Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Effect of surface recombination in high performance white-light CH3NH3PbI3 single crystal photodetectors

Open Access Open Access

Abstract

Methylammonium lead iodide (CH3NH3PbI3), with the organic-inorganic hybrid perovskite (OIHP) structure, has gained tremendous research interest due to its excellent photo-electron conversion ability in the application of photovoltaics. Despite its solution processed polycrystalline thin film form in solar cells, the single crystalline counterpart may offer some incredibly novel optoelectronic functionalities. In this work, a sizable (>5 mm) and high quality CH3NH3PbI3 single crystal has been synthesized by the inverse temperature crystallization method, and a white-light photodetector of the structure glass/ITO/Ga/ CH3NH3PbI3/Au was fabricated. Overbroad photo-excitation intensities ranging from 0.1 mW/cm2 to 100 mW/cm2 using a sun-light simulator, the on-off ratio is tunable in a wide-range from 65 to 2250 at zero bias voltage. The responsivity (R) and detectivity (D*) are 36.2 mA/W and 2.68×1011 Jones respectively at a weak white-light intensity such as 0.1 mW/cm2. Both the photodetective parameters decrease with the increase of the illumination intensity. Based on impedance spectra obtained at working condition and light intensity dependent Jsc measurements, the surface trap-assist recombination may play a dominating role. The corresponding lifetime (τsurf) and resistance (Rsurf_trap) exhibit fast decays at higher illumination intensities. This fundamental study may pave the way for exploring the contribution of the surface trap-assist recombination in the CH3NH3PbI3 single crystal based photodetector. We believe it is applicable for integration in micro-photonics for sensitive and weak white-light photo-detection.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Solution-processed organic-inorganic hybrid perovskites (OIHPs) have gained unprecedented attention in photonic and electronic applications for developing high performance solar cells, light emitting diodes (LEDs), field effect transistors (FETs) and solid state lasers. These are primarily due to their excellent optoelectronic properties, such as broad photon absorption spectra, direct band-gap, long charge carrier diffusion lengths and relatively small effective masses for electron and hole [1–3]. By far, the polycrystalline thin film form of the OIHPs still plays the dominating role in photonic devices. Many efforts have been devoted to synthesize and optimize high quality polycrystalline films via various methods such as post-treatment [4,5], newly synthesized perovskite precursors [6,7], organic additives [8,9], different fabrication methods [10,11] and organic solvent engineering [12,13], in order to extensively elaborate device performance. Nevertheless, it is generally agreed that the polycrystalline film form of the OIHPs may contain a large number of traps/defects existing at surfaces and grain boundaries; consequently, charge transport scattering, trap-assist recombination and ion migration via the grain boundaries are expected to be responsible for the device degradation [14–16].

Fortunately, some experimental studies have shown large single crystalline OIHPs can be synthesized via the solution method even in the ambient temperature. It brings us a new opportunity to investigate electronic transport behaviors in single crystalline OIHPs based optoelectronic devices in order to explore some novel and multi-functional properties of the devices. For example, a CH(NH2)2PbI3 single crystal based photodetector was made showing a low noise current at low frequencies [17]. A high gain of more than 104 electrons per photon, and large MAPbBr3 single crystals based photodetector was demonstrated [18]. The CH3NH3PbI3 single crystal was also reported for a self-powered photodetector [19]. A perovskite of the cubic structure was made in order to boost performance and stability of a photodiode [20]. By bulk defect reduction and surface passivation, a high mobility of 1.2 × 10−2 cm2V−1 and small surface charge recombination velocity of 64 cms−1 were achieved [21]. In comparison to polycrystalline OIHPs in photovoltaic and light emitting devices, the single crystalline counterparts are particularly favorable in the applications for photodetector and photodiode due to their low bulk defect density, strong visible light absorption abilities, fast photo-electrical responses, high mobility and excellent stability in the air.

In this work, we have successfully synthesized and fabricated a sizable (>5 mm) CH3NH3PbI3 single crystal with low trap density and high mobility based white-light photodetector comprising a simple vertical structure glass/ITO/Ga/CH3NH3PbI3/Au. A large tunable on-off ratio can be achieved from 65 to 2250 at zero bias. The photodetector exhibits a relatively high responsivity (R) with the best record of 36.2 mA/W under weak white-light excitation. Despite this, the detectivity (D*) was measured to be approximately 2.68 × 1011 Jones at the same condition. Both of them start to decrease with the increase of the illumination intensity. Based on ac-modulated impedance spectroscopic measurements and illumination intensity dependence of Jsc, the behavior has been attributed to the presence of surface trap-related electronic transport process.

