Open access peer-reviewed chapter

Two-Dimensional Transition Metal Dichalcogenide as Electron Transport Layer of Perovskite Solar Cells

Written By

Akrajas Ali Umar, Nurul Ain Abd Malek, Nabilah Alias and Abang Anuar Ehsan

Submitted: 23 December 2021 Reviewed: 22 February 2022 Published: 29 April 2022

DOI: 10.5772/intechopen.103854

From the Edited Volume

Chalcogenides - Preparation and Applications

Edited by Dhanasekaran Vikraman

Chapter metrics overview

187 Chapter Downloads

View Full Metrics

Abstract

Conventional perovskite solar cells utilize a combination of a compact and mesoporous layer of TiO2 or SnO2 as the electron transport layer. This structure is vulnerable to massive loss of photogenerated carriers due to grain boundary resistance in the layer. In this chapter, we will discuss a potential electron transport layer that might drive higher power conversion efficiency, i.e., thin and single-crystalline 2D transition metal dichalcogenide. Because of their ultimate thin structure, they facilitate rapid electron transport and enhanced carrier extraction in the solar cells device. We will also discuss the current state of the art of 2D transition metal dichalcogenide atomic layer application as an electron transport layer in the perovskite solar cells as well as our recent attempt in this field.

Keywords

  • 2D atomic layer
  • transition metal dichalcogenide
  • electron transport layer
  • perovskite solar cells

1. Introduction

Perovskite solar cells (PSCs) have received a great deal of attention in the past few decades due to their impressively high power conversion efficiency (PCE) [1]. To date, PCE as high as 25.6% has been successfully recorded. This performance has already been compared with the single-crystalline silicon solar cells system. With the advancement in the perovskite properties control, including the crystallinity properties, grain size, and stability properties, further improvement in the PCE is expected to be achieved soon. The continuous growth in the preparation of the high-performance charge selective layer in the perovskite solar cells further contributes to the rapid progress in the PCE improvement of the PSC [2].

Along with the transparent conducting electrode (TCE) and the top metal contact, a PSC device is composed of an electron-transport layer (ETL), an organometal-halide perovskite active layer, and a hole-transport layer (HTL). In these solar cells, the perovskite and its photoelectrical properties are the keys to the overall photovoltaic process. Its unique high-optical absorption constant drives massive photon absorption and exciton generation in the device. Despite this key fact, the carrier transport and interfacial charge transfer dynamics play another crucial factor for the generation of the overall PSC performance. These two parameters depend on the nature of the surface and the crystallinity properties of the charge-selective layers [3].

One of the serious problems in perovskite solar cell devices is the loss of charge carriers during the transport process in the carrier layer. This is because, the carrier layer has low crystallization, high grain boundary resistance as well as experiences loss of carrier charge during extraction to the outer electrode. The main factor of carrier charge lost during extraction to the outer electrode is due to the high interface resistance between the electrode and the carrier layer. Therefore, it is expected that when a carrier layer that has high crystallinity, very low thickness, and good coupling conditions with external electrodes is used, then the performance of the device will increase.

The electron transport layer (ETL), for example, TiO2, and other semiconducting oxides, such as SnO2, ZnO, have been widely applied in the perovskite solar cells fabrication. Despite the excellent performance demonstrated by them, this ETL suffers from large-density surface defects related to oxygen vacancy, particularly in the TiO2 system. The defect from such vacancy causes immense trap-limited (Shockley-Read-Hull) transport in the extraction of the photogenerated carrier to the external electrode. This in many cases degrades the photovoltaic performance of the PSC up to a certain degree, reducing the power conversion efficiency of the device. Even though there exist several methods in the passivation of such defects, such as acid passivation, etc., the improvement is minute. In addition, this method may add additional resistance to the photocarrier transport reducing the power conversion efficiency. Along with these crucial factors, the crystallinity properties of the ETL add an additional issue to the photocarrier transport dynamic in the device. As normal in the high-performance PSC fabrication, mesoporous TiO2 or SnO2 was used as ETL along with a compact layer of TiO2 or SnO2 (See Figure 1), [4]. As the figure reveals, the mesoporous layer is composed of a large number of interconnected small grain particles that produce grain boundary resistance due to lattice mismatch among the connected particles. This resistance should be massive due to their large-scale existence on the layer. This certainly complicates the transport of photogenerated electrons to the electrode layer, such as high internal resistance or radiationless recombination [5, 6]. Therefore, the selection of the right material for the carrier layer is important in determining the performance of a device. Such resistance boundary further augments the presence of mesoporous-compact layer interface resistance in the ETL system of the PSC. From this picture, we can estimate the loss would be suffered by the device during the photovoltaic process. This means that if such ETL is replaced with the single-crystalline ETL system, the performance of the perovskite solar cells can be improved.

