Next Article in Journal
Fatigue Cracking of Additively Manufactured Materials—Process and Material Perspectives
Previous Article in Journal
A Method for Reducing of the Overall Torsion for Reinforced Concrete Multi-Storey Irregular Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Contrasting Electron and Hole Transfer Dynamics from CH(NH2)2PbI3 Perovskite Quantum Dots to Charge Transport Layers

1
Shanghai Ultra-Precision Optical Manufacturing Engineering Research Center, and Key Laboratory of Micro and Nano Photonic Structures (Ministry of Education), Department of Optical Science and Engineering, Fudan University, Shanghai 200433, China
2
Institute of Functional Nano & Soft Materials (FUNSOM), Jiangsu Key Laboratory for Carbon-Based Functional Materials & Devices, Joint International Research Laboratory of Carbon-Based Functional Materials and Devices, Soochow University, Suzhou 215123, China
3
School of Electronic and Optical Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2020, 10(16), 5553; https://doi.org/10.3390/app10165553
Submission received: 21 July 2020 / Revised: 7 August 2020 / Accepted: 8 August 2020 / Published: 11 August 2020
(This article belongs to the Section Applied Physics General)

Abstract

:
In this work, the ultrafast transient absorption spectroscopy (TAs) was utilized to first investigate the charge transfer from the emerging FAPbI3 (FA = CH(NH2)2) perovskite quantum dots (PQDs) to charge transport layers. Specifically, we compared the TAs in pure FAPbI3 PQDs, PQDs grown with both electron and hole transfer layers (ETL and HTL), and PQDs with only ETL or HTL. The TA signals induced by photoexcited electrons decay much faster in PQDs samples with the ETL (~20 ps) compared to the pure FAPbI3 PQDs (>1 ns). These results reveal that electrons can effectively transport between coupled PQDs and transfer to the ETL (TiO2) at a time scale of 20 ps, much faster than the bimolecular charge recombination inside the PQDs (>1 ns), and the electron transfer efficiency is estimated to be close to 100%. In contrast, the temporal evolution of the TA signals in the PQDs with and without HTL exhibit negligible change, and no substantive hole transfer to the HTL (poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine], PTAA) occurs within 1 ns. The much slower hole transfer implies the further potential of increasing the overall photo-carrier conversion efficiency through enhancing the hole diffusion length and fine-tuning the coupling between the HTL and PQDs.

1. Introduction

Perovskite solar cells (PSCs) have made significant achievements in the certified power conversion efficiency (PCE), which increases from 3.8% to 25.2% [1] in recent several years. In particular, organolead halide (e.g., MAPbI3, MA = CH3NH3; FAPbI3, FA = CH(NH2)2) perovskite photovoltaic (PV) devices attract extensive investigation owing to the remarkable benefits of their large absorption coefficient [2], outstanding charge carrier mobility and long carrier diffusion lengths [3,4,5,6]. With the aim of designing higher-performance PSCs, it is critical to better understand the charge transfer (CT) from perovskite to charge transport layer. However, CT time scales in MAPbI3 PSCs reported to date vary from sub-picosecond [7,8,9] to hundreds of picoseconds [3,10,11,12] or nanoseconds [8,13,14], even for the same acceptor. For instance, Xing et al. reported the charge-carrier transfer time of 0.40 ns and efficiency of 92% for MAPbI3 with selective electron layers ([6,6]-phenyl-C61-butyric acid methyl ester, PCBM) [3]. Wang et al. revealed the electron transfer time from MAPbI3 to planar TiO2 to be 39.9 ± 2.5 ps [10]. Makuta et al. estimated the electron injection rate from MAPbI3 perovskite to TiO2 to be 11 ± 1 ns [14]. One of the main reasons for huge differences in measured CT time in MAPbI3 PSCs may arise from the material variations of the MAPbI3 perovskite films. To fabricate MAPbI3 films with the proper phase structure and crystalline orientation, essential for the light-electrical conversion, it is required sophisticated high temperature annealing and several other treatments [15,16]. Therefore, the uniformity of the film crystallinity and the material composition can be hardly ensured.
Compared with MAPbI3, FAPbI3 is superior in light absorption characteristics [17,18,19,20,21] and thermal stability [21,22]. In particular, FAPbI3 can be fabricated as quantum dots (QDs) which possess proper phase structure and crystalline orientation without the need of high temperature annealing [23,24,25]. Relative to bulk or thin-film perovskite materials, perovskite QDs (PQDs) possess high crystallinity during synthesis and exhibit unique features such as tunable bandgaps [26,27] and high quantum yield [28,29,30,31], which contribute to PV applications. The corresponding PSCs made of FAPbI3 PQDs have been demonstrated to exhibit reasonable high PCEs [25,32]. However, the investigation of CT processes to reveal the intrinsic transfer time and efficiency is still lacking, which hinders the clarification of the important factors limiting the PCEs. Unlike the exciton or charge diffusion in the MAPbI3 thin films as the active layer of PSCs, the diffusion process in the active layer of FAPbI3 PQDs critically relies on the electronic couplings between the PQDs before the free carriers are injected into the charge transport layer. Moreover, the charge injection processes strongly depend on the interfacial interaction between the PQDs and charge transport layers. It is unclear whether the electronic couplings between PQDs and at the interface of PQDs and charge transport layers are strong enough to ensure quick transport of the carriers to reach the interface within short time and yield effective charge injection before they are recombined within the PQDs.
Herein, in order to clarify the CT process in planar junctions, we studied for the first time the dynamics of interfacial CT from FAPbI3 PQDs to electron and hole transport layers (ETL and HTL) (tris(pentafluorophenyl)borane-doped poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine] (PTAA) HTL and TiO2 ETL) by utilizing the ultrafast transient absorption spectroscopy (TAs). The temporal evolution of the TA signals in pure FAPbI3 PQDs, PTAA/FAPbI3/TiO2, PTAA/FAPbI3, and FAPbI3/TiO2 were compared. Notably, we observed a fast electron transfer process at the time scale of 20 ps, and the transfer efficiency is close to the unity. In contrast, we did not observe significant hole transfer to PTAA within our experimental time window of ~1 ns.