2. Experiments

2.1 CH3NH3PbI3 single crystal synthesis

The CH3NH3PbI3 single crystal was grown by the inverse temperature crystallization method [22]. During the synthesis, a CH3NH3PbI3 precursor was prepared by mixing CH3NH3I and PbI2 with a molar rate of 1:1 in the GBL organic solvent. The solution was stirred at 60 °C for approximately 2 hours, and then it was filtered by a PTFE filter with a 0.2 μm pore size. A 2 ml filtrate was placed in a vial and it was then kept in an oil bath at 110 °C. It took approximately 6 hours to complete the growth of the single crystal. After this, the diethyl ether (DEE) and isopropyl alcohol (IPA) were used to clean the crystal in order to remove the remaining precursors.

2.2 Photodetector fabrication

During the device fabrication, the CH3NH3PbI3 single crystal was mounted on a lapping fixture using a paraffin wax. Then, it was polished by SiC abrasive papers (2000, 5000, and 7000 grit in order). After this, some residual paraffin waxes were thoroughly washed away by a warm (80 °C) isooctane. Finally, on top of the ITO coated glass, the Ga and Au were made as the bottom and top electrodes respectively for the photodetector.

2.3 Material and device characterizations

Steady state photoluminescence (SSPL) spectra were recorded by a fluorescence spectrometer (Jobinyvon Horiba Fluorolog-3). Time-resolved photoluminescence (TRPL) spectra were measured by a single photon counting system, and a 405 nm pulsed laser diode was utilized as the photo-excitation source. X-ray diffraction (XRD) was performed by an X-ray diffracted spectrometer (Bruker D-8 Advance). Photocurrent responses of the photodetector were recorded by a source meter unit (Keysight B2912A) and a standard solar simulator was used as the photo-excitation source. Different illumination intensities were adjusted by optical density (OD) filters. Impedance spectroscopic measurements were performed by an LCR impedance analyzer (Agilent E4990A).

3. Results and discussion

We started with basic characterizations of the single crystal CH3NH3PbI3. Figure 1(a) provides the photographic image of the solution-processed single crystal with the planar size of more than 5 mm. Owing to the single crystallinity, both the XRD and SSPL spectra are distinctly different from its polycrystalline form in Figs. 1(b) and 1(c) respectively. The XRD spectrum of the single crystal CH3NH3PbI3 (i.e. the bottom one) reveals less diffracted peaks by comparing to its polycrystalline film (i.e. the top one) indicating the formation of less impurity crystalline phases [23,24]. In Fig. 1(c), the normalized SSPL spectrum locates at 764 nm (red) for the CH3NH3PbI3 single crystal with respect to the one of the polycrystalline film at 760 nm (blue). The relative red-shift of the spectrum is ascribed to the increase of the grain size since a relatively weaker lattice distortion appears in the crystal with large grain size; as a consequence, it results in a smaller bandgap [24,25]. Figure 1(d) shows the TRPL spectra for both the polycrystalline film and the single crystal respectively. Their corresponding lifetimes are extracted by fitting the decay curves using the bi-exponential model Y = A1exp(tτ1)+A2exp(tτ2)+y0, in which, A1 and A2 denote fractions for a fast (τ1) and slow (τ2) decay times respectively. As we can see from Table 1, the CH3NH3PbI3 single crystal has the lifetime of approximately 67 ns, in contrast, the one of the polycrystalline film is decreased to 11 ns. It is, thus, possible to obtain a much longer electron-hole lifetime in the CH3NH3PbI3 single crystal.

 figure: Fig. 1

Fig. 1 (a) The photographic image of the solution-processed CH3NH3PbI3 single crystal. (b) XRD spectra for the CH3NH3PbI3 single crystal (bottom) and polycrystalline film (top) respectively. (c) Steady-state and (d) time-resolved photoluminescence spectra for the CH3NH3PbI3 single crystal and polycrystalline film.

Download Full Size | PDF

Tables Icon

Table 1. Fitting Parameters of the TRPL Spectra in Fig. 1(d) for the CH3NH3PbI3 Single Crystal and Polycrystalline Film.