Figure 1.

Mesoporous TiO2 ETL. (A and C) Top and side view of mesoporous TiO2 layer on compact layer TiO2. (B and D) Top and side of mesoporous TiO2 layer. (Reprinted from [4]. © 2017 American Chemical Society).

Advertisement

2. Two-dimensional transition metals ETL

Recently, materials of two-dimensional (2D) dichalcogenide transition metals (TMDs), such as MoS2, WS2, TiS2, CdTe, and others, have been used as carrier layers in perovskite solar cells due to their high charge carrier mobility, unique optoelectrical properties, large exciton binding energy, very fast interface charge transfer properties as well as excellent physical and chemical stability properties [7]. Their optoelectronic properties were also found to correspond to the number of layers, dopants, and strains (straining). The phenomenon of the massive charge transfer process in these van der Waals crystals driven by the collective motion of excitonic surfaces enables a high interfacial charge extraction and reduces charge recombination for an effective photovoltaic process [8]. One of the uniqueness of the TMDs layer is that it has an atomic-scale thickness (very thin) and has high crystallinity. With its planar (2D) structure, it will produce a strong coupling when grown on the electrode surface. Therefore, it has great potential for a carrier layer in perovskite solar cells.

Transition metal dichalcogenide (TMD) has the chemical formula of MX2 where M is the transition metal from groups 4 to 10 in the periodic table system, and X is a chalcogen atom such as sulfur (S), selenium (Se), or tellurium (Te). Figure 2 shows the typical structure of TMD. The structure has two layers of chalcogen that clamp a transition metal layer making this material have its uniqueness in electronic, optoelectronic properties, and chemical stability [10]. The electronic and optical properties of TMDs materials change significantly depending on the number of layers. For example, the MoS2 band gap increases from 1.29 eV (multilayered MoS2) to 1.59 eV (monolayer MoS2), and also this bandgap changes from an indirect bandgap to a direct bandgap as the number of layers decreases [11].

Figure 2.

Typical structure of transition metal dichalcogenide materials. (A) Typical layer stacking structure in bulk transition metal dichalcogenide structure. T and X represent the transition metal and chalcogen elements, respectively. (B) Top and side view of single-layer of TMD with 2H-phase. (C) Side view of single-layer TMD with 1T-phase. (Reprinted from [9]. © 2020, The Author(s)).

As is well known, most of these 2D TMD materials have ambipolar properties that enable the materials to transport both electrons and holes [12]. In other words, this allows 2D TMDs material to be used as ETL or HTL in n-i-p or p-i-n perovskite solar cells. However, most perovskite solar cell applications use these 2D TMD materials as HTL. Only MoS2 and TiS2 have been used as ETLs and have successfully produced efficiencies as high as 13.14% and 18.79% [7, 13]. Table 1 shows several PSC device structures utilizing TMD as ETL. Recently, there was a first simulation study on the photoelectric properties of WS2 as an ETL in perovskite solar cells reported with efficiencies as high as 25.70% [23]. By having high electron mobility as well as energy levels appropriate to the perovskite layer, the WS2 atomic layer is expected to function as an ETL capable of producing high-performance perovskite solar cell devices.

MaterialDevice structureJsc (mA cm−2)Voc (V)FFPCE (%)Ref.
TiS2FTO/TiS2/MAPbI3/spiro-OMeTAD/Au23.381.050.7117.37[14]
TiS2ITO/TiS2/ FAxMA1-xBrxClyI1-x-y/spiro-OMeTAD/Ag24.681.000.7518.79[7]
MoS2FTO/MoS2/MAPbI3/spiro-OMeTAD/Au21.700.890.6313.14[15]
MoS2ITO/MoS2/Csx(MAyFA1-y)1-xPb(IzBr1-z)3/spiro-OMeTAD/Au16.240.560.373.36[16]
MoS2/TiO2ITO/TiO2/MoS2/MAPbI3/spiro-OMeTAD/Au13.360.650.514.43[17]
MoS2/SnO2ITO/SnO2/MoS2/FAxMA1-xBrxClyI1-x-y/spiro-OMeTAD/Ag24.571.110.7921.73[18]
MoS2Graphene/MoS2/MAPbI3/PTAA/Au20.920.910.7614.42[19]
MoS2ITO/MoS2/MAPbI3/PCBM/Al12.500.850.576.01[20]
SnS2ITO/SnS2/MAPbI3/Spiro-OMeTAD/Au23.700.950.6113.63[21]
SnS2ITO/SnS2/MAPbI3/Spiro-OMeTAD/Au21.701.0110.6013.20[22]

Table 1.

Photovoltaic parameters of perovskite solar cell devices using dichalcogenide transition metals (TMDs) as ETLs.

