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

Greatly improved photoresponse in the MAPbBr3/Si heterojunction by introducing an ITO layer and optimizing MAPbBr3 layer thickness

Open Access Open Access

Abstract

In this paper, a CH3NH3(MA)PbBr3/Si heterojunction photodetector (PD) is prepared, and a simple method is proposed to improve the performance by introducing an ITO conductive layer and modulating thickness of the MAPbBr3 layer. The results indicate that the MAPbBr3/Si heterojunction PD exhibits an ultra-broadband photoresponse ranging from 405 to 1064 nm, and excellent performances with the responsivity (R) of 0.394 mA/W, detectivity (D) of 0.11×1010 Jones, and response times of ∼2176/∼257 ms. When adding the ITO layer, the R and D are greatly improved to 0.426 A/W and 5.17×1010 Jones, which gets an increment of 1.08×105% and 4.7×103%, respectively. Meanwhile, the response times are reduced to ∼130/∼125 ms, and a good environmental stability is obtained. Moreover, it is found that the photoresponse is strongly dependent on the thickness of the MAPbBr3 layer. By modulating the MAPbBr3 layer thickness from ∼85 to ∼590 nm, the performances are further improved with the best R of ∼0.87 A/W, D of ∼1.92×1011 Jones, and response times of ∼129/∼130 ms achieved in the ∼215 nm-thick PD.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Photodetectors (PDs), which can convert light signals into electrical signals through the interaction of light and matter, have aroused great interest in many application fields [1,2]. In particular, broadband PDs are an important part of photoelectric systems, optical communications, and environmental monitoring as they can be used in light detection of multiple colors due to their high light response in the entire wavelength range of interest [3,4]. On account of these outstanding features and the wide application fields, broadband PDs have increasingly attracted attentions in recent years, and researchers have been always seeking for different materials to design and prepare these kinds of devices and trying various modulation methods to improve the performances. Recently, metal halide organic-inorganic hybrid perovskites have been under the spotlight due to their outstanding optical and electrical properties of high light absorption, tunable bandgap, long diffusion length, large mobility, and low recombination rate of carriers, [5,6] and many different kinds of perovskite PDs, such as perovskite polycrystalline thin-film PD, [7] perovskite single-crystal PD, [8] perovskite quantum dot PD, [9] and perovskite nanowire PD, [10] have been successfully prepared and observed to exhibit outstanding photoresponses. However, the response spectrum mainly lies within the visible range due to the relatively wider bandgaps (>1.5 eV) of the perovskite materials, which is unfavorable to develop broadband PDs. Meanwhile, the performances are still very low and strongly dependent on the external bias considering the usual working mechanism of photoconductive effect [11]. To solve these issues, one of the best methods is to construct suitable heterostructures, as the combination with the other materials can make full potentials of optical and electrical properties of both materials, and the built-in field formed at the interface can also be utilized to accelerate the separation and transport of photo-excited electron-hole pairs [1214]. Therefore, some researchers have been trying best to build different perovskite-based heterojunctions, such as perovskite/MoS2, [15] perovskite/ZnO, [16,17] perovskite/CdS, [18,19] perovskite/C8BTBT, [20] perovskite/Ge, [1] perovskite/Si [21,22]. Among them, the perovskite/Si heterostructure shows some advantages as the Si is still a dominating semiconductor in the photovoltaic and photoelectric market because of its large abundance, high efficiency, and mature technology. However, the research of the perovskite/Si heterojunctions as PDs is still at the initial stage, and the photoresponses of the perovskite/Si heterojunction PDs are very low and urgently needed to explore.

In addition, considering the hygroscopicity of the amine salt, once the perovskites are exposed to the air with moisture and oxygen, it is prone to decomposition, so that the weak environmental stability is a usual drawback for most of perovskite materials [23,24]. Although several different methods have been proposed to improve the stability, such as interface engineering, [25] controllable growth, [26] solvent treatment, [27] and introduction of additive layers, [28,29] these methods are relatively complicated, and the synergistic improvement with the photoresponses is not easy to obtain. However, different from these methods, adding a transparent conductive layer is thought as a feasible method as the conductive layer can not only protect the fragile perovskite layer from contacting with the air but also acts as a carrier collection layer, which can accelerate the transport and collection of the carriers and then improve the photoelectric conversion efficiency [30,31]. Besides, it is suggested that the absorption, as well as the longitudinal separation and transport of the photo-generated carriers, is strongly dependent on the thickness of the perovskite layer, [32,33] thus suitable thickness should be of great importance in the devices. However, the thickness effect is scarcely considered in the perovskite-based heterojunction PDs.

In this work, a perovskite heterojunction PD is prepared with the structure of CH3NH3(MA)PbBr3/Si, and the photoelectric properties and the environmental stability are well studied by introducing an ITO conductive layer and modulating the MAPbBr3 layer thickness. It is found that the MAPbBr3/Si heterojunction has good photoresponses in the spectral range of 405 to 1064 nm, suggesting its great potential in broadband PD. More importantly, after introducing the ITO layer, the performances are greatly enhanced with the responsivity (R) increasing from 0.394×10−3 to 0.426 A/W, the detectivity (D) increasing from 0.11×1010 to 5.17×1010 Jones, and the response times reducing from ∼2176/∼257 to ∼130/∼125 ms. Meanwhile, a good environmental stability is obtained in the ITO/MAPbBr3/Si heterojunction PD as the photocurrent still remains above 85% after preserving in the ambient environment for 16 days. These results can be attributed to the efficient carrier extraction, collection and the external isolation protection of the ITO layer. Moreover, the ITO/MAPbBr3/Si heterojunctions with different MAPbBr3 layer thicknesses ranging from 85 to 590 nm are prepared and investigated. The photoresponse, which improves gradually to an optimum and then decreases subsequently, shows a nonlinear dependence on the thickness. It is demonstrated that the 215 nm-thick MAPbBr3 layer PD exhibits the best photoresponse performances with the R and D as high as ∼0.87 A/W and ∼1.92×1011 Jones, respectively, and the response speed as fast as ∼129/∼130 ms.

