] with permission from the American Chemical Society); (e,f) are for the sample prepared using 10 mM solution of the precursor at 200 ℃ for 1h showing flat and standing hexagonal nanodisks and (g) SAED. (h) TEM images for the sample prepared using 20 mM solution of the precursor at 280 ℃ (reproduced from ref. [42] with permission from the American Chemical Society).
Molecules 24 04223 g003
Figure 4. UV and PL absorption properties of PbS nanocrystals (a,b) (reproduced from ref. [48] with permission from the American Chemical Society).
Figure 4. UV and PL absorption properties of PbS nanocrystals (a,b) (reproduced from ref. [48] with permission from the American Chemical Society).
Molecules 24 04223 g004
Figure 5. PbS QDs (a) absorption spectra. (b) Comparison of the first exciton peak. (c) Representative J−V curves of QDSSC with PbS QDs. (d) PCE values (black legend) as a function of QDs optical εgap. (e) Peak EQE values measured at a 400 nm wavelength. (f) HRTEM images (reproduced from ref. [61] with permission from the American Chemical Society).
Figure 5. PbS QDs (a) absorption spectra. (b) Comparison of the first exciton peak. (c) Representative J−V curves of QDSSC with PbS QDs. (d) PCE values (black legend) as a function of QDs optical εgap. (e) Peak EQE values measured at a 400 nm wavelength. (f) HRTEM images (reproduced from ref. [61] with permission from the American Chemical Society).
Molecules 24 04223 g005
Figure 6. HRTEM images and SAED patterns of CuS-PVP nanoparticles (NPs) (a) and (c) and QDs (b) and (d). HRTEM image of CuS-PVP QDs showing abrupt edges (e). Dislocations, twins, and stacking faults are illustrated in (f). A twin region formed between the cubic and hexagonal phases of CuS is illustrated in (g) and (h). The stacking sequence is determined as ABCABABCAB in the direction indicated by the arrow (reproduced from ref. [69] with permission from the American Chemical Society).
Figure 6. HRTEM images and SAED patterns of CuS-PVP nanoparticles (NPs) (a) and (c) and QDs (b) and (d). HRTEM image of CuS-PVP QDs showing abrupt edges (e). Dislocations, twins, and stacking faults are illustrated in (f). A twin region formed between the cubic and hexagonal phases of CuS is illustrated in (g) and (h). The stacking sequence is determined as ABCABABCAB in the direction indicated by the arrow (reproduced from ref. [69] with permission from the American Chemical Society).
Molecules 24 04223 g006
Figure 7. (a) TEM image of star-shaped PbS nanocrystals synthesized at 230 °C. (Inset) HRTEM image of lattice fringes with zone axis of (111). (bd) TEM images and electron diffraction patterns with zone axis of (b) (110), (c) (112) and (d) (100), respectively. (e) HRTEM image of zoomed fringes with zone axis of (100). (fh) Dodecanethiol (DT) capping molecular system. The ratio of precursor to capping molecules is (a) 1/400, (b) 1/100 and (c) 1/50, respectively. (i) Dodecylamine (DDA) capping molecular system with the ratio of precursor to capping molecule of 1/100. (j) HRTEM image of a round truncated octahedron with zone axis of (110), and (k) a truncated octahedron with zone axis of (100). (l) The R factor depends on the relative ratio of precursor to capping molecules (reproduced from ref. [63] with permission from the American Chemical Society).