2.1 TiS2 electron transport layer

TiS2 is one of the TMDC family that has been intensively studied recently due to its semi-metallic properties with low-bandgap value, i.e., 0.2 eV. With high electrical conductivity, i.e., 1 x 104 S m−1, this material is potential as an electrode in many applications including lithium-ion batteries and solar cells. Despite its excellent electrical properties, the use of TiS2 as independent electrode material in the application is limitedly demonstrated. It is mainly stacked with other materials such as MoS2 [24] or TiO2 to improve the properties in applications. For the case of MoS2 stacked with TiS2, the TiS2 can form Schottky contact with MoS2 with barrier height [24] between these two atomic layers can be varied by the doping type and concentration either in the MoS2 or TiS2 side (Figure 3). This certainly provides a wider opportunity to modify the electrical properties of the system for desired performance in application. In the typical process, n-type-doped TiS2–MoS2 (ML) contacts exhibit a barrier height relatively larger, i.e., 1.0 eV below doping level degeneracy. Nevertheless, these n-type-doped contacts still have the potential as the switch in high-power as well as tunnel Schottky barrier MOSFETs. In contrary to the n-type doped system, the p-type-doped TiS2–MoS2 (ML) exhibits a zero barrier height at a particular doping concentration, i.e., 5 × 1018 cm−3. Under this condition, the depletion region width is zero and the band becomes flat, revealing that the contact is ohmic and the barrier height is small. These results reveal the unique unusual interfacial properties arising from this ultimate thin contact that promise a special function in the application. This phenomenon could be the driving factor for an efficient photocarrier extraction in the perovskite solar cells using ETL modified with MoS2 or TiS2 atomic layer.

Figure 3.

PLDOS of TiS2–MoS2 (ML) FET-like junctions doped with different doping concentrations and the variation of band structure at interface B. a–d The doping concentrations are: N = 5 × 1019 cm−3, N = 1 × 1019 cm−3, N = 5 × 1018 cm−3, and P = 5 × 1018 cm−3. The thickness of TiS2 is four layers. On the right side, the plot shows the variation of band structure under different doping concentrations. The scale bar is from 0.0 to 90.0 (1/eV). Interface A is the interface between TiS2–MoS2. (Reprinted from [24]. © 2020, The Author(s)).

For example, in the perovskite solar cells system with SnO2 ETL (Figure 4), there is an increase in the energy band alignment between the ETL and perovskite layer when the 2D TiS2 is attached to the surface of SnO2 [18]. The conduction band level of ETL (SnO2) reduced from 4.68 to 4.63 eV in the presence of 2D TiS2. This has narrowed the offset energy between the ETL and perovskite (conduction band level at 4.36 eV). As the result, the photogenerated carrier extraction becomes enhanced, improving the photocurrent and the power conversion efficiency. As shown in Figure 4C4F, the power conversion efficiency increases from 19.65% to 21.73% when the SnO2 ETL is modified with the 2D TiS2 atomic layer. The nature of interfacial photocarrier dynamic improvement in the presence of the 2D TiS2 atomic layer can be seen from the increase of the Voc, FF, and the IPCE of the device. This process is also reflected by the decrease in the device hysteresis and the improvement of the stability properties.

Figure 4.

(A) Cross-sectional SEM image of the PSC. (B) The energy level diagram. (C) Representative J-V curves of the PSCs with SnO2 or SnO2 /2D TiS2 as ETLs. (D) EQE curve and integrated current density of the PSC with SnO2 /2D TiS2 as the ETL. (E) Histogram of the PCE of PSCs with SnO2 and SnO2 /2D TiS2 as ETLs analyzed from 25 cells. (F) Steady-state efficiency of the PSCs with SnO2 and SnO2/2D TiS2 as ETLs measured under constant voltages of 0.86 V and 0.92 V, respectively. (Reprinted from [18]. © 2019 Royal Society of Chemistry).

Figure 5 explains in detail how the photocarrier dynamic in the device was impressively modified in the presence of a 2D TiS2 atomic layer on the surface of SnO2 ETL. As presented, the photocurrent is enhanced impressively. This is the result of enhanced interfacial charge transfer as indicated by the transient and steady-state photoluminescence analysis result, which is also supported by the electrochemical impedance spectroscopy result, showing decrease in the interfacial charge transfer resistance in the device.

Figure 5.

Comparison of SnO2 and SnO2/2D TiS2 as ETLs in PSCs: (A) Jph-V eff curves; (B) J-V curves in the dark (the dash-dot lines represent the fitting lines); (C) Nyquist plots; (D) steady-state PL spectra; (E) transient PL spectra; (F) C-V characteristics. (Reprinted from [18]. © 2019 Royal Society of Chemistry).