2. Experimental methods

2.1 Preparation of the ITO/MAPbBr3/Si heterojunction PD

The preparation process of the ITO/MAPbBr3/Si heterojunction is shown in Fig. 1(a). A 1.5 cm×1.5 cm n-type Si wafer was used as the substrate. Firstly, the Si substrate is cleaned, treated, and hydrophilic treated. Subsequently, the MAPbBr3 with different concentrations of 0.4 mol/L, 0.6 mol/L, 0.8 mol/L, 1.0 mol/L, 1.2 mol/L, and 1.5 mol/L was prepared on it, respectively. This is followed by an annealing process to remove the solvent and improve the quality at 90 °C for 10 min. Finally, a ∼150 nm-thick ITO conductive layer was deposited on the top of the MAPbBr3 layer by radio frequency (RF) magnetron sputtering at room temperature, and two In electrodes were prepared on the ITO and Si substrate with photoactive area of ∼0.785 mm2, respectively. For each thickness, at least six devices were prepared and measured.

 figure: Fig. 1.

Fig. 1. Schematic diagrams of the (a) fabrication process, (b) device structure, and (c) energy band alignment of the ITO/MAPbBr3/Si heterojunction PD under zero bias.

Download Full Size | PDF

2.2 Characterization of the heterojunctions

An X-ray diffractometer (XRD, Bruker D8 Advance) was used to characterize the crystal structure of the MAPbBr3. A field emission scanning electron microscope (SEM, FEI Nova NanoSE M450) was used to determine the top and cross-sectional view morphologies. A fluorescence spectrometer (Horiba LabRAM HR Evolution) was performed to identify the photoluminescence (PL) spectroscopy under a laser excitation source of 325 nm. A spectrophotometer (Hitachi U-4100) was used to record the ultraviolet-visible absorption spectrum.

2.3 Measurements of photoresponse

The current-voltage (I-V) and current-time (I-t) curves were measured with a source meter (Keithley 4200) in dark and light. For the illumination, six different continuous wave (CW) lasers with wavelengths of 405, 450, 532, 671, 808, and 1064 nm were used, and an optical power meter (Coherent FieldMax II) was used to determine the power density.

3. Results and discussion

The schematic diagram of the ITO/MAPbBr3/Si heterojunction is shown in Fig. 1(b). A MAPbBr3 polycrystalline layer is sandwiched between the ITO and Si substrate. Among them, the Si, the MAPbBr3, and the ITO are acted as the n-type layer, the intrinsic(/p)-type layer, and the conductive layer, respectively. The bandgap of the MAPbBr3 is ∼2.3 eV, and the valence band maximum (VBM) and the conduction band minimum (CBM) are located at −5.7 and −3.4 eV, respectively [34]. Based on the previous results, the band diagram and working mechanism of the ITO/MAPbBr3/Si heterojunction can be illustrated in Fig. 1(c). Therefore, when under the illumination of a suitable light, the photon energy is mainly absorbed by the MAPbBr3 or Si layer, then the photo-excited electron-hole pairs are separated by the interface built-in field of the MAPbBr3 and Si. Obviously, the ITO is very beneficial to the extraction and collection of the carriers due to the suitable band position and good conductive property, demonstrating the potential photoresponse improvement of this heterostructure PD.

Figure 2(a) gives the top-view SEM morphology of the MAPbBr3 layer. The perovskite layer is uniformly prepared, and the surface is generally composed of compact microcrystalline particles with an average size scale of hundreds of nanometers. To better understand the crystal quality and structure of the MAPbBr3 layer, XRD is measured, as shown in Fig. 2(b). Three main diffraction peaks are observed at 15.07°, 30.25°, and 46.05°, which can be assigned to the diffraction patterns of (001), (002), and (003), respectively, indicating the typical cubic structure of the MAPbBr3 [35]. Moreover, the very strong diffraction intensities of the (001) and (002) peaks, as well as the observation of the diffraction signal from the (003) crystalline plane, demonstrate the preferential growth of the film along the <001 > direction. To investigate the optical properties of the MAPbBr3 thin film, both the steady-state PL and absorption spectra are characterized, as shown in Fig. 2(c) and 2(d). The only luminescence peak observed at ∼545 nm (∼2.28 eV) can be thought as the recombination of the electrons at the CBM and holes at the VBM of the MAPbBr3 [34,35]. While the absorption results indicate that an additional steep absorption edge appears at ∼540 nm for the MAPbBr3/Si heterojunction as compared with the absorption spectra of the Si substrate, which directly confirms the synergistic effect of the MAPbBr3 and the Si.

 figure: Fig. 2.

Fig. 2. (a) Top view SEM image, and (b) XRD result of the MAPbBr3 film. (c) PL result of the MAPbBr3 film. (d) Absorption results of the Si and MAPbBr3/Si heterojunction.

