Elsevier

Thin Solid Films

Volume 714, 30 November 2020, 138393
Thin Solid Films

Ligand field states and defect levels synergism: A close look at the band alignment of 4T1‑Mn-CdS/Bi2S3-co-sensitized photoanodes

https://doi.org/10.1016/j.tsf.2020.138393Get rights and content

Highlights

  • Mn2+ 4T1 and Cd-Cd levels are formed into Mn-CdS after chalcogenide preparation.

  • Cd–Cd→Mn2+ 4T1 carrier transitions improve the light-harvesting of the photoanodes.

  • The load of Mn-CdS QDs impacts on the charge transport ability of the photoanodes.

  • Mn2+ 4T1 states provide a suitable band alignment in the composite heterostructure.

  • The Cd–Cd→Mn2+ 4T1 interaction increases the electron lifetime into photoanodes.

Abstract

Low-band gap chalcogenides such as Bi2S3 (<1.7 eV) have been widely used in optoelectronic devices such as quantum dot sensitized solar cells (QDSSCs), due to their high sunlight harvesting capability, absorbing low-energy photons close to the IR region. Nonetheless, Bi2S3 offers a poor band alignment with large-band gap semiconductors such as TiO2, achieving low photoconversion efficiencies. Accordingly, we studied how the presence of both Mn2+ 4T1 ligand field electronic states and structural defects as Cd–Cd energy levels produced during Mn-CdS synthesis influenced on the band structure and thereby, the charge carrier transport into co-sensitized boron, nitrogen and fluorine-co-doped TiO2 nanotubes (X–Mn–Y–CdS–Bi2S3). Carrier transfer pathways provided by both Cd–Cd defects and Mn2+ 4T1 states allowed to obtain a suitable 0.7–Mn–4–CdS–2–Bi2S3 based electrode, with a narrowed band gap of 2.16 eV and an appropriate II-type heterostructure. These features improved both the carrier separation and mobility from Mn-CdS/Bi2S3 interface to co-doped nanotubes. Additionally, the electron lifetime into the composite photoanode was 10 times higher compared with a Mn2+-absent 4–CdS–2–Bi2S3 material. Hence, the synergistic behavior between structural defects and ligand field electronic states explained here offers an insight for establishing adequate heterostructures to be useful in QDSSCs.

Introduction

The photoconversion efficiency (PCE) has been the most remarkable feature for obtaining promising solar cells during the last years. However, the actual research is focused on the improvement of the optical properties of several photoanode materials, and the decrease of charge carrier recombination into the devices [1]. By controlling these characteristics, a high energy conversion can be achieved. The latter is the case of quantum dot sensitized solar cells (QDSSCs), which have reached the PCE of their equivalent dye-sensitized solar cells around 11% [2]. As soon as the photophysical and electrical properties of sensitizers were progressively understood, the PCE was increased over 16.6% [3]. Advantages of narrow band gap-semiconductor based quantum dots (QDs) such as generation of multiple electron-hole pairs, high molar extinction coefficient, and tunable band structures have upgraded the sunlight harvesting capability and the carrier transport into solar cells [3, 4]. Despite of these improvements, some semiconductors do not show one of the latter two abilities, producing the eventual lowering of PCE from the devices, and the loss of interest from the industry. Thus, it has been explored low-cost and nontoxic (or low-toxicity) chalcogenides such as Sb2×3 (X = S, Se, S/Se), CuInSe2, Zn−Cu−In−Se alloys, among others, because these materials exhibit attractive optical properties (specially, a low band gap between 1 and 1.7 eV), providing solar devices with PCE values comparable with their Pb-counterparts [5, 6]. In this context, within the most attractive chalcogenide materials, we highlight bismuth sulfide (Bi2S3). This low-toxic chalcogenide displays a low band gap (Eg) around 1.2–1.7 eV, meaning a high visible-light absorption [7, 8]. Unfortunately, Bi2S3-based solar cells using TiO2 based photoanodes have produced PCEs lower than 1.0%, caused mainly for a poor band alignment in the TiO2/Bi2S3 interface [9], [10], [11], [12]. Therefore, the co-sensitization has been reported to be efficient for inducing a suitable band structure between binary sensitizers such as CdS and CdSe [13,14]. For instance, the introduction of CdS QDs has been widely studied as a suitable strategy to improve the light harvesting of QDSSCs in the visible region, with external quantum efficiencies near to 90% and increase the electron injection to reach the back contact, besides of its facile preparation [15], [16], [17], [18]. However, some reports have also demonstrated that the use of CdS mixed with other sensitizers can also provide an energy barrier for charge carrier separation and mobility, achieving low PCEs [19], [20], [21].