We also in our recent result have coupled the TiS2 atomic layer on top of the TiO2 surface to compensate for surface defect due to the oxygen vacancy, enhancing the interfacial charge transfer and transport dynamic when applied as ETL in perovskite solar cells [25]. The perovskite solar cells’ performance improves from 18.02 to 18.73% (Figure 6). Electrochemical impedance analysis revealed that there is an improvement as high as 13% in interfacial charge transfer in the ETL with 2D TiS2 and 43% improvement in the charge recombination resistance (Figure 7A). The latter is verified by the increase in the photocurrent (Figure 7B) and the decrease in the leakage current of the device when 2D TiS2 passivates the TiO2 surface (Figure 7C). We can relate this process to the reduction in the trap density in the device as shown by the value of VTFL of the double-log J-V curve as depicted in Figure 7D where the VTFL value depends linearly with the trap density in the device.

Figure 6.

Photovoltaic performance of the 2D TiS2-TiO2 NG and TiO2 NG-based PSC. (A) Schematic structure of 2D TiS2-TiO2 NG-based PSC. (B) J-V curves of the champion device. (C-F) The comparison of the photovoltaic parameters, i.e., PCE, Voc, Jsc, and FF, for the two devices. (Reprinted from [25]. © 2021 The American Chemical Society).

Figure 7.

Photoelectrical properties of the PSC device. (A) Electrochemical impedance spectra and equivalent circuit of the device. (B) Photogenerated current of the PSC device (Jph-Veff curve). (C) Semilog J-V curve of the PSC in the dark. The green lines represent the fitting line. (D) Double log J-V curve in the dark for photoelectrical dynamic in the device. Three distinct regimes of (i) the ohmic response, (ii) filled trap transition, and (iii) SCLC are shown by different colored regions. (Reprinted from [25]. © 2021 The American Chemical Society).

2.2 MoS2 electron transport layer

MoS2 atomic layer is the most studied TMD system because of its excellent optical and electrical properties [26, 27, 28] and has been used widely in perovskite solar cells as a hole-transport layer (HTL) and an electron-transport layer (ETL) [11, 15, 26] in the form of colloidal or flakes thin film [15, 28, 29, 30]. Table 1 lists down several perovskite solar cells using MoS2 as ETL with a particular device configuration. For example, Singh, Giri, et al. [13] have obtained power conversion efficiency as high as 13.2% from PSC devices using MoS2 material as ETL. In this study, they synthesized the MoS2 film directly on FTO substrate using microwave irradiation-assisted reduction method. It is found that the efficiency obtained by MoS2 material is close to the efficiency value obtained from TiO2 and SnO2 material making MoS2 material comparable to other ETL materials. Abd Malek et al. [16] have also developed different structures of MoS2 ETL on the ITO substrate. Instead of colloidal or flake structured film, an ultrathin layer of MoS2 prepared from ultrasonic spray pyrolysis was fabricated to obtain its functionalities as ultrathin ETL in the PSC device. The result showed that the PCE device performance depended on the condition during the preparation of the MoS2 atomic layer, particularly the substrate temperature. It is demonstrated that substrate temperature of 200°C is suitable for growing high-quality MoS2 atomic layer on ITO surface, thus, optimizing the power conversion efficiency of the PSC (Figure 8). This MoS2 thin-film-based device as ETL has shown high-stability properties where its efficiency can be maintained as much as 90.24% of the original efficiency after 80 s exposure continuously under simulated solar light illumination (AM1.5).

Figure 8.

The photovoltaic parameter for MoS2 as ETL in PSC. (A) Schematic structure of the PSC device. (B) The J-V curves for the champion device, (C-F) PCE, Voc, Jsc, and FF of the MoS2 based PSC devices with different substrate temperatures, namely MoS180 (a), MoS200 (b), MoS220 (c), and MoS250 (d). (Reprinted from [16]. © 2020 Elsevier).

In addition to being used singly in the ETL, TMD materials can also be combined with other organic or inorganic electron transport materials to form electron transport materials. For example, Ahmed et al. [31] have added a MoS2 layer on top of the TiO2 layer to be used as ETL in perovskite solar cells. The use of MoS2/TiO2 as ETL has successfully increased the efficiency of the device by 16% higher than the device that only uses TiO2 as ETL. Similarly, Huang et al. [18] have successfully produced an n-i-p type plane device using SnO2 and 2D TiS2 as ETL. High efficiency was recorded by this group, which was as high as 21.73% with a relatively small hysteresis value. The increase in efficiency in this device is due to the matching of the ETL energy level and the appropriate perovskite layer as well as the lack of electron trap density in the ETL.