2. Materials and Methods

The FAPbI3 PQDs were synthesized according to the recent report [25]. For fabrication of the PTAA/FAPbI3/TiO2 sample, a 50 nm compact hydrothermal TiO2 layer was deposited above the quartz substrate. The TiO2 film was exposed to UV-ozone for 10 min before FAPbI3 PQDs deposition. The FAPbI3 PQDs (45 mg/mL in chloroform) were then spin-cast on top of as-prepared TiO2 film at 2000 rpm for 40 s. Then the PQDs film was swiftly treated by 200 µL MeOAc for 5 s to remove the surface ligands and spun at 2000 rpm for 20 s. The PQDs film deposition process, performed in the dry air-filled glovebox with relative humidity below 10%, could build up a ~160 nm thick film. Before the fabrication of HTL, the PQDs film was annealed at 70 °C for 7 min in a nitrogen-filled glovebox, which can further remove the solvent. By spin-coating the tris(pentafluorophenyl)borane-doped PTAA toluene solution (15 mg/mL, dopant/PTAA = 5% in weight) at 3000 rpm for 40 s, the 40 nm HTL could be deposited above the FAPbI3 PQDs layer in the N2 glovebox.
We used a Perkin Elmer model Lambda 950 spectrophotometer to characterize the ultraviolet-visible (UV-vis) absorption spectra of the FAPbI3 PQDs. Their photoluminescence (PL) spectra were measured with a FluoroMax-4 spectrofluorometer (HORIBA Scientific). The transmission electron microscopy (TEM) images were obtained by using Tecnai G2 F20 S-Twin transmission electron microscope.
The TAs were measured by using a pump−probe experimental system. The pump and probe pulses were delivered by a Ti:sapphire amplifier laser with the pulse width of 100 fs and repetition rate of 1 kHz. We used a mechanical chopper to modulate the pump beam at ~300 Hz and focused the laser on the sample with a spot diameter of ~3 mm. The pump beam, converted from the fundamental laser beam by using of a β-barium borate (BBO) crystal, has the central wavelength of 400 nm. Supercontinuum white light was generated by a sapphire plate irradiated by 800 nm pulses and it then passed though different color filters with 10 nm bandwidth to serve as the probe pulses with the wavelength range from 500 to 1000 nm. A silicon photodiode was used to collect the probe light transmitted through the sample. A lock-in amplifier was connected to the photodiode to extract the transmission change (ΔT) between the states with and without the pump excitation.