Download Full Size | PDF

The I-V characteristics of the MAPbBr3/Si heterojunction with a MAPbBr3 layer thickness of 85 nm under a typical 532 nm, and several other lasers (405, 450, 671, 808, and 1064 nm) illuminations of different power densities are shown in Fig. 3(a) and S1, respectively. All the curves show typical rectifying behaviors in dark, indicating the formation of the heterostructure between the MAPbBr3 and the Si with good quality. Besides, for all laser illuminations, obvious current responses can be observed and the output currents increase gradually with the incident power density increasing from 6.36 to 127.38 mW/cm2 and reverse bias adding from 0 to −1 V. The photoresponse in the whole wavelength range demonstrates the broadband spectral property of this heterojunction, and the large power density range suggests that this heterojunction PD can detect incident light power over a wide range. Then, the photocurrents (Iph) of different power densities at −1 V are extracted, as shown in Fig. 3(b) and S2 (solid dots). Obvious nonlinear relationships can be observed between the Iph and the power density, which can be theoretically fitted with a power-law, [36,37]

$${I_{\textrm{ph}}} = \alpha {P^\beta }$$
where Iph is calculated by subtracting the current in dark from the current in light (Iph = Ilight-Idark, Ilight represents the output current in light, and Idark represents the output current in dark), α is a proportional coefficient, P is the power density, and β is an important exponent parameter that is used to determine the photoresponse behavior of the PD. By fitting the Iph results with the Eq. (1), the best fitted parameters of β are obtained to be 0.8883, 0.5264, 0.3331, 0.2192, 0.2062, and 0.1804 for the 405, 450, 532, 671, 808, and 1064 nm, respectively. The serious departure of the β from the ideal factor of 1, especially in the long wavelength range, demonstrates the complex process of the carriers’ generation, recombination, separation, and trapping in the PDs [36]. Then, the R, which is a key parameter of one PD and defined by the Iph response generated at a unit laser power, is determined for the MAPbBr3/Si heterojunction PD in the whole response range based on the relation of [36,37]
$$R = \frac{{{I_{\textrm{ph}}}}}{{PS}}$$
where S represents the laser irradiation area of the PD. It is suggested that the R reflects the photoelectric conversion capability of the PD, and is generally strongly dependent on both the incident laser wavelength and power density as they may result in a sharp change in the R when the linear dynamic response range is beyond the PDs [14]. As shown in Fig. 3(c), the R decreases quickly with an increase in the power density for all laser wavelengths, and the maximum R of 0.394 mA/W can be observed for the 532 nm laser at 6.36 mW/cm2. Besides, another critical paremeter of one PD is the D, which shows the detection capability and can be extracted by [36,37]
$$D = \frac{R}{{\sqrt {2q{J_d}} }}$$
where Jd is the current density in dark and q is the quantity of a unit electric charge. Dominated by the R, the D also decreases gradually with an increase in the power density, and the corresponding maximum value of 1.10×109 Jones is also obtained at 6.36 mW/cm2 of the 532 nm laser illumination, as shown in Fig. 3(d). As is well known that the decaying of the R and D as a function of the laser power density is a common phenomenon in most of PDs, which can be mainly attributed to the enhanced scattering and recombination rate of the photo-excited carriers [14,15,19,22,37].

 figure: Fig. 3.

Fig. 3. (a) I-V results of the MAPbBr3/Si heterojuncion PD under a 532 nm laser illumination of different power densities, with inset the I-V curve in dark. (b) The corresponding photocurrent as a function of illumination power density at −1 V. Extracted (c) R and (d) D results under the illumination of different lasers.

Download Full Size | PDF

Considering the potential improvement of the photoelectric properties of the MAPbBr3/Si heterojunction by introducing an ITO conductive layer, the I-V characteristics are also well measured for the ITO/MAPbBr3/Si heterojunction PD under the illumination of a 532 nm laser, as shown in Fig. 4(a). Typical rectifying behavior and photoresponses can still be obtained in dark and light conditions, respectively. However, the output current is largely improved as compared with that without the ITO layer. Figure 4(b) plots the extracted Iph results of the different power densities at −1 V. It can be seen that the Iph is ∼380-∼1010 times larger than that of the MAPbBr3/Si heterojunction PD. Besides, based on the Eq. (1), the best-fitted β, which is gotten to be 0.5478, is also remarkably improved, indicating the significant improvement of the photoelectric conversion performance by adding the ITO layer. Moreover, the I-V characteristics of the ITO/MAPbBr3/Si heterojunction PD under the illumination of other five lasers (405, 450, 671, 808, and 1064 nm) are also measured and analyzed, as shown in Fig. S3. The photocurrents of different wavelengths, as well as the corresponding β, are all greatly enhanced with different degrees, as shown in Fig. S4. Based on the Iph results, the R and D parameters are deduced and given in Fig. 4(c) and 4(d). Both the R and the D still show a decreasing tendency with increasing the laser power density, but they are improved by two or three orders of magnitude as compared with these without the ITO layer in the whole power density range. In addition, it can be seen that the photoresponses in both the MAPbBr3/Si and ITO/MAPbBr3/Si heterojunctions are strongly dependent on the wavelength. However, a significant change in the wavelength dependency happens before and after introducing the ITO layer. In order to show it more clearly, the maximum R and D parameters of different wavelengths at 6.36 mW/cm2 are summarized in Fig. 5(a) and 5(b). For the MAPbBr3/Si heterojunction PD, the optimal photoresponse is observed at the 532 nm (the optical absorption edge of the MAPbBr3), which can be ascribed to the strong absorption and high photoelectric conversion efficiency of the MAPbBr3. While, after adding the ITO layer, the optimal response wavelength changes to 671 nm for the ITO/MAPbBr3/Si heterojunction PD with the maximum R and D of 0.426 A/W and 5.17×1010 Jones, which gets an enhancement of 1.08×105% and 4.7×103%, respectively. In order to well understand the red shift of the optimal response wavelength, the absorbance spectra of the ITO/MAPbBr3/Si heterojunction are measured, as shown in Fig. S5. Different from the absorbance results in Fig. 2(d), there is a very strong absorption peak appearing between 600 and 900 nm, meaning that a relatively larger number of carriers would be excited in this range when under illumination, which is extremely beneficial to the output photocurrent. That may be why the optimal photoresponse wavelength is gotten at the 671 nm. From the above discussion, it is indicated that the photoresponses can be greatly enhanced by introducing the ITO layer, which can be attributed to its good capability of carrier extraction and collection due to the suitable band structure and high conductive property.