2.3 WS2 electron transport layer

Tungsten disulfide (WS2) share common basic properties of TMD with other systems, such as high-mobility properties, unique optoelectronic properties, large exciton-binding energy, and good physical and chemical stability as well as ambipolar properties [11]. In addition, WS2 has an energy level that is suitable for the perovskite layer of three types of cations (Figure 9) and can be easily synthesized by the ultrasonic spray pyrolysis method. WS2 also has high stability as well as having fast interface charge transfer properties [32]. Among the available 2D TMD, the energy band structure of WS2 is a much better match with the common perovskite of MAPbI3 (Figure 10). Furthermore, it also has a relatively larger bandgap if compared with the other system in this class of materials, promising facile excitonic separation during the photovoltaic process and producing better power conversion efficiency.

Figure 9.

Energy levels of dichalcogenide transition metal materials (TMDs) as ETLs and MAPbI3 as perovskite layers in perovskite solar cells.

Figure 10.

Energy level diagram for n-i-p perovskite solar cells using WS2 ETL.

Recently, we have realized the PSC device utilizing the WS2 layer as ETL and evaluated how the number of layers of WS2 influences the carrier dynamic in the device [5]. We prepared the WS2 atomic layer via ultrasonic spray pyrolysis. Figure 11 shows a schematic diagram of the 2D atomic layer preparation. A modified commercially available ultrasonic spray system (Daiso, Japan) was used. A homemade solution container was placed on the top of the ultrasonic membrane of the system (Figure 11). Ultrasmall solution precursor mist can be produced from the process and fall on the ITO substrate surface that is positioned approximately 5 cm below the membrane. The temperature of the substrate was set at 350°C.

Figure 11.

Schematic diagram of ultrasonic spray pyrolysis for the preparation of TMD ETL.

The typical morphology of the WS2 atomic layer on the ITO substrate is shown in Figure 12A. The WS2 nanosheet’s morphology resembles a circular structure that is produced from the precursors’ mist that emerged from the ultrasonic spray membrane. Confocal Raman imaging further indicated the existence of a very thin layer of structure from the circular structure as shown in Figure 12B. Raman analysis then confirmed the phase crystallinity of the WS2 (Figure 12C). As the figure reveals, there are two sharp peaks obtained from the Raman spectrum that is centered at 348.9 cm−1and 412.3 cm−1, which are associated with the in-plane (E2g) and the out-of-plane (A1g) vibration modes of the lattice (see inset in Figure 12C) [33, 34, 35, 36, 37, 38, 39]. According to the value of the separation between these two peaks, the thickness of the atomic layer is estimated to be in the range of 10 L. The X-ray diffraction analysis further confirmed the phase crystallinity of the WS2 layer (Figure 12D) [40, 41, 42]. The high-resolution transmission electron microscopy (HRTEM) and selected area electron diffraction (SAED) analysis results (Figure 12F and G) show that the sample is single crystalline. However, the presence of SAED composed of a triple spot is related to the stacking of the WS2 atomic layer during the transfer to the lacey grid for HRTEM analysis. The XPS analysis then further confirmed the Raman and XRD analysis results on the phase crystallinity of the sample of which it belongs to WS2 (Figure 12HI).

Figure 12.

The morphology, phase crystallinity, chemical state properties of WS2 nanosheet. (A) FESEM image of WS2 nanosheet on the ITO substrate. (B-C) Raman imaging and spectrum of WS2 were obtained using 532 nm laser excitation. The inset in (C) shows the corresponding main vibration mode of Raman. (D) XRD spectrum for WS2 nanosheet showing 2H phase. (E-F) Low and high-resolution TEM image of WS2 nanosheet. (G) SAED pattern of WS2 nanosheet showing at least three stacking WS2 nanosheets. (H-I) High-resolution scan of XPS at W and S binding energy of WS2 nanosheet. (Reprinted from [5]. © 2020 Wiley-VCH GmbH).

PSCs device was fabricated using the WS2 atomic layer as ETL and investigated how the thickness of the WS2 ETL influenced the photovoltaic process. The structure of the PSC device is ITO/WS2 nanosheets/Perovskite/Spiro-OMeTAD/Au. Perovskite used was triple cations system of Cs0.05[MA0.13FA0.87]0.95Pb (I0.87Br0.13)3 [43].

It was found that the thickness, represented by the number of layers, of the WS2 atomic layer ETL, strongly influences the power conversion efficiency of the PSC device (Figure 13). The results show that the PCE performance improves with the increase of thickness from 4 L to the optimum thickness of 7 L (WS30 sample in the figure). The optimized WS2 ETL thickness can produce a PSC device with PCE as high as 18.21% with Jsc, Voc, and FF as high as 22.24 mA cm−2, 1.12 V, and 0.731, respectively. The average performance was 17.84%, 22.33 mA cm−2, 1.10 V, 0.731 for PCE, Jsc, Voc, and FF, respectively. However, due to an increase in the energetic disorder when using the WS2 ETL, the device performance then declined when the thickness of the ETL increased above 7 L. From Figure 13, we can also see that the values of Voc and fill factor (FF) are impressively high, which is higher than 1.1 V for Voc and approximately 74% for FF. This reflects that the photogenerated carrier dynamic in the device is high and the photogenerated carrier is effectively extracted to the external circuit to produce photocurrent [44]. This is verified by the high-external quantum efficiency (EQE) of the device as shown in Figure 13F.