The trap density (ntrap) of the CH3NH3PbI3 single crystal can be evaluated by the J-V characterization method for a hole-only device such as Au/CH3NH3PbI3/Au [Fig. 2(a)] in the dark condition. The corresponding J-V characteristic curve is shown in Fig. 2(b), and it can be distinguished by three different regions (I, II and III). At the low electric field strength (Region I), Jsc increases linearly with V indicating the Ohmic response for the device. The conductivity (σ) of the single crystal was estimated to be approximately 4.6×104 Scm1  for n = 1, which is higher than the one of a typical CH3NH3PbI3 polycrystalline film (1.5×105 Scm1) [25]. At the moderate bias range (Region II), the rate of the increase of the Jsc becomes fast for n > 2. When V > 3.7 V at the kink point, it is expected that all traps in the single crystal are completely filled by injected charge carriers. The applied voltage at the point is defined as the trap-filled limit voltage (VTFL), which can be found by the expression VTFL=entrapL2εε0 [26], where e is the elementary charge of the electron, L is the thickness of the CH3NH3PbI3 single crystal, ε is the relative dielectric constant for CH3NH3PbI3 (ε = 32.1) [27], ε0 is the dielectric constant in the free space. In this case, the VTFL was estimated to be approximately 3.7 V for the CH3NH3PbI3 single crystal, and ntrap was calculated to be approximately 1.3 × 1010 cm−3. It is much lower than the one of the CH3NH3PbI3 polycrystalline film (8.9 × 1016 cm−3) [27]. At higher bias voltage (Region III), the J-V characteristic curve is quadratic dependent (n = 2), μ can be calculated by the Mott-Gurney Law μ=8JDL39εε0V2 after fitting [28], and JD is the dark current density. In this case, μ is approximately equal to 37.6 cm2V−1s−1 which is 4 orders of magnitude higher than the one of the polycrystalline CH3NH3PbI3 film (3.9 × 10−3 cm2V−1s−1) [29]. In spite of this, the properties of some CH3NH3PbI3 single crystals that have been reported in literatures are summarized in Table 2. In comparisons, our single crystal exhibits the lowest ntrap using the ITC method. Reference [30] shows even higher μ than ours; however, σ is known from that work. Other methods, such as AVC and TSSG have been reported for the syntheses of the single crystal as well. As the data shown in Table 2, the CH3NH3PbI3 single crystals made by these two methods contain higher ntrap than ours.

 figure: Fig. 2

Fig. 2 (a) A schematic drawing for the hole-only device structure. (b) A J-V characteristic curve for the hole-only device, VTFL stands for trap-filled limit voltage.

Download Full Size | PDF

Tables Icon

Table 2. Parameters were Obtained by Fitting the J-V Curve of Fig. 2(b). ITC, AVC and TSSG Denote Inverted Temperature Crystallization, Antisolvent Vapor-assisted Crystallization and Top-Seeded Solution Growth, Respectively.

Based on the as-prepared single crystal, the photodetector was fabricated and Fig. 3(a) schematically shows the device structure consisting of glass/ITO/Ga/CH3NH3PbI3/Au. The illumination intensity dependence of the J-V characteristic curves for the device is given in Fig. 3(b). Apparently, it produces a very low JD of approximately 1.9 × 10−4 mA/cm2 at 0 V (inset). Such property is the essential prerequisite for a photodetector to achieve an outstanding performance and reliability [34]. With the increase of the illumination intensity, J increases dramatically with the increase of V, for example J is about 1×101 mA/cm2 under 100 mW/cm2 illumination at 0 V. Figure 3(c) provides light intensity dependence of Jsc and their linear relationship can be fitted by the formula Jsc=Pα [35]. The deviation of the α parameter from 0.5 indicates the presence of trap-assist recombination process. In this case, we found α=0.62. Because of the lack of bulk trap states in the CH3NH3PbI3 single crystal, we have assumed that surface trap states may dominate.

 figure: Fig. 3

Fig. 3 (a) A schematic drawing of the photodetector glass/ITO/Ga/CH3NH3PbI3/Au. (b) Illumination intensity dependence of J-V measurements, the inset shows the zoom-in on V close to 0. (c) Illumination intensity dependence of Jsc.