 figure: Fig. 4.

Fig. 4. (a) I-V results of the ITO/MAPbBr3/Si heterojunction PD under a 532 nm laser illumination of different power densities, with inset the I-V curve in dark. (b) The corresponding photocurrent as a function of illumination power density at −1 V. Extracted (c) R and (d) D results under the illumination of different lasers.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Extracted maximum R and D results as a function of the laser wavelength for the (a) MAPbBr3/Si heterojunction PD and (b) ITO/MAPbBr3/Si heterojunction PD. The corresponding rising and falling edges of the (c) MAPbBr3/Si heterojunction PD and (d) ITO/MAPbBr3/Si heterojunction PD. Transient photocurrent responses of (e) MAPbBr3/Si heterojunction PD and (f) ITO/MAPbBr3/Si heterojunction PD after maintaining in the air for several days.

Download Full Size | PDF

Then, the response speed, which is a key factor to reflect the potential to respond to a fast-changing optical signal, is also identified. The I-t results of the two devices with and without the ITO layer were measured under the illumination of a mechanically modulated continuous wave laser by a chopper with the transient laser on or off time of 1 s and the pulse interval of 10 s at 6.36 mW/cm2. The rise time (tr) and the fall time (tf) are determined by the duration of the photocurrent increasing or decreasing from 10%(/90%) to 90%(/10%) of the maximum value.37 As shown in Fig. 5(c), the tr and the tf are deduced to be ∼2176 and ∼257 ms for the MAPbBr3/Si heterojunction PD, respectively. While, after introducing the ITO layer, the response times are greatly reduced to ∼130 and ∼125 ms correspondingly for the ITO/MAPbBr3/Si heterojunction PD (Fig. 5(d)), further confirming the important carrier extraction capability of the ITO layer. Subsequently, the environmental stability are conducted for the MAPbBr3/Si and ITO/MAPbBr3/Si heterojunctions with the PDs placing in an ambient environment of humidity of about 40% and a temperature of ∼25 °C without any encapsulation and protection, with the transient photocurrent results of different days shown in Fig. 5(e) and 5(f). It can be seen that the photocurrent of the MAPbBr3/Si heterojunction PD decays quickly with the placing time and nearly tends to zero after about two days. While the ITO/MAPbBr3/Si heterojunctions shows high environmental stability as the photocurrent nearly keeps constant in the first eight days, and there is only a ∼15% decreasing after exposure to air for 16 days. This result shows that the ITO layer can also bring about an improvement of the environmental stability in the MAPbBr3/Si heterojunction due to the protection of the MAPbBr3 layer against the O2 and H2O.

As previously reported, the thickness of the photoactive layer is thought to have important effects on the absorption, carriers’ generation, separation and diffusion, [31,32] so that optimizing MAPbBr3 layer thickness should be very crucial to the photoresponse of the ITO/MAPbBr3/Si heterojunction PD and can make full potential of the heterostructure. Therefore, to determine the thickness effect of the MAPbBr3 layer on the performance, a series of ITO/MAPbBr3/Si heterojunction PDs are prepared with the MAPbBr3 layer thickness of ∼85, ∼160, ∼215, ∼380, ∼440, and ∼590 nm, respectively, with the cross-sectional SEM morphologies shown in Fig. S6(a-f). In order to well distinguish them, these PDs are denoted as #1, #2, #3, #4, #5, and #6, correspondingly. With an increase in the MAPbBr3 layer thickness, the XRD diffraction intensity, the relative absorbance, and the PL intensity get an enhancement gradually, but there is nearly no any changes in their shapes and positions or edges, as shown in Fig. S6(g-i).