Figure 13.

The photovoltaic performance of PSC using different thicknesses of WS2 ETL. (A) J-V curves for the champion PSC device. (B–E) The statistic plot for PCE, Voc, Jsc, and FF, respectively. (F) EQE and integrated current density (J) of the corresponding device. (Reprinted from [5]. © 2020 Wiley-VCH GmbH).

To understand the extent effect of the WS2 atomic layer as ETL in the PSC device, the device performance was compared with the reference PSC utilizing well-known SnO2 ETL. In the typical process, the performance of SnO2-based PSC shows lower performance than the WS2 atomic layer–based device (Figure 14). Steady-state and transient photoluminescence analysis revealed that the interfacial charge transfer from the perovskite to ETL is high in the WS2 atomic layer [45], the result of optimized coupling due to ultra-flat surface morphology offered by the WS2 atomic layer. This phenomenon is further confirmed by the electrochemical impedance spectroscopy analysis result where it is obtained that the interface charge transfer resistance is lower in the WS2-based PSC device than the SnO2-based device. Thus, it can be remarked that the WS2 atomic layer enables highly active interfacial charge transfer for a high-performance PSC device.

Figure 14.

The comparison of the photovoltaic parameter between WS2 (7 L, WS30 sample) and SnO2-based PSC device. (A) J-V curves for the champion device. (B–E) The comparison of PCE, Voc, Jsc, and FF for the two devices, respectively. (F–H) Steady-state PL, time-resolved PL spectra (TRPL), and electrochemical impedance spectra for WS2 and SnO2-based PSC devices, respectively. (Reprinted from [5]. © 2020 Wiley-VCH GmbH).

Advertisement

3. Conclusions

2D atom thick TMD promises facile charge extraction and transport in the perovskite solar cells due to its ultimate thin and single-crystalline nature. The optimization of the 2D TMD layer to obtain a large dimension on the substrate surface is necessary to further promote a highly dynamic photogenerated carrier in the perovskite solar cells device. These materials may become a potential platform for high-performance perovskite solar cells.

Advertisement

Acknowledgments

We acknowledged the financial support from the Universiti Kebangsaan Malaysia for supporting this project under GUP-2019-071 and DIP-2021-025.”