Download Full Size | PDF

The photodetective switching behavior is a unique feature to evaluate the performance of the photodetector. Here, we have measured the time-dependent photocurrent at the Jsc condition under different illumination intensities using the white-light. In Fig. 4(a), the Jsc increases with the rising illumination intensity from 0.1 mW/cm2 to 100 mW/cm2. Here, it undergoes the switching on-off for 4 cycles. An average Jsc of 132 μA/cm2 was achieved under 100 mW/cm2 illumination intensity, showing an excellent on-off ratio of 2250. It is worthwhile to mention that the photodetector did not show any noticeable degradation during the measurements and all the JD coincide at the same level. In order to quantitatively evaluate the optoelectronic performance of the CH3NH3PbI3 single crystal based photodetector, the three critical parameters such as responsivity (R), detectivity (D*) and on-off ratio were calculated at different illumination intensities ranging from 0.1 mW/cm2 to 100 mW/cm2 under white-light at 0 V bias, and the results are given in Figs. 4(b)-4(d) respectively [36]. R is defined as = JlightJDPin , in which Pin is incident light intensity. D* can be expressed as D*= R2eJD. Clearly, both R and D* decrease with the increase of JD. The maximum R and D* are approximately equal to 36.2 mA/W and 2.68 × 1011 Jones, respectively at 0.1 mW/cm2. In comparison, the on-off ratio keeps rising until a maximum value of 2250 at 100 mW/cm2.

 figure: Fig. 4

Fig. 4 (a) Illumination intensity dependence of the switching behavior at 0 V for the CH3NH3PbI3 single crystal based photodetector. Extracted values are for (b) responsivity R, (c) detectivity D* and (d) on-off ratios.

Download Full Size | PDF

The illumination intensity dependence of Jsc [Fig. 3(c)] can be used to evaluate the trap-assist recombination process in an optoelectronic device. As it has been discussed above, the white-light intensity dependence of the on-off ratio, D* and R, are of critically importance for the determinations of the outputs of the single crystal photodetector. A complete understanding of the physical phenomena behind requires the non-destructive ac-modulation impedance spectroscopic characterization for the photodetector at working condition. Figure 5(a) displays the illumination intensity dependence of the impedance spectra for the photodetector consisting of glass/ITO/Ga/CH3NH3PbI3/Au, and all the spectra were fitted by the equivalent electronic circuit in Fig. 5(b). In general, it contains two sets of the R-C circuits connected in parallel, Rs denotes the series resistance, the two capacitors stand for the surface (Csurf) and bulk capacitance (Cbulk), respectively [37]. In this case, Cbulk is represented by a constant phase elements (CPE) and it is usually determined by two major components, CPE-T and CPE-P. The value of CPE-P that varies from 0.5 to 1 decides whether the CPE-T behaves like a capacitor or a resistor. Rbulk and Rsurf are the bulk and surface resistances respectively. Despite these, Rsurf_trap and Csurf_trap are the surface trap-related resistance and capacitance respectively [38]. All the fitting parameters have been summarized in Table 3. The surface charge and bulk charge recombination lifetimes (τsurf,  τbulk) that could be extracted from the impedance spectra were plotted in Fig. 5(c) [39]. As we can see, both of them tend to decrease with the increase of the white-light intensity; nevertheless, τsurf exhibits two distinct features by comparing with  τbulk, (i) τsurf is much shorter than  τbulk within the complete white-light intensity range; (ii) the reduction of τsurf is faster than  τbulk at higher illumination intensities. The results elucidate that the surface recombination is expected to play dominating role for the performance of the photodetector at the Jsc condition. Moreover, Rsurf_trap and Csurf_trap of Fig. 5(d) evidently reveals that the surface-trap associated recombination resistance decreases with the rising white-light intensity. Owing to the occupancy of the trap-states by photo-excited charge carriers at higher illumination intensities, Csurf_trap is increased. In the comparison with Figs. 4(b) and 4(c), the decrease of the photodeteictive parameters R and D* at higher illumination intensities can be explained by considering the effect of surface trap-assist recombination process at higher white-light intensities causing a significant reduction of the photo-generated current. In fact, the effect can be also reflected from the aforementioned result in Fig. 3(c) since the σ-parameter is in-between 0.5 and 1.0 indicating the presence of trap-related recombination process.

 figure: Fig. 5

Fig. 5 (a) White-light illumination intensity dependence of Nyquist plots for the photodetector consisting of glass/ITO/Ga/ CH3NH3PbI3/Au. (b) An equivalent electronic circuit for fitting the spectra. (c) Light intensity dependence of  τbulk and τsurf. (d) Light intensity dependent of Rsurf_trap and Csurf_trap.