Figure 6(a) gives the I-V curves of the ITO/MAPbBr3/Si heterojunction PDs with different MAPbBr3 layer thicknesses under the illumination of a 671 nm laser at 6.36 mW/cm2. The Iph at −1 V, which increases from 23.01 µA to a maximum of 43.89 µA with the layer thickness changing from ∼85 to ∼215 nm, then decreases quickly to 3.51 µA when the layer thickness reaching ∼590 nm, exhibits a nonlinear dependence on the thickness of the MAPbBr3 layer, as shown in Fig. 6(b). Not only that, the photoresponses of these PDs are also measured under the illumination of other larger power densities (as shown in Fig. S7), and nearly the same thickness dependence of the photocurrent (Iph) can still be observed (as shown in Fig. S8), indicating the intrinsic property of the ITO/MAPbBr3/Si heterojunction. Figure 6(c) and 6(d) gives the R and D results of the devices as a function of the power density. For all PDs, both the R and the D show a gradual decreasing tendency with the power density, and are strongly dependent on the MAPbBr3 layer thickness as similar as the photocurrent, with the best results of each thickness summarized in Fig. 6(e) and 6(f). Obviously, the 215 nm-thick PD exhibits the optimal photoresponse with the R as high as ∼0.87 A/W and D as high as ∼1.92×1011 Jones, which gets an increment of 204% and 371%, respectively, as compared with that of the 85 nm-thick PD. Notably, here the optimal R value of the ITO/MAPbBr3/Si heterojunction is much larger than most of previous results based on the perovskite films, as summarized in Table 1, indicating its greatly potential application in photoelectric devices. At last, the response speeds are also well investigated in these PDs, with the typical I-t result of the 215 nm-thick PD under the illumination of 6.36 mW/cm2 shown in Fig. 7(a). No any attenuation or oscillation can be observed in the multiple cycles, suggesting the outstanding repeatability and stability of the photoresponse in this heterojunction. By magnifying the laser on and off stages in one period, the tr and the tf of ∼129/∼130 ms are deduced, as shown in Fig. 7(b) and 7(c). Besides, the transient I-t curves of the other thickness PDs are also well measured, with the extracted response time results summarized in Fig. 7(d). It seems that the response time increases slightly with increasing the MAPbBr3 layer thickness mainly due to the increased longitudinal separation and transport distance, but these response speeds are all much larger than that without the ITO layer, further demonstrating the important effect of the ITO layer on the photoresponse performances.

 figure: Fig. 6.

Fig. 6. (a) I-V results of the ITO/MAPbBr3/Si heterojunction PDs with different MAPbBr3 layer thicknesses under the illumination of a 671 nm laser at 6.36 mW/cm2. (b) The corresponding extracted photocurrent results at −1 V. The calculated (c) R, and (d) D results as a function of power density for different MAPbBr3 layer thickness devices. The extracted maximum (e) R and (f) D results of the ITO/MAPbBr3/Si heterojunction PD as a function of the MAPbBr3 layer thickness.

Download Full Size | PDF

 figure: Fig. 7.

Fig. 7. (a) Time-dependent photoresponse of the ITO/MAPbBr3(215 nm)/Si heterojunction PD at −1 V. The corresponding (b) rising and (c) falling edges of one response cycle for estimating the response time. (d) Extracted response times of the different MAPbBr3 layer thickness PDs.

Download Full Size | PDF

Tables Icon

Table 1. Performance comparisons of perovskite PDs with different structures

4. Conclusions

In conclusion, a broadband, high-performance and stable MAPbBr3/Si heterojunction PD is successfully prepared by introducing an ITO conductive layer and optimizing the thickness of the MAPbBr3 layer. It is found that the PD can exhibit a broadband photoresponse ranging from 405 to 1064 nm. After introducing the ITO layer, the R is increased from 0.394×10−3 to 0.426 A/W with an increment of 1.08×105%, the D is improved from 0.11×1010 to 5.17×1010 Jones with an enhancement of 4.7×103%, the response times are reduced from ∼2176/∼257 to ∼130/∼125 ms, and meanwhile the environmental stability is greatly improved. These results can be attributed to the efficient carrier extraction, collection and the external isolation protection of the ITO layer. In addition, the ITO/MAPbBr3/Si heterojunctions with different MAPbBr3 layer thicknesses ranging from 85 to 590 nm are prepared, and the photoresponses are found to be strongly dependent on the layer thickness. By modulating the thickness of the MAPbBr3 layer, the performances are further improved with the best R and D as high as ∼0.87 A/W and ∼1.92×1011 Jones, respectively, and the response speed as fast as ∼129/∼130 ms obtained in the 215 nm-thick MAPbBr3 layer PD at −1 V. This result not only gives a simple and effective method to enhance the performance of the MAPbBr3/Si heterojunction but also brings an insight on designing broadband, high-photoresponse, fast speed and low-cost perovskite-based optoelectronic devices.

Funding

Science and Technology Plan Project of Hebei Province (216Z1703G); Nature Science Foundation of Hebei Province (F2018201198, F2019201047); National Natural Science Foundation of China (11704094, 62175058, U20A20166).

Acknowledgments

This work was supported by the National Nature Science Foundation of China (Grant Nos. 62175058, U20A20166, and 11704094), the Nature Science Foundation of Hebei Province (Grant Nos. F2019201047, and F2018201198), and the Science and Technology Plan Project of Hebei Province (Grant No. 216Z1703G).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. W. Hu, H. Cong, W. Huang, L. Chen, A. Pan, and C. Xue, “Germanium/perovskite Heterostructure for High-Performance and Broadband Photodetector From Visible to Infrared Telecommunication Band,” Light: Sci. Appl. 8(1), 106 (2019). [CrossRef]  

2. L. T. Dou, Y. Yang, J. B. You, Z. R. Hong, W. H. Chang, G. Li, and Y. Yang, “Solution-Processed Hybrid Perovskite Photodetectors with High Detectivity,” Nat. Commun. 5(1), 5404 (2014). [CrossRef]  

3. X. Hu, X. Zhang, L. Liang, J. Bao, S. Li, W. Yang, and Y. Xie, “High-Performance Flexible Broadband Photodetector Based on Organolead Halide Perovskite,” Adv. Funct. Mater. 24(46), 7373–7380 (2014). [CrossRef]  

4. H. Sun, W. Tian, F. Cao, J. Xiong, and L. Li, “Ultrahigh-Performance Self-Powered Flexible Double-Twisted Fibrous Broadband Perovskite Photodetector,” Adv. Mater. 30(21), 1706986 (2018). [CrossRef]  