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Jeong J, Kim M, Seo J, Lu H, Ahlawat P, Mishra A, et al. Pseudo-halide anion engineering for α-FAPbI 3 perovskite solar cells. Nature. 2021;592(7854):381-385
  2. 2. Lin PY, Loganathan A, Raifuku I, Li MH, Chiu YY, Chang ST, et al. Pseudo-Halide Perovskite Solar Cells. Advanced Energy Materials. 2021;2100818
  3. 3. Deng X, Cao Z, Yuan Y, Chee MOL, Xie L, Wang A, et al. Coordination modulated crystallization and defect passivation in high quality perovskite film for efficient solar cells. Coordination Chemistry Reviews. 2020;420:213408
  4. 4. Jeong I, Park YH, Bae S, Park M, Jeong H, Lee P, et al. Solution-Processed Ultrathin TiO2 Compact Layer Hybridized with Mesoporous TiO2 for High-Performance Perovskite Solar Cells. ACS Applied Materials & Interfaces. 2017;9(42):36865-36874
  5. 5. Malek NAA, Alias N, Umar AA, Zhang X, Li X, Saad SKM, et al. Enhanced Charge Transfer in Atom-Thick 2H–WS2 Nanosheets’ Electron Transport Layers of Perovskite Solar Cells. Solar RRL. 2020;4(10):2000260
  6. 6. Umar AA, Zhang X, Saad SKM, Abd Malek NA, Liu K, Alias N, et al. Enhancing the interfacial carrier dynamic in perovskite solar cells with an ultra-thin single-crystalline nanograss-like TiO 2 electron transport layer. Journal of Materials Chemistry A. 2020;8(27):13820-13831
  7. 7. Huang P, Yuan L, Zhang K, Chen Q , Zhou Y, Song B, et al. Room-temperature and aqueous solution-processed two-dimensional TiS2 as an electron transport layer for highly efficient and stable planar n–i–p perovskite solar cells. ACS Applied Materials & Interfaces. 2018;10(17):14796-14802
  8. 8. Sumpter BG, Liang L, Nicolai A, Meunier V. Interfacial properties and design of functional energy materials. Accounts of Chemical Research. 2014;47(11):3395-3405
  9. 9. Kuc A, Heine T. On the Stability and Electronic Structure of Transition-Metal Dichalcogenide Monolayer Alloys Mo1− xXxS2− ySey with X= W, Nb. Electronics. 2016;5(1):1
  10. 10. Furchi MM, Pospischil A, Libisch F, Burgdörfer J, Mueller T. Photovoltaic effect in an electrically tunable van der Waals heterojunction. Nano Letters. 2014;14(8):4785-4791
  11. 11. Van Le Q , Choi J-Y, Kim SY. Recent advances in the application of two-dimensional materials as charge transport layers in organic and perovskite solar cells. FlatChem. 2017;2:54-66
  12. 12. Ren Y, Yang X, Zhou L, Mao JY, Han ST, Zhou Y. Recent advances in ambipolar transistors for functional applications. Advanced Functional Materials. 2019;29(40):1902105
  13. 13. Singh R, Giri A, Pal M, 7Thiyagarajan K, Kwak J, Lee J-J, et al. Perovskite solar cells with an MoS 2 electron transport layer. Journal of Materials Chemistry A. 2019;7(12):7151-7158
  14. 14. Yin G, Zhao H, Feng J, Sun J, Yan J, Liu Z, et al. Low-temperature and facile solution-processed two-dimensional TiS2 as an effective electron transport layer for UV-stable planar perovskite solar cells. Journal of Materials Chemistry A. 2018;6(19):9132-9138
  15. 15. Singh R, Giri A, Pal M, Thiyagarajan K, Kwak J, Lee J-J, et al. Perovskite solar cells with an MoS2 electron transport layer. Journal of Materials Chemistry A. 2019;7(12):7151-7158
  16. 16. Abd Malek NA, Alias N, Saad SKM, Abdullah NA, Zhang X, Li X, et al. Ultra-thin MoS2 nanosheet for electron transport layer of perovskite solar cells. Optical Materials. 2020;104:109933
  17. 17. Ahmed MI, Hussain Z, Khalid A, Amin HMN, Habib A. Absorption enhancement in CH3NH3PbI3solar cell using a TiO2/MoS2nanocomposite electron selective contact. Materials Research Express. 2016;3(4):045022
  18. 18. Huang P, Chen Q , Zhang K, Yuan L, Zhou Y, Song B, et al. 21.7% efficiency achieved in planar n–i–p perovskite solar cells via interface engineering with water-soluble 2D TiS 2. Journal of Materials Chemistry A. 2019;7(11):6213-6219
  19. 19. Jung DH, Oh YJ, Nam YS, Lee H. Effect of layer number on the properties of stable and flexible perovskite solar cells using two dimensional material. Journal of Alloys and Compounds. 2021;850:156752
  20. 20. Dasgupta U, Chatterjee S, Pal AJ. Thin-film formation of 2D MoS2 and its application as a hole-transport layer in planar perovskite solar cells. Solar Energy Materials & Solar Cells. 2017;172:353-360
  21. 21. Zhao E, Gao L, Yang S, Wang L, Cao J, Ma T. In situ fabrication of 2D SnS2 nanosheets as a new electron transport layer for perovskite solar cells. Nano Research. 2018;11(11):5913-5923
  22. 22. Chu W, Li X, Li S, Hou J, Jiang Q , Yang J. High-Performance Flexible Perovskite Solar Cells with a Metal Sulfide Electron Transport Layer of SnS2 by Room-Temperature Vacuum Deposition. ACS Applied Energy Materials. 2019;2(1):382-388