Download Full Size | PDF

Tables Icon

Table 3. A Summary of Impedance Parameters Fitted from Fig. 5(a).

4. Conclusion

In summary, we have synthesized the CH3NH3PbI3 single crystal with relatively lower trap density ntrap of approximately 1.3 × 1010 cm−3, higher conductivity σ=4.6×104 Scm1  and mobility μ=37.6 cm2V1s1 by the ITC method. Based on this, the white-light photodetector of the structure glass/ITO/Ga/CH3NH3PbI3/Au was fabricated. It exhibits remarkably large R (36.2 mA/W) and D* (2.68×1011 Jones) under weak white-light photo-excitation intensity such as 0.1 mW/cm2 at zero bias. With the assist of impedance spectroscopic measurements, the reductions of the R and D* are mainly due to the enhancement surface trap-related recombination process. We believe this work serves as a fundamental study for the surface trap associated recombination mechanism in the CH3NH3PbI3 single crystal based photodetector. The device is applicable for sensitive and weak white-light photo-detection.

Funding

The National Natural Science Foundation of China (Grant No. 61604010, 61634001, U1601651), and the research funding from Beijing Jiaotong University Research Program (S18JB00020).

References

1. H. Cho, S.-H. Jeong, M.-H. Park, Y.-H. Kim, C. Wolf, C.-L. Lee, J. H. Heo, A. Sadhanala, N. Myoung, S. Yoo, S. H. Im, R. H. Friend, and T. W. Lee, “Overcoming the electroluminescence efficiency limitations of perovskite light-emitting diodes,” Science 350(6265), 1222–1225 (2015). [CrossRef]   [PubMed]  

2. H. Zhu, Y. Fu, F. Meng, X. Wu, Z. Gong, Q. Ding, M. V. Gustafsson, M. T. Trinh, S. Jin, and X. Y. Zhu, “Lead halide perovskite nanowire lasers with low lasing thresholds and high quality factors,” Nat. Mater. 14(6), 636–642 (2015). [CrossRef]   [PubMed]  

3. W. S. Yang, B.-W. Park, E. H. Jung, N. J. Jeon, Y. C. Kim, D. U. Lee, S. S. Shin, J. Seo, E. K. Kim, J. H. Noh, and S. I. Seok, “Iodide management in formamidinium-lead-halide-based perovskite layers for efficient solar cells,” Science 356(6345), 1376–1379 (2017). [CrossRef]   [PubMed]  

4. A. Dualeh, N. Tétreault, T. Moehl, P. Gao, M. K. Nazeeruddin, and M. Grätzel, “Effect of annealing temperature on film morphology of organic-inorganic hybrid pervoskite solid-state solar cells,” Adv. Funct. Mater. 24(21), 3250–3258 (2014). [CrossRef]  

5. S. Aharon, A. Dymshits, A. Rotem, and L. Etgar, “Temperature dependence of hole conductor free formamidinium lead iodide perovskite based solar cells,” J. Mater. Chem. A Mater. Energy Sustain. 3(17), 9171–9178 (2015). [CrossRef]  

6. Q. Chen, H. Zhou, Y. Fang, A. Z. Stieg, T. B. Song, H. H. Wang, X. Xu, Y. Liu, S. Lu, J. You, P. Sun, J. McKay, M. S. Goorsky, and Y. Yang, “The optoelectronic role of chlorine in CH3NH3PbI3(Cl)-based perovskite solar cells,” Nat. Commun. 6(1), 7269 (2015). [CrossRef]   [PubMed]  

7. F. Jiang, Y. Rong, H. Liu, T. Liu, L. Mao, W. Meng, F. Qin, Y. Jiang, B. Luo, S. Xiong, J. Tong, Y. Liu, Z. Li, H. Han, and Y. Zhou, “Synergistic effect of PbI2passivation and chlorine inclusion yielding high open-circuit voltage exceeding 1.15 V in both mesoscopic and inverted planar CH3NH3PbI3(Cl)-based perovskite solar cells,” Adv. Funct. Mater. 26(44), 8119–8127 (2016). [CrossRef]  