3. Results and Discussion

Figure 1a shows the structure of PTAA/FAPbI3/TiO2 sample, where the thickness of the ETL TiO2, the light harvester layer FAPbI3 PQDs, and the HTL PTAA doped by tris(pentafluorophenyl)borane is ~50 nm, ~160 nm, and ~40 nm, respectively. As shown in Figure 1b, FAPbI3 PQDs exhibit deep valence band (VB) energy level of −5.46 eV, whereas the PTAA exhibits shallower highest occupied molecular orbital (HOMO) energy level of −5.20 eV [24,25], and the conduction band (CB) energy level values of FAPbI3 PQDs and TiO2 are −3.94 eV and −4.15 eV [25,33,34], respectively, which are in favor of the charge transfer. From the TEM image of the pure FAPbI3 PQDs (Figure 1c), we found the PQDs have an average size of about 15 nm. The UV−vis absorption spectra show a broad absorption with the first absorption peak at ~770 nm as a result of optical transition from the ground state to the first excitonic state (Figure 1d). A narrow PL peak centered at ~770nm can also be identified in the PL spectra. Based on this structure, the PCE of our FAPbI3 PQDs solar cells has reached 11.6% [25]. Using PTAA as the hole transport material (HTM) instead of the conventional 2,2,7,7-tetrakis-(N,N-di-p-methoxyphenylamine)-9,9-bifluorene (Spiro-OMeTAD) can reduce manufacturing costs and avoid device instability caused by complex doping and oxidation processes [15,35,36].
We then performed the ultrafast TA measurements in the pure FAPbI3 PQDs and PTAA/FAPbI3/TiO2. Figure 2a,b show their transmission modulation (ΔT/T) spectra under excitation of 400-nm pump laser pulses. The spectra of pure FAPbI3 display positive photoinduced bleaching (PB) signals in the range of 650–1000 nm. These PB signals are pronouncedly enhanced around 750 nm, which coincides with the wavelength of the first absorption peak, shown in Figure 1d, resulting from the transition from the ground state to the first excitonic state. We can thus ascribe the strong PB signals to the ground state bleaching (GSB). Actually, previous works have demonstrated that in the FAPbI3 PQDs the strong PB signals was mainly contributed by the holes in the valence band after laser excitation [37,38,39]. We note a slight red shift of the PB peak with increasing time delay. This shift may be attributed to the bandgap renormalization (BGR) as a result of the Coulomb interaction of photoexcited free carriers near the band edge [40,41].
In the range of 500–650 nm, the ΔT/T spectra show negative photoinduced absorption (PA) signals. The time dependences of the PA signals at different wavelengths are nearly identical, as shown in Figure 2c. This indicates that the PA signals are mainly originated from one component of the photoexcited species, since different components typically have distinct dynamics and spectra weight. In principle, both transient absorption of conduction band electron and BGR may contribute to the PA signals [42,43]. However, both contributions in the pure FAPbI3 PQDs are proportional to the number of electrons or holes which exist as pairs, therefore, their dynamics are expected to behave in the same way.
Compared to the pure FAPbI3 PQDs, the ΔT/T spectra in PTAA/FAPbI3/TiO2 also display the apparent PB and PA features. However, the PA wavelength range in the latter is broadened to above 700 nm and the PA peak is shifted to the longer wavelength of ~650 nm. Consequently the strong PB signals only appear in a much narrower wavelength range, and the PB peaks show negligible red shift over a long time delay. We may postulate at this stage that a significant number of photoexcited carriers in the PQDs are quickly transferred to the charge transport layer, which leads to a strong internal electric field, resulting in a large Stark effect to modify the bandgap of the PQDs and therefore the structure of transient absorption spectra [40,41]. Another possibility is that the strong electronic coupling between the PbI2 and the TiO2, favoring a fast electron transfer process, leads to a broadened exciton band in the TA spectra [34]. The quick electron transfer may result in an imbalance of the electron and hole numbers in the PQDs. This imbalance may affect the time dependences of the PA signals probed at various wavelengths in a different way. Actually, we can see from Figure 2d some small differences of the normalized time dependences of the PA signals in 500–700 nm. We also note significant PB and PA signals remained at long delay time of 800 ps in Figure 2b, pointing to the existence of long-lived holes residing in the PQDs as a result of the quick electron transfer. Assured evidences of quick electron transfer will be given later.
Because the average size of FAPbI3 PQDs is about 15 nm, much larger than its exciton Bohr radius (~8 nm), the photoexcited excitons in the FAPbI3 PQDs quickly dissociate to become free carriers [44,45]. The free carriers then decay via bimolecular recombination [37,46]. However, the Auger recombination may emerge for large carrier density when the PQDs are excited with high pump fluences [41,46]. The Auger recombination is normally much faster than the bimolecular recombination, and thus the decay of ΔT/T signals may display two stages with different time scales [41,46]. With increasing pump fluences, an even faster decaying process at the time scale of several picoseconds owing to the quick biexciton recombination may occur as previously discovered in some PQDs [41,46,47,48,49,50]. This was also observed in our FAPbI3 PQDs, as shown in the supplemental Figures S1 and S2 and Table S1 (Supplementary Materials). To avoid the very quick biexciton recombination and reduce the effect of the fast Auger recombination on the charge transfer processes, we have measured the temporal evolution of the ΔT/T signals for both pure FAPbI3 PQDs and PTAA/FAPbI3/TiO2 at a low pump fluence of 3.5 μJ/cm2. The measured time dependences of the PA signals at 570 nm and the PB signals at 770 nm are shown in Figure 3a,b. We then used a formula with the sum of two exponential decaying terms, T / T = A 1 e t / τ 1 + A 2 e t / τ 2 , to fit the temporal evolution of the PA and PB decay signals (see solid curves as fitting results in the Figures). Both PA and PB in pure FAPbI3 PQDs show a fast component with τ 1 120   ps and a slow component with τ 2 1600   ps (see Table 1), attributed to the Auger and bimolecular recombination, respectively. The slow component has much larger amplitude, indicating that the bimolecular recombination is the dominant decay mechanism.
We then turn to the time dependence of the PA signal at 570 nm for PTAA/FAPbI3/TiO2 shown in Figure 3a. The PA in PTAA/FAPbI3/TiO2 displays a much faster decaying process compared to that of the pure FAPbI3 PQDs. The fitting using the above formula reveals τ 1 20   ps , and this fast decay component contributes more than 53% of the total PA amplitude. In contrast, the other exponential decay component becomes much slower at the time scale of τ 2 2600   ps , compared to that in the pure FAPbI3 PQDs. The emergence of fast decay component gives important evidence that free charge carriers in the PQDs are quickly transferred to the charge transport layer. To distinguish which charge carriers, electrons or holes, are effectively transferred within 20 ps, we may turn to the time dependences of PB signals, which are mainly contributed from the holes. The time dependences of the PB signals in Figure 3b clearly show a much slower decay process in PTAA/FAPbI3/TiO2 than in the pure FAPbI3. The fitting reveals the absence of fast component but only one slow component with τ 2 2800   ps . These results strongly support that electrons are quickly transferred to the TiO2 layer. Such transfer reduces the conduction electron absorption in the PQDs, and consequently attenuates the PA signals at 570 nm within τ 1 20   ps . The quick electron transfer leaves the holes in the valence band of the PQDs, thus markedly slowing down the PB decaying process. Actually, we also note a small rising signal in the PB of 770 nm at a time scale of 20 ps, as can be better seen in Figure 4b with expanded time scale. This is likely due to the diminishment of conduction electron absorption at this probe wavelength in PQDs as a result of the fast electron transfer to TiO2. Since the PB signals and the slow decay component of the PA signals have an approximately equal decaying time scale (~2600–2800 ps), we believe that in addition to the origin of conduction electron absorption, the PA signals at 570 nm also have a significant contribution from the BGR caused by the holes. In other words, the slow decay component of the PA at 570 nm in PTAA/FAPbI3/TiO2 is mainly caused by the holes residing in the PQDs. This implies that the electrons have transfer efficiency much larger than 53% at the time scale of 20 ps. Moreover, in PTAA/FAPbI3/TiO2, we did not observe the decay component at ~1600 ps caused by the bimolecular recombination inside the PQDs. Hence we believe nearly all electrons have been transferred to TiO2, which means the transfer efficiency is close to 100%. On the contrary, in the time window of our experiments (~1 ns), we have not observed substantive hole transfer from the FAPbI3 PQDs to PTAA.
To confirm the above discussions that electrons have quick and efficient transfer from FAPbI3 PQDs to TiO2, whereas the hole transfer from FAPbI3 PQDs to PTAA is not significant within 1 ns, we also have measured the time dependences of PA and PB signals for FAPbI3 PQDs covered with only TiO2, allowing only electron transfer, or just PTAA, allowing only hole transfer. The ΔT/T spectra of FAPbI3/TiO2 and PTAA/FAPbI3 allow us to separately investigate electron transfer and hole transfer respectively. The measured results are shown in Figure 4a,b. We can see that the PA decaying processes in FAPbI3/TiO2 and PTAA/FAPbI3/TiO2 resemble each other very closely, and the PA decaying processes in PTAA/FAPbI3 and pure FAPbI3 PQDs are quite similar. Consistently, the variations of PB decaying processes in these samples behave in the same manner as those of PA decaying processes. Detailed fitting parameters shown in Table 1 further confirm such behaviors. These comparisons therefore confirm the above discussions of effective electron transfer at the time scale of 20 ps and insignificant hole transfer within 1 ns.
The most reasonable explanations for the different electron and hole transfer processes are described below. The strong electronic interactions between the FAPbI3 and the TiO2 is beneficial to the electron transfer [34], and the coupling between PQDs also favors the electron diffusion. Owing to the suitable energy levels and ideal interface between the FAPbI3 PQDs and TiO2 layers, the electron transfer is sufficiently faster than the bimolecular charge recombination. In contrast, the coupling between the PTAA and FAPbI3 is likely not strong enough to yield quick hole injection, though their HOMO energy levels are matched. One possibility is related to the defective physical contact. Less physical contact, due to the roughness of the interfacial planar heterojunction between the FAPbI3 PQDs and PTAA layers, would increase the need of the hole diffusion before finding a suitable site for transfer [35]. Another possibility is that the coupling between PQDs are not strong enough to support the quick transport of holes to reach the interface. Based on the above CT results and the underlying limiting factors, we therefore propose to make proper interface modification between the FAPbI3 PQDs and PTAA to largely enhance the hole injection rate. In addition, the coupling strength affecting the hole transport between neighboring FAPbI3 PQDs also needs to be further examined and improved. If the hole transfer rate and efficiency are greatly enhanced, a large increase of PCE in FAPbI3 PQDs based solar cells can be expected.