5. Y. Zhao and K. Zhu, “Organic-Inorganic Hybrid Lead Halide Perovskites for Optoelectronic and Electronic Applications,” Chem. Soc. Rev. 45(3), 655–689 (2016). [CrossRef]  

6. Y. Zhao, C. Li, and L. Shen, “Recent Advances on Organic-Inorganic Hybrid Perovskite Photodetectors with Fast Response,” InfoMat 1(2), inf2.12010 (2019). [CrossRef]  

7. S. P. Senanayak, B. Yang, T. H. Thomas, N. Giesbrecht, W. Huang, E. Gann, B. Nair, K. Goedel, S. Guha, X. Moya, C. R. Mcneill, P. Docampo, A. Sadhanala, R. H. Friend, and H. Sirringhaus, “Understanding Charge Transport in Lead Iodide Perovskite Thin-Film Field-Effect Transistors,” Sci. Adv. 3(1), e1601935 (2017). [CrossRef]  

8. Y. Liu, Y. Zhang, K. Zhao, Z. Yang, J. Feng, X. Zhang, K. Wang, L. Meng, H. Ye, M. Liu, and S. Liu, “A 1300 mm2 Ultrahigh-Performance Digital Imaging Assembly using High-Quality Perovskite Single Crystals,” Adv. Mater. 30(29), 1707314 (2018). [CrossRef]  

9. Y. Chen, Y. Chu, X. Wu, W. OuYang, and J. Huang, “High-Performance Inorganic Perovskite Quantum Dot-Organic Semiconductor Hybrid Phototransistors,” Adv. Mater. 29(44), 1704062 (2017). [CrossRef]  

10. L. Gao, K. Zeng, J. Guo, C. Ge, J. Du, Y. Zhao, C. Chen, H. Deng, Y. He, H. Song, G. Niu, and J. Tang, “Passivated Single-Crystalline CH3NH3PbI3 Nanowire Photodetector with High Detectivity and Polarization Sensitivity,” Nano Lett. 16(12), 7446–7454 (2016). [CrossRef]  

11. W. Tian, H. Zhou, and L. Li, “Hybrid Organic-Inorganic Perovskite Photodetectors,” Small 13(41), 1702107 (2017). [CrossRef]  

12. L. Shen, Y. Fang, D. Wang, Y. Bai, Y. Deng, M. Wang, Y. Lu, and J. Huang, “A Self-Powered, Sub-Nanosecond-Response Solution-Processed Hybrid Perovskite Photodetector for Time-Resolved Photoluminescence-Lifetime Detection,” Adv. Mater. 28(48), 10794–10800 (2016). [CrossRef]  

13. M. Cao, J. Tian, Z. Cai, L. Peng, L. Yang, and D. Wei, “Perovskite Heterojunction Based on CH3NH3PbBr3 Single Crystal for High-Sensitive Self-Powered Photodetector,” Appl. Phys. Lett. 109(23), 233303 (2016). [CrossRef]  

14. S. Qiao, Y. Liu, J. Liu, G. Fu, and S. Wang, “High-Responsivity, Fast, and Self-Powered Narrowband Perovskite Heterojunction Photodetectors with a Tunable Response Range in the Visible and Near-Infrared Region,” ACS Appl. Mater. Interfaces 13(29), 34625–34636 (2021). [CrossRef]  

15. F. Bai, J. Qi, F. Li, Y. Fang, W. Han, H. Wu, and Y. Zhang, “A High-Performance Self-Powered Photodetector Based on Monolayer MoS2/Perovskite Heterostructures,” Adv. Mater. Interfaces 5(6), 1701275 (2018). [CrossRef]  

16. J. Yu, X. Chen, Y. Wang, H. Zhou, M. Xue, Y. Xu, Z. Li, C. Ye, J. Zhang, P. A. Aken, P. D. Lund, and H. Wang, “A High-Performance Self-powered Broadband Photodetector Based on a Ch3NH3PbI3 perovskite/ZnO nanorod array heterostructure,” J. Mater. Chem. C 4(30), 7302–7308 (2016). [CrossRef]  

17. H. Liu, X. Zhang, L. Zhang, Z. Yin, D. Wang, J. Meng, J. Jiang, Y. Wang, and J. You, “A High-Performance Photodetector Based on Inorganic Perovskite-ZnO Heterostructure,” J. Mater. Chem. C 5(25), 6115–6122 (2017). [CrossRef]  

18. F. Cao, L. Meng, M. Wang, W. Tian, and L. Li, “Gradient Energy Band Driven High-Performance Self-Powered Perovskite/CdS Photodetector,” Adv. Mater. 31(12), 1806725 (2019). [CrossRef]  

19. Z. Li, H. Li, K. Jiang, D. Ding, J. Li, C. Ma, S. Jiang, Y. Wang, T. D. Anthopoulos, and Y. Shi, “Self-Powered Perovskite/CdS Heterostructure Photodetectors,” ACS Appl. Mater. Interfaces 11(43), 40204–40213 (2019). [CrossRef]  

20. S. Tong, J. Sun, C. Wang, Y. Huang, C. Zhang, J. Shen, H. Xie, D. Niu, S. Xiao, Y. Yuan, J. He, J. Yang, and Y. Gao, “High-Performance Broadband Perovskite Photodetectors Based on CH3NH3PbI3/C8BTBT Heterojunction,” Adv. Electron. Mater. 3(7), 1700058 (2017). [CrossRef]  