8. P. W. Liang, C. Y. Liao, C. C. Chueh, F. Zuo, S. T. Williams, X. K. Xin, J. Lin, and A. K. Jen, “Additive enhanced crystallization of solution-processed perovskite for highly efficient planar-heterojunction solar cells,” Adv. Mater. 26(22), 3748–3754 (2014). [CrossRef]   [PubMed]  

9. C.-G. Wu, C.-H. Chiang, Z.-L. Tseng, M. K. Nazeeruddin, A. Hagfeldt, and M. Grätzel, “High efficiency stable inverted perovskite solar cells without current hysteresis,” Energy Environ. Sci. 8(9), 2725–2733 (2015). [CrossRef]  

10. S. D. Stranks, P. K. Nayak, W. Zhang, T. Stergiopoulos, and H. J. Snaith, “Formation of thin films of organic-inorganic perovskites for high-efficiency solar cells,” Angew. Chem. Int. Ed. Engl. 54(11), 3240–3248 (2015). [CrossRef]   [PubMed]  

11. T.-B. Song, Q. Chen, H. Zhou, C. Jiang, H.-H. Wang, Y. M. Yang, Y. Liu, J. You, and Y. Yang, “Perovskite solar cells: Film formation and properties,” J. Mater. Chem. A Mater. Energy Sustain. 3(17), 9032–9050 (2015). [CrossRef]  

12. N. Ahn, D. Y. Son, I. H. Jang, S. M. Kang, M. Choi, and N. G. Park, “Highly reproducible perovskite solar cells with average efficiency of 18.3% and best efficiency of 19.7% fabricated via lewis base adduct of lead(II) iodide,” J. Am. Chem. Soc. 137(27), 8696–8699 (2015). [CrossRef]   [PubMed]  

13. Q. Chen, N. D. Marco, Y. M. Yang, T.-B. Song, C.-C. Chen, H. Zhao, Z. Hong, H. Zhou, and Y. Yang, “Under the spotlight: The organic-inorganic hybrid halide perovskite for optoelectronic applications,” Nano Today 10(3), 355–396 (2015). [CrossRef]  

14. J. Yang, B. D. Siempelkamp, D. Liu, and T. L. Kelly, “Investigation of CH3NH3PbI3 degradation rates and mechanisms in controlled humidity environments using in situ techniques,” ACS Nano 9(2), 1955–1963 (2015). [CrossRef]   [PubMed]  

15. Y. Shao, Z. Xiao, C. Bi, Y. Yuan, and J. Huang, “Origin and elimination of photocurrent hysteresis by fullerene passivation in CH3NH3PbI3 planar heterojunction solar cells,” Nat. Commun. 5(1), 5784 (2014). [CrossRef]   [PubMed]  

16. Q. Wang, Y. Shao, Q. Dong, Z. Xiao, Y. Yuan, and J. Huang, “Large fill-factor bilayer iodine perovskite solar cells fabricated by a low-temperature solution-process,” Energy Environ. Sci. 7(7), 2359–2365 (2014). [CrossRef]  

17. Q. Han, S. H. Bae, P. Sun, Y. T. Hsieh, Y. M. Yang, Y. S. Rim, H. Zhao, Q. Chen, W. Shi, G. Li, and Y. Yang, “Single crystal formamidinium lead iodide (FAPbI3): Insight into the structural, optical, and electrical properties,” Adv. Mater. 28(11), 2253–2258 (2016). [CrossRef]   [PubMed]  

18. M. I. Saidaminov, V. Adinolfi, R. Comin, A. L. Abdelhady, W. Peng, I. Dursun, M. Yuan, S. Hoogland, E. H. Sargent, and O. M. Bakr, “Planar-integrated single-crystalline perovskite photodetectors,” Nat. Commun. 6(1), 8724 (2015). [CrossRef]   [PubMed]  

19. J. Ding, H. Fang, Z. Lian, J. Li, Q. Lv, L. Wang, J.-L. Sun, and Q. Yan, “A self-powered photodetector based on a CH3NH3PbI3 single crystal with asymmetric electrodes,” CrystEngComm 18(23), 4405–4411 (2016). [CrossRef]  