An alternative to overcome the latter issue is the incorporation of new electronic states into the binary CdS coming from Mn-doping. Interestingly, Mn2+ electronic states are considered as efficient electron storage sites, where their carrier transfer mechanism can be explained according to the ligand field theory [22, 23]. Mn2+ cation shows a coordination number of 4 into chalcogenide rich-ambients (e.g. S2−, Se2−), generating a weak tetrahedral field. In this context, 4T16A1 ligand field emission transitions into Mn2+ can be achieved. This type of d-d transitions is known to be spin and orbitally forbidden, resulting in a photoexcited electron lifetime around milliseconds [22], [23], [24]. Additionally, these d-d transitions are located into the band gap of CdS, exhibiting a low energy gap around 2.1 eV [25]. This feature has been revealed to red-shift the Eg of CdS [22], which can explain the enhancement of its charge carrier separation and electron transfer dynamics [4,[26], [27], [28], [29]]. However, an excess of this cation also provides non-radiative recombination sites and a poor band alignment into the composites [22,26,30]. Therefore, the amount of Mn2+ into CdS plays an important role in the photovoltaic performance of the QDSSCs. The density of Mn2+ can be easily adjusted by controlling the Mn precursor through low-cost successive ionic layer adsorption and reaction (SILAR), using alcoholic solutions during the formation of Mn–doped CdS QDs. Furthermore, an adequate amount of Mn2+ into inner CdS can increase electron lifetime into the composite material, preventing charge carrier recombination and loss of PCE in QDSSCs [31].

The generation of photocurrent into Mn–CdS sensitized solar cells has been studied in function of the density of Mn2+, achieving higher PCE values than that of CdS-devices [27,29,31]. Nonetheless, the literature is limited about how the Mn2+ligand field electronic states such as 4T1 or/and 6A1 impact on the carrier transfer ability of Mn–CdS. It seems that the presence of Mn2+ is the only key factor to dictate the carrier mobility. But the CdS QDs synthesis itself encloses the formation of hidden energy levels that mitigate the carrier dynamics into photoanodes [32,33]. Maybe, this is one of the main aspects that have not allowed to reach high PCEs in CdS-QDSSCs. By considering some parameters as the concentration of cationic/anionic solutions, washing process and the number of deposition cycles during SILAR process, CdS can form structural defects known as Cd-Cd energy levels. These states are produced by direct interaction between two adjacent Cd atoms caused by sulfur vacancies into CdS lattice. The sulfur release transfers two electrons to Cd atoms for establishing a Cd-Cd bond through the Cd sp2 hybridization [32]. In terms of energy, Cd-Cd states are located above to the valence band (VB) of CdS, lowering its Eg [32,33]. Cd-Cd energy levels have been early reported to impact the photovoltaic properties of CdS/CdS1-xSex co-sensitized photoanodes [33]. Clearly, if the influence of the Cd-Cd energy levels has not been deeply investigated, their interaction with Mn2+ electronic states for improving the carrier transport/separation into heterostructures has not been revealed yet. By using photoelectrochemistry to analyze both the light harvesting and charge carrier transport abilities of composites, we can observe how the synergism between Cd-Cd energy levels and Mn2+ ligand field electronic states (1) facilitates the band engineering to produce a more suitable band structure using unmatched sensitizers, and (2) promotes carrier flow in order to achieve a high photocurrent. Thus, we consider that the photovoltaic performance of low-band gap based QDSSCs could be improved.