4. Conclusions

In conclusion, based on the comparative studies of the photoexcited charge carrier dynamics in FAPbI3, PTAA/FAPbI3, FAPbI3/TiO2, and PTAA/FAPbI3/TiO2 planar junctions by using ultrafast TAs, we unveiled evidences of quick electron transfer process from FAPbI3 PQDs to TiO2 ETL at the time scale of 20 ps and the transfer efficiency is estimated to be close to 100%. In contrast, no significant hole transfer to the PTAA HTL were observed within ~1 ns, which is likely due to the fact that the interface coupling is not strong enough to drive quick hole transport through PQDs and fast hole injection into PTAA. Our results suggest that improving the ultrafast hole transfer at PQDs/HTL interface is essential for the high-performance of FAPbI3 PQDs solar cells.

Supplementary Materials

The following are available online at https://www.mdpi.com/2076-3417/10/16/5553/s1, Figure S1: Temporal evolution of ΔT/T for pure FAPbI3 PQDs film with the probe wavelength of 770 nm under different pump fluences. Solid curves represent the fitting results using the formula of T / T = A 0 e t / τ 0 + A 1 e t / τ 1 + A 2 e t / τ 2 . The fitting parameters are summarized in Table S1; Figure S2: Temporal evolution of ΔT/T for pure FAPbI3 PQDs film with the probe wavelength of 570 nm under different pump fluences. Solid curves represent the fitting results using the formula of T / T = A 0 e t / τ 0 + A 1 e t / τ 1 + A 2 e t / τ 2 . Relative to the case of pump fluence of ≈14 μJ/cm2, the quick biexciton recombination is absent under the low fluences; Table S1: Parameters used to fit the PB decay processes of the pure FAPbI3 PQDs film probed at 770 nm and excited at 400 nm with different pump fluences, shown in Figure S1.

Author Contributions

Conceptualization, Z.L., J.Y. (Jianyu Yuan), C.S., W.M. and H.Z. (Haibin Zhao); Formal analysis, Z.L., S.L., J.Y. (Jiabei Yuan), K.J., H.Z. (Hongchang Zhao), H.X. and G.N.; Investigation, Z.L.; Methodology, Z.L., J.Y.(Jianyu Yuan), C.S., W.M. and H.Z. (Haibin Zhao); Project administration, G.N., L.C. and H.Z. (Haibin Zhao); Resources, J.Y. (Jiabei Yuan), K.J., J.Y. (Jianyu Yuan) and W.M.; Supervision, G.N., L.C. and H.Z. (Haibin Zhao); Validation, Z.L.; Writing—original draft, Z.L. and H.Z.(Haibin Zhao); Writing—review & editing, Z.L., J.Y. (Jianyu Yuan), W.M. and H.Z.(Haibin Zhao) All authors have read and agreed to the published version of the manuscript.