20. Z. Shi, Y. Zhang, C. Cui, B. Li, W. Zhou, Z. Ning, and Q. Mi, “Symmetrization of the crystal lattice of MAPbI3 boosts the performance and stability of metal-perovskite photodiodes,” Adv. Mater. 29(30), 1701656 (2017). [CrossRef]   [PubMed]  

21. H. Wei, Y. Fang, P. Mulligan, W. Chuirazzi, H.-H. Fang, C. Wang, B. R. Ecker, Y. Gao, M. A. Loi, L. Cao, and J. Huang, “Sensitive X-ray detectors made of methylammonium lead tribromide perovskite single crystals,” Nat. Photonics 10(5), 333–339 (2016). [CrossRef]  

22. M. I. Saidaminov, A. L. Abdelhady, B. Murali, E. Alarousu, V. M. Burlakov, W. Peng, I. Dursun, L. Wang, Y. He, G. Maculan, A. Goriely, T. Wu, O. F. Mohammed, and O. M. Bakr, “High-quality bulk hybrid perovskite single crystals within minutes by inverse temperature crystallization,” Nat. Commun. 6(1), 7586 (2015). [CrossRef]   [PubMed]  

23. B. Wu, H. T. Nguyen, Z. Ku, G. Han, D. Giovanni, N. Mathews, H. J. Fan, and T. C. Sum, “Discerning the surface and bulk recombination kinetics of organic-inorganic halide perovskite single crystals,” Adv. Energy Mater. 6(14), 1600551 (2016). [CrossRef]  

24. V. D’Innocenzo, A. R. Srimath Kandada, M. De Bastiani, M. Gandini, and A. Petrozza, “Tuning the light emission properties by band gap engineering in hybrid lead halide perovskite,” J. Am. Chem. Soc. 136(51), 17730–17733 (2014). [CrossRef]   [PubMed]  

25. J. H. Heo, H. J. Han, D. Kim, T. K. Ahn, and S. H. Im, “Hysteresis-less inverted CH3NH3PbI3 planar perovskite hybrid solar cells with 18.1% power conversion efficiency,” Energy Environ. Sci. 8(5), 1602–1608 (2015). [CrossRef]  

26. R. H. Bube, “Trap density determination by space-charge-limited currents,” J. Appl. Phys. 33(5), 1733–1737 (1962). [CrossRef]  

27. C. Han, K. Wang, X. Zhu, H. Yu, X. Sun, Q. Yang, and B. Hu, “Unraveling surface and bulk trap states in lead halide perovskite solar cells using impedance spectroscopy,” J. Phys. D Appl. Phys. 51(9), 095501 (2018). [CrossRef]  

28. Y. Liu, J. Sun, Z. Yang, D. Yang, X. Ren, H. Xu, Z. Yang, and S. F. Liu, “20-mm-large single-crystalline formamidinium-perovskite wafer for mass production of integrated photodetectors,” Adv. Opt. Mater. 4(11), 1829–1837 (2016). [CrossRef]  

29. G. Giorgi and K. Yamashita, “Organic-inorganic halide perovskites: An ambipolar class of materials with enhanced photovoltaic performances,” J. Mater. Chem. A Mater. Energy Sustain. 3(17), 8981–8991 (2015). [CrossRef]  

30. M. I. Saidaminov, A. L. Abdelhady, B. Murali, E. Alarousu, V. M. Burlakov, W. Peng, I. Dursun, L. Wang, Y. He, G. Maculan, A. Goriely, T. Wu, O. F. Mohammed, and O. M. Bakr, “High-quality bulk hybrid perovskite single crystals within minutes by inverse temperature crystallization,” Nat. Commun. 6(1), 7586 (2015). [CrossRef]   [PubMed]  

31. Y. Liu, Z. Yang, D. Cui, X. Ren, J. Sun, X. Liu, J. Zhang, Q. Wei, H. Fan, F. Yu, X. Zhang, C. Zhao, and S. F. Liu, “Two-Inch-Sized Perovskite CH3 NH3 PbX3 (X = Cl, Br, I) Crystals: Growth and Characterization,” Adv. Mater. 27(35), 5176–5183 (2015). [CrossRef]   [PubMed]  