In this work, we studied the impact of the synergism between Mn2+ 4T1 electronic states and Cd-Cd defects on the photophysical and photo(electro)chemical properties of Mn–CdS/Bi2S3 co-sensitized boron, nitrogen and fluorine-co-doped TiO2 nanotubes (X–Mn–Y–CdS–Bi2S3). For understating the band structure and carrier transfer processes. Materials were modified by the presence of ligand field electronic states and structural defects, an efficient carrier transport pathway, high electron lifetime and photocurrent were achieved into X–Mn–Y–CdS–Bi2S3 composite photoanodes. This contribution opens the door to visualize more clearly the carrier transfer mechanisms occurring into composite photoanodes that limit their potential application in QDSSCs.

Section snippets

Fabrication of BNF–TNT arrays

B-, N- F-co-doped TiO2 nanotubes (BNF-TNT) were grown on Ti foils according to an established procedure [33,34]. Briefly, Ti anodization was performed in an electrolyte solution composed by 0.06 wt% H3BO3, 0.45 wt% NH4F and 2.0 wt% deionized (DI) water in ethylene glycol, applying a bias potential of 60 V for 2.5 h. A two-electrode cell, using a Cu foil as the cathode, were used for the material preparation. The as-prepared BNF-TNT were rinsed with DI water, dried at 100 °C and annealed at

Morphological and physicochemical characterization of composite photoanodes

FESEM images of the BNF–TNT, 0.7–Mn–4–CdS and 0.7–Mn-4–CdS–2–Bi2S3 photoanodes were obtained, where it was observed the top-view and cross section of the materials surface. BNF–TNT were vertically oriented, showing cavities with an average diameter around 116.6 ± 2.6 nm (Fig. 1a) and an average tube length around 16.7 ± 1.0 nm (Fig. 1a’). The width of the cavities was decreased after the deposition of sensitizer nanoparticles, achieving an increase of the wall thickness in the tubular

Conclusions

The incorporation of Mn2+ into CdS for synthesizing Mn-CdS provided the improvement of the light harvesting and electron transport ability into X–Mn–Y–CdS–2–Bi2S3 composites. This fact was caused by the presence of both structural defects such as Cd-Cd energy levels and ligand field electronic states such as Mn2+ 4T1 into the Mn-CdS QDs. In this context, the photophysical and photoelectrochemical properties of composites were modified, altering their band structure. The formation of an adequate

CRediT authorship contribution statement

Andrés F. Gualdrón-Reyes: Conceptualization, Methodology, Data curation, Formal analysis, Investigation, Visualization, Writing - original draft. Johan S. Ríos-Niño: Methodology, Data curation, Investigation. Angel M. Meléndez: Conceptualization, Supervision, Investigation, Validation, Writing - review & editing. Jhonatan Rodríguez-Pereira: Methodology, Data curation, Investigation, Validation. Mario Alejandro Mejía-Escobar: Methodology, Investigation, Writing - review & editing. Franklin

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

Authors are graceful with Universidad Industrial de Santander-Colciencias project 110265843664 (VIE 8836). Andrés F. Gualdrón-Reyes acknowledges to COLCIENCIAS for the Ph.D. 617-scholarship (Doctorado Nacional COLCIENCIAS).

References (66)

  • C. Chen et al.

    Double-sided transparent TiO2 nanotube/ITO electrodes for efficient CdS/CuInS2 quantum dot-sensitized solar cells

    Nanoscale Res. Lett.

    (2017)
  • C. Chen et al.

    Photocurrent enhancement of the CdS/TiO2/ITO photoelectrodes achieved by controlling the deposition amount of Ag2S nanocrystals

    Appl. Surf. Sci.

    (2015)
  • D. Esparza et al.

    Photovoltaic Properties of Bi2S3 and CdS Quantum Dot Sensitized TiO2 Solar Cells

    Electrochim. Acta

    (2015)
  • J.L. Qiao et al.