Acknowledgments

This work was supported by National Key Research and Development Program of China (2016YFA0300703, 2016YFA0202402), National Natural Science Foundation of China (11774064, 51971064, 61674111), the China Postdoctoral Science Foundation (Grant No. 2019M651942) and the “111” Project. The authors also acknowledge the financial support from Collaborative Innovation Center of Suzhou Nano Science and Technology, Soochow University and Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. National Renewable Energy Laboratory (NREL). Available online: https://www.nrel.gov/pv/cell-efficiency.html (accessed on 3 June 2020).
  2. Green, M.A.; Ho-Baillie, A.; Snaith, H.J. The emergence of perovskite solar cells. Nat. Photon. 2014, 8, 506–514. [Google Scholar] [CrossRef]
  3. Xing, G.; Mathews, N.; Sun, S.; Lim, S.S.; Lam, Y.M.; Grätzel, M.; Mhaisalkar, S.G.; Sum, T.C. Long-Range Balanced Electron-and Hole-Transport Lengths in Organic-Inorganic CH3NH3PbI3. Science 2013, 342, 344–347. [Google Scholar] [CrossRef] [PubMed]
  4. Stranks, S.D.; Eperon, G.E.; Grancini, G.; Menelaou, C.; Alcocer, M.J.P.; Leijtens, T.; Herz, L.M.; Petrozza, A.; Snaith, H.J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Stavrakas, C.; Delport, G.; Zhumekenov, A.A.; Anaya, M.; Chahbazian, R.; Bakr, O.M.; Barnard, E.S.; Stranks, S.D. Visualizing Buried Local Carrier Diffusion in Halide Perovskite Crystals via Two-Photon Microscopy. ACS Energy Lett. 2019, 5, 117–123. [Google Scholar] [CrossRef] [PubMed]
  6. Ambrosio, F.; Meggiolaro, D.; Mosconi, E.; De Angelis, F. Charge Localization, Stabilization, and Hopping in Lead Halide Perovskites: Competition between Polaron Stabilization and Cation Disorder. ACS Energy Lett. 2019, 4, 2013–2020. [Google Scholar] [CrossRef] [Green Version]
  7. Ponseca, C.S.; Savenije, T.J.; Abdellah, M.; Zheng, K.; Yartsev, A.; Pascher, T.; Harlang, T.C.B.; Chabera, P.; Pullerits, T.; Stepanov, A.; et al. Organometal Halide Perovskite Solar Cell Materials Rationalized: Ultrafast Charge Generation, High and Microsecond-Long Balanced Mobilities, and Slow Recombination. J. Am. Chem. Soc. 2014, 136, 5189–5192. [Google Scholar] [CrossRef]
  8. Ponseca, C.S.; Hutter, E.M.; Piatkowski, P.; Cohen, B.; Pascher, T.; Douhal, A.; Yartsev, A.; Sundström, V.; Savenije, T.J. Mechanism of Charge Transfer and Recombination Dynamics in Organo Metal Halide Perovskites and Organic Electrodes, PCBM, and Spiro-OMeTAD: Role of Dark Carriers. J. Am. Chem. Soc. 2015, 137, 16043–16048. [Google Scholar] [CrossRef]
  9. Piatkowski, P.; Cohen, B.; Ramos, F.J.; Di Nunzio, M.; Nazeeruddin, M.K.; Grätzel, M.; Ahmad, S.; Douhal, A. Direct monitoring of ultrafast electron and hole dynamics in perovskite solar cells. Phys. Chem. Chem. Phys. 2015, 17, 14674–14684. [Google Scholar] [CrossRef]
  10. Wang, L.; McCleese, C.; Kovalsky, A.; Zhao, Y.; Burda, C. Femtosecond Time-Resolved Transient Absorption Spectroscopy of CH3NH3PbI3 Perovskite Films: Evidence for Passivation Effect of PbI2. J. Am. Chem. Soc. 2014, 136, 12205–12208. [Google Scholar] [CrossRef]
  11. Zhu, Z.; Ma, J.; Wang, Z.; Mu, C.; Fan, Z.; Du, L.; Bai, Y.; Fan, L.; Yan, H.; Phillips, D.L.; et al. Efficiency Enhancement of Perovskite Solar Cells through Fast Electron Extraction: The Role of Graphene Quantum Dots. J. Am. Chem. Soc. 2014, 136, 3760–3763. [Google Scholar] [CrossRef]
  12. Xing, G.; Wu, B.; Chen, S.; Chua, J.; Yantara, N.; Mhaisalkar, S.; Mathews, N.; Sum, T.C. Interfacial electron transfer barrier at compact TiO2/CH3NH3PbI3 heterojunction. Small 2015, 11, 3606–3613. [Google Scholar] [CrossRef] [PubMed]
  13. Cha, J.-M.; Lee, J.-W.; Son, D.-Y.; Kim, H.-S.; Jang, I.-H.; Park, N.-G. Mesoscopic perovskite solar cells with an admixture of nanocrystalline TiO2 and Al2O3: Role of interconnectivity of TiO2 in charge collection. Nanoscale 2016, 8, 6341–6351. [Google Scholar] [CrossRef] [PubMed]
  14. Makuta, S.; Liu, M.; Nishimura, H.; Endo, M.; Wakamiya, A.; Tachibana, Y. Photo-excitation intensity dependent electron and hole injections from lead iodide perovskite to nanocrystalline TiO2 and spiro-OMeTAD. Chem. Commun. 2016, 52, 673–676. [Google Scholar] [CrossRef]
  15. Yuan, J.; Ling, X.; Yang, D.; Li, F.; Zhou, S.; Shi, J.; Qian, Y.; Hu, J.; Sun, Y.; Yang, Y.-G.; et al. Band-Aligned Polymeric Hole Transport Materials for Extremely Low Energy Loss α-CsPbI3 Perovskite Nanocrystal Solar Cells. Joule 2018, 2, 2450–2463. [Google Scholar] [CrossRef] [Green Version]
  16. Li, F.; Yuan, J.; Ling, X.; Zhang, Y.; Yang, Y.-G.; Cheung, S.H.; Ho, C.H.Y.; Gao, X.; Ma, W. A Universal Strategy to Utilize Polymeric Semiconductors for Perovskite Solar Cells with Enhanced Efficiency and Longevity. Adv. Funct. Mater. 2018, 28, 1706377. [Google Scholar] [CrossRef]
  17. Lee, J.-W.; Seol, D.-J.; Cho, A.-N.; Park, N.-G. High-Efficiency Perovskite Solar Cells Based on the Black Polymorph of HC(NH2)2PbI3. Adv. Mater. 2014, 26, 4991–4998. [Google Scholar] [CrossRef]
  18. Jeon, N.J.; Noh, J.H.; Yang, W.S.; Kim, Y.C.; Ryu, S.; Seo, J.; Seok, S.I. Compositional engineering of perovskite materials for high-performance solar cells. Nature 2015, 517, 476–480. [Google Scholar] [CrossRef]
  19. Eperon, G.E.; Stranks, S.D.; Menelaou, C.; Johnston, M.B.; Herz, L.M.; Snaith, H.J. Formamidinium lead trihalide: A broadly tunable perovskite for efficient planar heterojunction solar cells. Energy Environ. Sci. 2014, 7, 982. [Google Scholar] [CrossRef]