21. X. Geng, F. Wang, H. Tian, Q. Feng, H. Zhang, R. Liang, Y. Shen, Z. Ju, G. Y. Gou, N. Deng, Y. T. Li, J. Ren, D. Xie, Y. Yang, and T. L. Ren, “Ultrafast Photodetector by Integrating Perovskite Directly on Silicon Wafer,” ACS Nano 14(3), 2860–2868 (2020). [CrossRef]  

22. J. Liu, Y. Gao, G. Wu, X. Tong, C. Xie, L. Luo, L. Liang, and Y. Wu, “Silicon/Perovskite Core-Shell Heterojunctions with Light-Trapping Effect for Sensitive Self-Driven Near-Infrared Photodetectors,” ACS Appl. Mater. Interfaces 10(33), 27850–27857 (2018). [CrossRef]  

23. X. Dong, X. Fang, M. Lv, B. Lin, S. Zhang, J. Ding, and N. Yuan, “Improvement of the Humidity Stability of Organic-Inorganic Perovskite Solar Cells Using Ultrathin Al2O3 Layer Prepared by Atomic Layer Deposition,” J. Mater. Chem. A 3(10), 5360–5367 (2015). [CrossRef]  

24. Z. Wang, Z. Shi, T. Li, Y. Chen, and W. Huang, “Stability of Perovskite Solar Cells: A Prospective on the Substitution of the A Cation and X Anion,” Angew. Chem. 56(5), 1190–1212 (2017). [CrossRef]  

25. G. Grancini, C. Roldan-Carmona, I. Zimmermann, E. Mosconi, X. Lee, D. Martineau, S. Narbey, F. Oswald, F. Angelis, M. Graetzel, and M. K. Nazeeruddin, “One-Year Stable Perovskite Solar Cells by 2D/3D Interface Engineering,” Nat. Commun. 8(1), 15684 (2017). [CrossRef]  

26. X. Gong, M. Li, X. Shi, H. Ma, Z. Wang, and L. Liao, “Controllable Perovskite Crystallization by Water Additive for High-Performance Solar Cells,” Adv. Funct. Mater. 25(42), 6671–6678 (2015). [CrossRef]  

27. M. Li, X. Yan, Z. Kang, X. Liao, Y. Li, X. Zheng, P. Lin, J. Meng, and Y. Zhang, “Enhanced Efficiency and Stability of Perovskite Solar Cells via Anti-Solvent Treatment in Two-Step Deposition Method,” ACS Appl. Mater. Interfaces 9(8), 7224–7231 (2017). [CrossRef]  

28. F. Zhang and K. Zhu, “Additive Engineering for Efficient and Stable Perovskite Solar Cells,” Adv. Energy Mater. 10(13), 1902579 (2020). [CrossRef]  

29. Y. Liu, S. Qiao, H. Sun, J. Liu, L. Guo, Z. Yang, G. Fu, and S. Wang, “Large-Area Heterojunction Photodetectors Based on Nanometer-Thick CH3NH3PbI3 Films Modified with Poly(methylmethacrylate) Nanofilms,” ACS Appl. Nano Mater. 4(2), 1682–1691 (2021). [CrossRef]  

30. X. Zhao, H. Shen, Y. Zhang, X. Li, X. Zhao, M. Tai, J. Li, J. Li, X. Li, and H. Lin, “Aluminum-Doped Zinc Oxide as Highly Stable Electron Collection Layer for Perovskite Solar Cells,” ACS Appl. Mater. Interfaces 8(12), 7826–7833 (2016). [CrossRef]  

31. H. Yang, H. Kwon, S. Ma, K. Kim, S. C. Yun, G. Jang, J. Park, H. Lee, S. Goh, and J. Moon, “Energy Level-Graded Al-Doped ZnO Protection Layers for Copper Nanowire-Based Window Electrodes for Efficient Flexible Perovskite Solar Cells,” ACS Appl. Mater. Interfaces 12(12), 13824–13835 (2020). [CrossRef]  

32. M. Rai, L. H. Wong, and L. Etgar, “Effect of Perovskite Thickness on Electroluminescence and Solar Cell Conversion Efficiency,” J. Phys. Chem. Lett. 11(19), 8189–8194 (2020). [CrossRef]  

33. T. Du, W. Xu, S. Xu, S. R. Ratnasingham, C. T. Lin, J. Kim, J. Briscoe, M. A. McLachlan, and J. R. Durrant, “Light-Intensity and Thickness Dependent Efficiency of Planar Perovskite Solar Cells: Charge Recombination versus Extraction,” J. Mater. Chem. C 8(36), 12648–12655 (2020). [CrossRef]  

34. Y. Liang, Y. Wang, C. Mu, S. Wang, X. Wang, D. Xu, and L. Sun, “Achieving High Open-Circuit Voltages up to 1.57 V in Hole-Transport-Material-Free MAPbBr3 Solar Cells with Carbon Electrodes,” Adv. Energy Mater. 8(4), 1701159 (2018). [CrossRef]  

35. S. Ryu, J. H. Noh, N. J. Jeon, Y. C. Kim, W. S. Yang, J. Seo, and S. Seok, “Voltage Output of Efficient Perovskite Solar Cells with high Open-Circuit Voltage and Fill Factor,” Energy Environ. Sci. 7(8), 2614–2618 (2014). [CrossRef]  

36. S. Qiao, J. Liu, X. Niu, B. Liang, G. Fu, Z. Li, S. Wang, K. Ren, and C. Pan, “Piezophototronic Effect Enhanced Photoresponse of the Flexible Cu(In,Ga)Se2 (CIGS) Heterojunction Photodetectors,” Adv. Funct. Mater. 28(19), 1707311 (2018). [CrossRef]  