  23. 23. Sobayel K, Akhtaruzzaman M, Rahman K, Ferdaous M, Al-Mutairi ZA, Alharbi HF, et al. A comprehensive defect study of tungsten disulfide (WS2) as electron transport layer in perovskite solar cells by numerical simulation. Results in Physics. 2019;12:1097-1103
  24. 24. Gao J, Gupta M. Titanium disulfide as Schottky/ohmic contact for monolayer molybdenum disulfide. npj 2D Materials and Applications. 2020;4(1):26
  25. 25. Alias N, Ali Umar A, Malek NAA, Liu K, Li X, Abdullah NA, et al. Photoelectrical Dynamics Uplift in Perovskite Solar Cells by Atoms Thick 2D TiS2 Layer Passivation of TiO2 Nanograss Electron Transport Layer. ACS Applied Materials & Interfaces. 2021;13(2):3051-3061
  26. 26. Capasso A, Matteocci F, Najafi L, Prato M, Buha J, Cinà L, et al. Few-Layer MoS2 Flakes as Active Buffer Layer for Stable Perovskite Solar Cells. Advanced Energy Materials. 2016;6(16):1600920
  27. 27. Ambrosi A, Sofer Z, Pumera M. 2H --> 1T phase transition and hydrogen evolution activity of MoS2, MoSe2, WS2 and WSe2 strongly depends on the MX2 composition. Chemical Communications (Cambridge). 2015;51(40):8450-8453
  28. 28. Xiao S, Xiao P, Zhang X, Yan D, Gu X, Qin F, et al. Atomic-layer soft plasma etching of MoS2. Scientific Reports. 2016;6:19945
  29. 29. Wang L, Hu P, Long Y, Liu Z, He X. Recent advances in ternary two-dimensional materials: synthesis, properties and applications. Journal of Materials Chemistry A. 2017;5(44):22855-22876
  30. 30. Jeon J, Jang SK, Jeon SM, Yoo G, Jang YH, Park JH, et al. Layer-controlled CVD growth of large-area two-dimensional MoS2 films. Nanoscale. 2015;7(5):1688-1695
  31. 31. Ahmed MI, Hussain Z, Khalid A, Amin HMN, Habib A. Absorption enhancement in CH3NH3PbI3 solar cell using a TiO2/MoS2 nanocomposite electron selective contact. Materials Research Express. 2016;3(4):045022
  32. 32. Park J, Woo H, Jeon S. Impact of fast transient charging and ambient on mobility of WS2 field-effect transistor. Journal of Vacuum Science & Technology, B: Nanotechnology & Microelectronics: Materials, Processing, Measurement, & Phenomena. 2017;35(5):050601
  33. 33. Gutiérrez HR, Perea-López N, Elías AL, Berkdemir A, Wang B, Lv R, et al. Extraordinary Room-Temperature Photoluminescence in Triangular WS2 Monolayers. Nano Letters. 2013;13(8):3447-3454
  34. 34. Qiao S, Yang H, Bai Z, Peng G, Zhang X. Identifying the number of WS2 layers via Raman and photoluminescence spectrum. Advances in Engineering Research. 2017;141:1408-1413
  35. 35. Park J, Kim MS, Cha E, Kim J, Choi W. Synthesis of uniform single layer WS2 for tunable photoluminescence. Scientific Reports. 2017;7(1):16121
  36. 36. Zhang Y, Shi J, Han G-F, Li M, Ji Q , Ma D, et al. Chemical vapor deposition of monolayer WS2 nanosheets on Au foils towards direct application in hydrogen evolution. Nano Research. 2015;8:2881-2890
  37. 37. Molas MR, Nogajewski K, Potemski M, Babinski A. Raman scattering excitation spectroscopy of monolayer WS2. Scientific Reports. 2017;7(1):5036
  38. 38. Zhao W, Ghorannevis Z, Amara KK, Pang JR, Toh M, Zhang X, et al. Lattice dynamics in mono- and few-layer sheets of WS2 and WSe2. Nanoscale. 2013;5:9677-9683
  39. 39. Berkdemir A, Gutiérrez HR, Botello-Méndez AR, Perea-López N, Elías AL, Chia C-I, et al. Identification of individual and few layers of WS2 using Raman Spectroscopy. Scientific Reports. 2013;3:1755
  40. 40. Jian-Ping Z, Jun M, Jin-Ming L, Jian Y, Junkai H, Yongtao M, et al. Fabrication of novel heterostructured few layered WS 2-Bi 2 WO 6/Bi 3.84 W 0.16 O 6.24 composites with enhanced photocatalytic performance. Applied Catalysis B: Environmental. 2015;179:220-228
  41. 41. Sang Y, Zhao Z, Zhao M, Hao P, Leng Y, Liu H. From UV to near-infrared, WS2 nanosheet: a novel photocatalyst for full solar light spectrum photodegradation. Advanced Materials. 2015;27(2):363-369
  42. 42. Hai X, Chang K, Pang H, Li M, Li P, Liu H, et al. Engineering the Edges of MoS2 (WS2) Crystals for Direct Exfoliation into Monolayers in Polar Micromolecular Solvents. Journal of the American Chemical Society. 2016;138(45):14962-14969
  43. 43. Saliba M, Matsui T, Seo JY, Domanski K, Correa-Baena JP, Nazeeruddin MK, et al. Cesium-containing triple cation perovskite solar cells: improved stability, reproducibility and high efficiency. Energy & Environmental Science. 2016;9(6):1989-1997
  44. 44. Kim G-W, Shinde DV, Park T. Thickness of the hole transport layer in perovskite solar cells: performance versus reproducibility. RSC Advances. 2015;5(120):99356-99360
  45. 45. Wang F, Chen Y, Han G, Zhang Q , Ma Q. The interface and its role in carrier transfer/recombination dynamics for the planar perovskite solar cells prepared under fully open air conditions. Current Applied Physics. 2016;16(10):1353-1363

Written By

Akrajas Ali Umar, Nurul Ain Abd Malek, Nabilah Alias and Abang Anuar Ehsan

Submitted: 23 December 2021 Reviewed: 22 February 2022 Published: 29 April 2022