  20. Koh, T.M.; Fu, K.; Fang, Y.; Chen, S.; Sum, T.C.; Mathews, N.; Mhaisalkar, S.G.; Boix, P.P.; Baikie, T. Formamidinium-Containing Metal-Halide: An Alternative Material for Near-IR Absorption Perovskite Solar Cells. J. Phys. Chem. C 2013, 118, 16458–16462. [Google Scholar] [CrossRef]
  21. Pang, S.; Hu, H.; Zhang, J.; Lv, S.; Yu, Y.; Wei, F.; Qin, T.; Xu, H.; Liu, Z.; Cui, G. NH2CH=NH2PbI3: An Alternative Organolead Iodide Perovskite Sensitizer for Mesoscopic Solar Cells. Chem. Mater. 2014, 26, 1485–1491. [Google Scholar] [CrossRef]
  22. Zhao, Y.; Zhu, K. ChemInform Abstract: Organic-Inorganic Hybrid Lead Halide Perovskites for Optoelectronic and Electronic Applications. Chem. Soc. Rev. 2016, 47, 655–689. [Google Scholar] [CrossRef] [PubMed]
  23. Xue, J.; Lee, J.-W.; Dai, Z.; Wang, R.; Nuryyeva, S.; Liao, M.E.; Chang, S.-Y.; Meng, L.; Meng, D.; Sun, P.; et al. Surface Ligand Management for Stable FAPbI3 Perovskite Quantum Dot Solar Cells. Joule 2018, 2, 1866–1878. [Google Scholar] [CrossRef] [Green Version]
  24. Li, F.; Zhou, S.; Yuan, J.; Qin, C.; Yang, Y.; Shi, J.; Li, X.; Li, Y.; Ma, W. Perovskite quantum dot solar cells with 15.6% efficiency and improved stability enabled by an α-CsPbI3/FAPbI3 bilayer structure. ACS Energy Lett. 2019, 4, 2571–2578. [Google Scholar] [CrossRef]
  25. Ji, K.; Yuan, J.; Li, F.; Shi, Y.; Ling, X.; Zhang, X.; Zhang, Y.; Lu, H.; Yuan, J.; Ma, W. High-efficiency perovskite quantum dot solar cells benefiting from a conjugated polymer-quantum dot bulk heterojunction connecting layer. J. Mater. Chem. A 2020, 8, 8104–8112. [Google Scholar] [CrossRef]
  26. Ho-Baillie, A.W.Y.; Zhang, M.; Lau, C.F.J.; Ma, F.-J.; Huang, S. Untapped Potentials of Inorganic Metal Halide Perovskite Solar Cells. Joule 2019, 3, 938–955. [Google Scholar] [CrossRef] [Green Version]
  27. Yuan, M.; Liu, M.; Sargent, E.H. Colloidal quantum dot solids for solution-processed solar cells. Nat. Energy 2016, 1, 16016. [Google Scholar] [CrossRef]
  28. Shamsi, J.; Urban, A.S.; Imran, M.; De Trizio, L.; Manna, L. Metal Halide Perovskite Nanocrystals: Synthesis, Post-Synthesis Modifications, and Their Optical Properties. Chem. Rev. 2019, 119, 3296–3348. [Google Scholar] [CrossRef]
  29. Kovalenko, M.V.; Protesescu, L.; Bodnarchuk, M.I. Properties and potential optoelectronic applications of lead halide perovskite nanocrystals. Science 2017, 358, 745–750. [Google Scholar] [CrossRef] [Green Version]
  30. Akkerman, Q.A.; Rainò, G.; Kovalenko, M.V.; Manna, L. Genesis, challenges and opportunities for colloidal lead halide perovskite nanocrystals. Nat. Mater. 2018, 17, 394–405. [Google Scholar] [CrossRef]
  31. Miyata, A.; Mitioglu, A.; Plochocka, P.; Portugall, O.; Wang, J.T.-W.; Stranks, S.D.; Snaith, H.J.; Nicholas, R.J. Direct measurement of the exciton binding energy and effective masses for charge carriers in organic–inorganic tri-halide perovskites. Nat. Phys. 2015, 11, 582–587. [Google Scholar] [CrossRef] [Green Version]
  32. Xue, J.; Wang, R.; Chen, L.; Nuryyeva, S.; Han, T.; Huang, T.; Tan, S.; Zhu, J.; Wang, M.; Wang, Z.; et al. A Small-Molecule “Charge Driver” enables Perovskite Quantum Dot Solar Cells with Efficiency Approaching 13%. Adv. Mater. 2019, 31, 1900111. [Google Scholar] [CrossRef] [PubMed]
  33. Leventis, H.C.; O’Mahony, F.; Akhtar, J.; Afzaal, M.; O’Brien, K.; Haque, S.A. Transient Optical Studies of Interfacial Charge Transfer at Nanostructured Metal Oxide/PbS Quantum Dot/Organic Hole Conductor Heterojunctions. J. Am. Chem. Soc. 2010, 132, 2743–2750. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, Y.; Rodríguez-Córdoba, W.E.; Xiang, X.; Lian, T. Strong Electronic Coupling and Ultrafast Electron Transfer between PbS Quantum Dots and TiO2 Nanocrystalline Films. Nano Lett. 2011, 12, 303–309. [Google Scholar] [CrossRef]
  35. Brauer, J.C.; Lee, Y.H.; Nazeeruddin, M.K.; Banerji, N. Charge Transfer Dynamics from Organometal Halide Perovskite to Polymeric Hole Transport Materials in Hybrid Solar Cells. J. Phys. Chem. Lett. 2015, 6, 3675–3681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Wang, S.; Luo, Q.; Fang, W.-H.; Long, R. Interfacial Engineering Determines Band Alignment and Steers Charge Separation and Recombination at an Inorganic Perovskite Quantum Dot/WS2 Junction: A Time Domain Ab Initio Study. J. Phys. Chem. Lett. 2019, 10, 1234–1241. [Google Scholar] [CrossRef] [PubMed]
  37. Piatkowski, P.; Cohen, B.; Ponseca, C.S.; Salado, M.; Kazim, S.; Ahmad, S.; Sundström, V.; Douhal, A. Unraveling Charge Carriers Generation, Diffusion, and Recombination in Formamidinium Lead Triiodide Perovskite Polycrystalline Thin Film. J. Phys. Chem. Lett. 2015, 7, 204–210. [Google Scholar] [CrossRef]
  38. Stamplecoskie, K.G.; Manser, J.S.; Kamat, P. Dual nature of the excited state in organic–inorganic lead halide perovskites. Energy Environ. Sci. 2015, 8, 208–215. [Google Scholar] [CrossRef]
  39. Christians, J.A.; Manser, J.; Kamat, P.V. Multifaceted Excited State of CH3NH3PbI3. Charge Separation, Recombination, and Trapping. J. Phys. Chem. Lett. 2015, 6, 2086–2095. [Google Scholar] [CrossRef] [Green Version]
  40. Klimov, V.; Hunsche, S.; Kurz, H. Biexciton effects in femtosecond nonlinear transmission of semiconductor quantum dots. Phys. Rev. B 1994, 50, 8110–8113. [Google Scholar] [CrossRef]
  41. Klimov, V.I. Spectral and Dynamical Properties of Multiexcitons in Semiconductor Nanocrystals. Annu. Rev. Phys. Chem. 2007, 58, 635–673. [Google Scholar] [CrossRef]