32. D. Shi, V. Adinolfi, R. Comin, M. Yuan, E. Alarousu, A. Buin, Y. Chen, S. Hoogland, A. Rothenberger, K. Katsiev, Y. Losovyj, X. Zhang, P. A. Dowben, O. F. Mohammed, E. H. Sargent, and O. M. Bakr, “Solar cells. Low trap-state density and long carrier diffusion in organolead trihalide perovskite single crystals,” Science 347(6221), 519–522 (2015). [CrossRef]   [PubMed]  

33. Q. Dong, Y. Fang, Y. Shao, P. Mulligan, J. Qiu, L. Cao, and J. Huang, “Electron-hole diffusion lengths > 175 μm in solution-grown CH3NH3PbI3 single crystals,” Science 347(6225), 967–970 (2015). [CrossRef]   [PubMed]  

34. H. Huang, Y. Xie, W. Yang, F. Zhang, J. Cai, and Z. Wu, “Low-dark-current TiO2 msm uv photodetectors with Pt schottky contacts,” IEEE Electron Device Lett. 32(4), 530–532 (2011). [CrossRef]  

35. I. Riedel, J. Parisi, V. Dyakonov, L. Lutsen, D. Vanderzande, and J. C. Hummelen, “Effect of temperature and illumination on the electrical characteristics of polymer-fullerene bulk-heterojunction solar cells,” Adv. Funct. Mater. 14(1), 38–44 (2004). [CrossRef]  

36. J. Zhou and J. Huang, “Photodetectors based on organic-inorganic hybrid lead halide perovskites,” Adv Sci (Weinh) 5(1), 1700256 (2018). [CrossRef]   [PubMed]  

37. W. L. Leong, S. R. Cowan, and A. J. Heeger, “Differential resistance analysis of charge carrier losses in organic bulk heterojunction solar cells: Observing the transition from bimolecular to trap-assisted recombination and quantifying the order of recombination,” Adv. Energy Mater. 1(4), 517–522 (2011). [CrossRef]  

38. J. Bisquert, “Beyond the quasistatic approximation: Impedance and capacitance of an exponential distribution of traps,” Phys. Rev. B 77(23), 235203 (2008). [CrossRef]  

39. I. Zarazua, G. Han, P. P. Boix, S. Mhaisalkar, F. Fabregat-Santiago, I. Mora-Seró, J. Bisquert, and G. Garcia-Belmonte, “Surface recombination and collection efficiency in perovskite solar cells from impedance analysis,” J. Phys. Chem. Lett. 7(24), 5105–5113 (2016). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 (a) The photographic image of the solution-processed CH3NH3PbI3 single crystal. (b) XRD spectra for the CH3NH3PbI3 single crystal (bottom) and polycrystalline film (top) respectively. (c) Steady-state and (d) time-resolved photoluminescence spectra for the CH3NH3PbI3 single crystal and polycrystalline film.
Fig. 2
Fig. 2 (a) A schematic drawing for the hole-only device structure. (b) A J-V characteristic curve for the hole-only device, V TFL stands for trap-filled limit voltage.
Fig. 3
Fig. 3 (a) A schematic drawing of the photodetector glass/ITO/Ga/CH3NH3PbI3/Au. (b) Illumination intensity dependence of J-V measurements, the inset shows the zoom-in on V close to 0. (c) Illumination intensity dependence of J sc .
Fig. 4
Fig. 4 (a) Illumination intensity dependence of the switching behavior at 0 V for the CH3NH3PbI3 single crystal based photodetector. Extracted values are for (b) responsivity R, (c) detectivity D* and (d) on-off ratios.
Fig. 5
Fig. 5 (a) White-light illumination intensity dependence of Nyquist plots for the photodetector consisting of glass/ITO/Ga/ CH3NH3PbI3/Au. (b) An equivalent electronic circuit for fitting the spectra. (c) Light intensity dependence of   τ bulk and τ surf . (d) Light intensity dependent of R surf_trap and C surf_trap .

Tables (3)

Tables Icon

Table 1 Fitting Parameters of the TRPL Spectra in Fig. 1(d) for the CH3NH3PbI3 Single Crystal and Polycrystalline Film.

Tables Icon

Table 2 Parameters were Obtained by Fitting the J-V Curve of Fig. 2(b). ITC, AVC and TSSG Denote Inverted Temperature Crystallization, Antisolvent Vapor-assisted Crystallization and Top-Seeded Solution Growth, Respectively.

Tables Icon

Table 3 A Summary of Impedance Parameters Fitted from Fig. 5(a).

Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.