37. S. Qiao, R. Cong, J. Liu, B. Liang, G. Fu, W. Yu, K. Ren, S. Wang, and C. Pan, “A Vertically Layered MoS2/Si Heterojunction for an Ultrahigh and Ultrafast Photoresponse Photodetector,” J. Mater. Chem. C 6(13), 3233–3239 (2018). [CrossRef]  

38. H. Zhou, J. Mei, M. Xue, Z. Song, and H. Wang, “High-Stability, Self-Powered Perovskite Photodetector Based on a CH3NH3PbI3/GaN Heterojunction with C60 as an Electron Transport Layer,” J. Phys. Chem. C 121(39), 21541–21545 (2017). [CrossRef]  

39. C. Bao, W. Zhu, J. Yang, F. Li, S. Gu, Y. Wang, T. Yu, J. Zhu, Y. Zhou, and Z. Zou, “Highly Flexible Self-Powered Organolead Trihalide Perovskite Photodetectors with Gold Nanowire Networks as Transparent Electrodes,” ACS Appl. Mater. Interfaces 8(36), 23868–23875 (2016). [CrossRef]  

40. B. R. Sutherland, A. K. Johnston, A. H. Ip, J. Xu, V. Adinolfi, P. Kanjanaboos, and E. H. Sargent, “Sensitive, Fast, and Stable Perovskite Photodetectors Exploiting Interface Engineering,” ACS Photonics 2(8), 1117–1123 (2015). [CrossRef]  

41. Y. Wang, T. Zhang, P. Zhang, D. Liu, L. Ji, H. Chen, Z. D. Chen, J. Wu, and S. Li, “Solution Processed PCBM-CH3NH3PbI3 Heterojunction Photodetectors with Enhanced Performance and Stability,” Org. Electron. 57, 263–268 (2018). [CrossRef]  

42. B. Xin, Y. Pak, S. Mitra, D. Almalawi, N. Alwadai, Y. H. Zhang, and I. S. Roqan, “Self-Patterned CsPbBr3 Nanocrystals for High-Performance Optoelectronics,” ACS Appl. Mater. Interfaces 11(5), 5223–5231 (2019). [CrossRef]  

43. I. M. Asuo, D. Banerjee, A. Pignolet, R. Nechache, and S. G. Cloutier, “Perovskite/Silicon-Nanowire-Based Hybrid Heterojunctions for Fast and Broadband Photodetectors,” Phys. Status Solidi RRL 15(4), 2000537 (2021). [CrossRef]  

44. I. M. Asuo, D. Gedamu, I. Ka, L. F. Gerlein, F. X. Fortier, A. Pignolet, S. G. Cloutier, and R. Nechache, “High-Performance Pseudo-Halide Perovskite Nanowire Networks for Stable and Fast-Response Photodetector,” Nano Energy 51, 324–332 (2018). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental Document

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (7)

Fig. 1.
Fig. 1. Schematic diagrams of the (a) fabrication process, (b) device structure, and (c) energy band alignment of the ITO/MAPbBr3/Si heterojunction PD under zero bias.
Fig. 2.
Fig. 2. (a) Top view SEM image, and (b) XRD result of the MAPbBr3 film. (c) PL result of the MAPbBr3 film. (d) Absorption results of the Si and MAPbBr3/Si heterojunction.
Fig. 3.
Fig. 3. (a) I-V results of the MAPbBr3/Si heterojuncion PD under a 532 nm laser illumination of different power densities, with inset the I-V curve in dark. (b) The corresponding photocurrent as a function of illumination power density at −1 V. Extracted (c) R and (d) D results under the illumination of different lasers.
Fig. 4.
Fig. 4. (a) I-V results of the ITO/MAPbBr3/Si heterojunction PD under a 532 nm laser illumination of different power densities, with inset the I-V curve in dark. (b) The corresponding photocurrent as a function of illumination power density at −1 V. Extracted (c) R and (d) D results under the illumination of different lasers.
Fig. 5.
Fig. 5. Extracted maximum R and D results as a function of the laser wavelength for the (a) MAPbBr3/Si heterojunction PD and (b) ITO/MAPbBr3/Si heterojunction PD. The corresponding rising and falling edges of the (c) MAPbBr3/Si heterojunction PD and (d) ITO/MAPbBr3/Si heterojunction PD. Transient photocurrent responses of (e) MAPbBr3/Si heterojunction PD and (f) ITO/MAPbBr3/Si heterojunction PD after maintaining in the air for several days.
Fig. 6.
Fig. 6. (a) I-V results of the ITO/MAPbBr3/Si heterojunction PDs with different MAPbBr3 layer thicknesses under the illumination of a 671 nm laser at 6.36 mW/cm2. (b) The corresponding extracted photocurrent results at −1 V. The calculated (c) R, and (d) D results as a function of power density for different MAPbBr3 layer thickness devices. The extracted maximum (e) R and (f) D results of the ITO/MAPbBr3/Si heterojunction PD as a function of the MAPbBr3 layer thickness.
Fig. 7.
Fig. 7. (a) Time-dependent photoresponse of the ITO/MAPbBr3(215 nm)/Si heterojunction PD at −1 V. The corresponding (b) rising and (c) falling edges of one response cycle for estimating the response time. (d) Extracted response times of the different MAPbBr3 layer thickness PDs.

Tables (1)

Tables Icon

Table 1. Performance comparisons of perovskite PDs with different structures

Equations (3)

Equations on this page are rendered with MathJax. Learn more.

I ph = α P β
R = I ph P S
D = R 2 q J d
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.