  42. Yang, Y.; Ostrowski, D.P.; France, R.M.; Zhu, K.; Van De Lagemaat, J.; Luther, J.M.; Beard, M. Observation of a hot-phonon bottleneck in lead-iodide perovskites. Nat. Photon. 2015, 10, 53–59. [Google Scholar] [CrossRef]
  43. Yang, Y.; Yang, M.; Li, Z.; Crisp, R.W.; Zhu, K.; Beard, M. Comparison of Recombination Dynamics in CH3NH3PbBr3 and CH3NH3PbI3 Perovskite Films: Influence of Exciton Binding Energy. J. Phys. Chem. Lett. 2015, 6, 4688–4692. [Google Scholar] [CrossRef]
  44. Telfah, H.; Jamhawi, A.; Teunis, M.B.; Sardar, R.; Liu, J. Ultrafast Exciton Dynamics in Shape-Controlled Methylammonium Lead Bromide Perovskite Nanostructures: Effect of Quantum Confinement on Charge Carrier Recombination. J. Phys. Chem. C 2017, 121, 28556–28565. [Google Scholar] [CrossRef]
  45. Einevoll, G.T. Confinement of excitons in quantum dots. Phys. Rev. B 1992, 45, 3410–3417. [Google Scholar] [CrossRef]
  46. Piatkowski, P.; Cohen, B.; Kazim, S.; Ahmad, S.; Douhal, A. How photon pump fluence changes the charge carrier relaxation mechanism in an organic–inorganic hybrid lead triiodide perovskite. Phys. Chem. Chem. Phys. 2016, 18, 27090–27101. [Google Scholar] [CrossRef]
  47. Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J.-E.; et al. Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar Cell with Efficiency Exceeding 9%. Sci. Rep. 2012, 2, 591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Istrate, E.; Hoogland, S.; Sukhovatkin, V.; Levina, L.; Myrskog, S.; Smith, P.W.E.; Sargent, E.H. Carrier Relaxation Dynamics in Lead Sulfide Colloidal Quantum Dots. J. Phys. Chem. B 2008, 112, 2757–2760. [Google Scholar] [CrossRef] [PubMed]
  49. Simpson, M.J.; Doughty, B.; Yang, B.; Xiao, K.; Ma, Y.-Z. Spatial Localization of Excitons and Charge Carriers in Hybrid Perovskite Thin Films. J. Phys. Chem. Lett. 2015, 6, 3041–3047. [Google Scholar] [CrossRef]
  50. Saba, M.; Cadelano, M.; Marongiu, D.; Chen, F.; Sarritzu, V.; Sestu, N.; Figus, C.; Aresti, M.; Piras, R.; Lehmann, A.G.; et al. Correlated electron–hole plasma in organometal perovskites. Nat. Commun. 2014, 5, 5049. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) The structure of FAPbI3 PQDs layer grown with PTAA HTL and TiO2 ETL; (b) energy levels of TiO2, FAPbI3 PQDs, and PTAA used in this work; (c) transmission electron microscope (TEM) image of pure FAPbI3 PQDs film; (d) normalized UV−vis absorption spectra and PL spectra of FAPbI3 PQDs solutions.
Figure 1. (a) The structure of FAPbI3 PQDs layer grown with PTAA HTL and TiO2 ETL; (b) energy levels of TiO2, FAPbI3 PQDs, and PTAA used in this work; (c) transmission electron microscope (TEM) image of pure FAPbI3 PQDs film; (d) normalized UV−vis absorption spectra and PL spectra of FAPbI3 PQDs solutions.
Applsci 10 05553 g001
Figure 2. Differential transmission spectra of (a) pure FAPbI3 PQDs film excited at 400 nm with pump fluence of ≈70 μJ/cm2 and (b) PTAA HTL/FAPbI3 PQDs/TiO2 ETL excited at 400 nm with pump fluence of ≈21 μJ/cm2. The numbers marked in the upper right corner indicate the delay times between the pump and probe pulses. Transient ΔT/T traces of (c) pure FAPbI3 PQDs film (400 nm pump 14 μJ/cm2) and (d) PTAA HTL/FAPbI3 PQDs/TiO2 ETL (400 nm pump 21 μJ/cm2) probed at different wavelengths.
Figure 2. Differential transmission spectra of (a) pure FAPbI3 PQDs film excited at 400 nm with pump fluence of ≈70 μJ/cm2 and (b) PTAA HTL/FAPbI3 PQDs/TiO2 ETL excited at 400 nm with pump fluence of ≈21 μJ/cm2. The numbers marked in the upper right corner indicate the delay times between the pump and probe pulses. Transient ΔT/T traces of (c) pure FAPbI3 PQDs film (400 nm pump 14 μJ/cm2) and (d) PTAA HTL/FAPbI3 PQDs/TiO2 ETL (400 nm pump 21 μJ/cm2) probed at different wavelengths.
Applsci 10 05553 g002
Figure 3. Normalized kinetic traces of (a) PA probed at 570 nm and (b) PB probed at 770 nm for the films in vacuum after excitation at 400 nm with pump fluence of ≈3.5 μJ/cm2. The solid curves represent the fitting results.
Figure 3. Normalized kinetic traces of (a) PA probed at 570 nm and (b) PB probed at 770 nm for the films in vacuum after excitation at 400 nm with pump fluence of ≈3.5 μJ/cm2. The solid curves represent the fitting results.
Applsci 10 05553 g003
Figure 4. Experiment data of normalized (a) PA kinetic traces probed at 570 nm and (b) PB kinetic traces probed at 770 nm. All the measurements were carried out under the pump pulses of 400 nm with fluence of ≈3.5 μJ/cm2. The solid curves represent the fitting results. The fitting parameters are summarized in Table 1.
Figure 4. Experiment data of normalized (a) PA kinetic traces probed at 570 nm and (b) PB kinetic traces probed at 770 nm. All the measurements were carried out under the pump pulses of 400 nm with fluence of ≈3.5 μJ/cm2. The solid curves represent the fitting results. The fitting parameters are summarized in Table 1.
Applsci 10 05553 g004
Table 1. Parameters used to fit the temporal evolution of the PA and PB decay signals of the samples excited at 400 nm with pump fluence of ≈3.5 μJ/cm2.
Table 1. Parameters used to fit the temporal evolution of the PA and PB decay signals of the samples excited at 400 nm with pump fluence of ≈3.5 μJ/cm2.
SignalSamples A 1 A 1 + A 2 (%) τ 1   ( p s ) A 2 A 1 + A 2 (%) τ 2 ( p s )
PAFAPbI332.89121.6567.111563.98
PTAA/FAPbI332.31118.5367.691553.29
FAPbI3/TiO253.7118.9346.292602.56
PTAA/FAPbI3/TiO253.6819.9746.322567.01
PBFAPbI312.63117.1787.371656.98
PTAA/FAPbI314.01117.2185.991667.77
FAPbI3/TiO21002805.96
PTAA/FAPbI3/TiO21002873.01