Figure 7. (a) TEM image of star-shaped PbS nanocrystals synthesized at 230 °C. (Inset) HRTEM image of lattice fringes with zone axis of (111). (bd) TEM images and electron diffraction patterns with zone axis of (b) (110), (c) (112) and (d) (100), respectively. (e) HRTEM image of zoomed fringes with zone axis of (100). (fh) Dodecanethiol (DT) capping molecular system. The ratio of precursor to capping molecules is (a) 1/400, (b) 1/100 and (c) 1/50, respectively. (i) Dodecylamine (DDA) capping molecular system with the ratio of precursor to capping molecule of 1/100. (j) HRTEM image of a round truncated octahedron with zone axis of (110), and (k) a truncated octahedron with zone axis of (100). (l) The R factor depends on the relative ratio of precursor to capping molecules (reproduced from ref. [63] with permission from the American Chemical Society).
Molecules 24 04223 g007
Figure 8. (ac) PbS CEs SEM images, (d) J–V, (e) Tafel and (f) Nyquist plots (reproduced from ref. [111] with permission from the American Chemical Society).
Figure 8. (ac) PbS CEs SEM images, (d) J–V, (e) Tafel and (f) Nyquist plots (reproduced from ref. [111] with permission from the American Chemical Society).
Molecules 24 04223 g008
Figure 9. (J–V) of TiO2/CdS/CdSe/ZnS QDSSCs with CuS and Pt CEs (reproduced from ref. [130] with permission from the Elsevier).
Figure 9. (J–V) of TiO2/CdS/CdSe/ZnS QDSSCs with CuS and Pt CEs (reproduced from ref. [130] with permission from the Elsevier).
Molecules 24 04223 g009
Table 1. J–V properties of PbS, CuS and SnS QDSSCs with various photoanodes and CEs.
Table 1. J–V properties of PbS, CuS and SnS QDSSCs with various photoanodes and CEs.
Quantum DotsPhotoanodeElectrolyteCEsECE (%)YearRef.
CuSTiO2/CdS/CdSe/ZnSS2−/Sx2−Pt3.182015[29]
Cu2S-CBD/N719TiO24-tert-butylpyridinePt8.342016[30]
CuInSe/ZnSTiO2S2−/Sx2−Cu2S8.102016[31]
PbS QDsTiO2S2−/Sx2−Cu2S2.242016[32]
PbS/CdS/ZnSTiO2S2−/Sx2−Cu2S7.192016[33]
PbSTiO2S2−/Sx2−Cu2S1.302018[34]
PbS CQD ZnOTBAICarbon3.62017[35]
SnS QDsTiO24-tert-butylpyridinePt0.3262015[36]
SnSTiO24-tert-butylpyridineAu0.01022014[37]
SnS QDsTiO2S2−/Sx2−Pt0.0722019[38]
Table 2. Photovoltaic parameters of PbS, CuS and SnS solar cells.
Table 2. Photovoltaic parameters of PbS, CuS and SnS solar cells.
CellVOC (mV)JSC (mA/cm2)FF (%)H (%)RS (kΩ cm2)RSH (kΩ cm2)nJ0 × 10−7 (mA/cm2)YearRef.
CuS5/FTO5 μm0.375.81521.1213.626.2--2017[116]
CuS 2h0.61215.520.4524.297.651.681.332.6762017[117]
CuS/NF0.6117.80.544.937.47.262.1-2017[118]
CuS0.580.71481.38----2017[119]
CuS6889.37674.31----2019[120]
PbS900.64412.170.5884.61 7.561.450.5 5.01 2015[121]
ITO/PbS-TBAI/Al0.45 17.86 52.0 4.18----2017[122]
PbS 2 h0.56011.200.553.488.582.6 1.213.042015[123]
PbS49517.460.6395.52----2019[124]
PbS/CuS/HNS QD0.6123.864.49.3----2019[125]
PbS QD0.5625.0167.39.43----2018[111]
SnS0.391.70.722.36----2017[126]
SnS QDs3427838.20.0102----2014[44]
Mo/SnS/CdS/i-ZnO/ITO29017.2562.8----2019[127]
SnS-S74618.090.678.96----2019[128]
FTO/πSnS/CdS/i-ZnO/ITO1133.40420.15----2018[129]

Share and Cite

MDPI and ACS Style

Meyer, E.L.; Mbese, J.Z.; Agoro, M.A. The Frontiers of Nanomaterials (SnS, PbS and CuS) for Dye-Sensitized Solar Cell Applications: An Exciting New Infrared Material. Molecules 2019, 24, 4223. https://doi.org/10.3390/molecules24234223

AMA Style

Meyer EL, Mbese JZ, Agoro MA. The Frontiers of Nanomaterials (SnS, PbS and CuS) for Dye-Sensitized Solar Cell Applications: An Exciting New Infrared Material. Molecules. 2019; 24(23):4223. https://doi.org/10.3390/molecules24234223

Chicago/Turabian Style

Meyer, Edson L., Johannes Z. Mbese, and Mojeed A. Agoro. 2019. "The Frontiers of Nanomaterials (SnS, PbS and CuS) for Dye-Sensitized Solar Cell Applications: An Exciting New Infrared Material" Molecules 24, no. 23: 4223. https://doi.org/10.3390/molecules24234223

Article Metrics

Back to TopTop