    Enhancing photoelectrochemical performance of TiO2 nanotube arrays by CdS and Bi2S3 co-sensitization

    J. Photochem. Photobiol. A

    (2016)
  • I. Mehmood et al.

    Mn-doped CdS passivated CuInSe2 quantum dot sensitized solar cells with remarkably enhanced photovoltaic efficiency

    RSC Adv.

    (2017)
  • Z. Du et al.

    Carbon counter-electrode-based quantum-dot-sensitized solar cells with certified efficiency exceeding 11%

    J. Phys. Chem. Lett.

    (2016)
  • M. Hao et al.

    Ligand-assisted cation-exchange engineering for high-efficiency colloidal Cs1−xFAxPbI3 quantum dot solar cells with reduced phase segregation

    Nat. Energy

    (2020)
  • J. Wang et al.

    Mn doped quantum dots sensitized solar cells with power conversion efficiency exceeding 9%

    J. Mater. Chem. A

    (2016)
  • C. hen et al.

    Open-circuit voltage loss of antimony chalcogenide solar cells: status, origin and possible solution

    ACS Energy Lett.

    (2020)
  • J. Du et al.

    Zn−Cu−In−Se quantum dot solar cells with a certified power conversion efficiency of 11.6%

    J. Am. Chem. Soc.

    (2016)
  • F. Guo et al.

    Flowerlike Bi2S3 microspheres: facile synthesis and application in the catalytic reduction of 4-nitroaniline

    New J. Chem.

    (2014)
  • J.L.T. Chen, V. Nalla, G. Kannaiyan, V. Mamidala, W. Ji, J.J. Vittal. Synthesis and nonlinear optical switching of...
  • Y.C. Lin et al.

    Bi2S3 liquid-junction semiconductor-sensitized SnO2 solar cells

    J. Electrochem. Soc.

    (2014)
  • A.N. Kulkarni et al.

    TiO2 Photoanode sensitized with nanocrystalline Bi2S3: the effect of sensitization time and annealing on its photovoltaic performance

    Appl. Nanosci.

    (2016)
  • I. Zumeta-Dubé et al.

    TiO2 sensitization with Bi2S3 quantum dots: the inconvenience of sodium ions in the deposition procedure

    J. Phys. Chem. C

    (2014)
  • R. Zhou et al.

    Influence of deposition strategies on CdSe quantum dot-sensitized solar cells: a comparison between successive ionic layer adsorption and reaction and chemical bath deposition

    J. Mater. Chem. A

    (2015)
  • F. Huang et al.

    High efficiency CdS/CdSe quantum dot sensitized solar cells with two ZnSe layers

    ACS Appl. Mater. Interfaces

    (2016)
  • C. Rosiles-Pereza et al.

    Improved performance of CdS quantum dot sensitized solar cell by solvent modified SILAR approach

    Sol. Energy

    (2018)
  • S. Hachiya et al.

    Dependences of the optical absorption and photovoltaic properties of CdS quantum dot-sensitized solar cells on the CdS quantum dot adsorption time

    J. Appl. Phys.

    (2011)
  • D. Esparza et al.

    Effect of different sensitization technique on the photoconversion efficiency of CdS quantum dot and CdSe quantum rod sensitized TiO2 solar cells

    J. Phys. Chem. C

    (2015)
  • Z. Chen et al.

    Band alignment by ternary crystalline potential tuning interlayer for efficient electron injection in quantum dot-sensitized solar cells

    J. Mater. Chem. A

    (2014)
  • S. Naskar et al.

    Synthesis of ternary and quaternary Au and Pt decorated CdSe/CdS heteronanoplatelets with controllable morphology

    Adv. Funct. Mater.

    (2017)
  • P. Wang et al.

    The important role of surface ligand on CdSe/CdS core/shell nanocrystals in affecting the efficiency of H2 photogeneration from water

    Nanoscale

    (2015)
  • View full text