Share and Cite

MDPI and ACS Style

Lou, Z.; Liang, S.; Yuan, J.; Ji, K.; Yuan, J.; Zhao, H.; Xia, H.; Ni, G.; Sheng, C.; Ma, W.; et al. Contrasting Electron and Hole Transfer Dynamics from CH(NH2)2PbI3 Perovskite Quantum Dots to Charge Transport Layers. Appl. Sci. 2020, 10, 5553. https://doi.org/10.3390/app10165553

AMA Style

Lou Z, Liang S, Yuan J, Ji K, Yuan J, Zhao H, Xia H, Ni G, Sheng C, Ma W, et al. Contrasting Electron and Hole Transfer Dynamics from CH(NH2)2PbI3 Perovskite Quantum Dots to Charge Transport Layers. Applied Sciences. 2020; 10(16):5553. https://doi.org/10.3390/app10165553

Chicago/Turabian Style

Lou, Zhigang, Shuyan Liang, Jiabei Yuan, Kang Ji, Jianyu Yuan, Hongchang Zhao, Hong Xia, Gang Ni, Chuanxiang Sheng, Wanli Ma, and et al. 2020. "Contrasting Electron and Hole Transfer Dynamics from CH(NH2)2PbI3 Perovskite Quantum Dots to Charge Transport Layers" Applied Sciences 10, no. 16: 5553. https://doi.org/10.3390/app10165553

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop