Realizing photomultiplication-type organic photodetectors based on C60-doped bulk heterojunction structure at low bias
Gong Wei, An Tao, Liu Xinying, Lu Gang
Department of Electronic Engineering, Xi’an University of Technology, Xi’an 710048, China

 

† Corresponding author. E-mail: antao@xaut.edu.cn

Project supported by the National Natural Science Foundation of China (Grant No. 61106043), the Natural Science Foundation of Shaanxi Province, China (Grant No. 2015JM6267), and Program of Xi’an University of Technology, China (Grant No. 103-4515013).

Abstract

Photomultiplication (PM) structure has been widely employed to improve the optoelectronic performance of organic photodetectors (OPDs). However, most PM-type OPDs require a high negative operating voltage or complex fabrication. For obtaining high-efficiency OPDs with low voltage and easy process, here the bulk heterojunction (BHJ) structure of high exciton dissociation efficiency combined with the method of trap-assisted PM are applied to the OPDs. In this paper, we investigate the operating mechanism of OPDs based on poly(3-hexylthiophene) (P3HT): (phenyl-C61-butyric-acid-methyl-ester) (PC61BM), and poly-{[4,8-bis[(2-ethylhexyl)oxy]-benzo[1,2-b:4,5-b]dithiophene-2,6-diyl]-alt-[3-fluore-2-(octyloxy)carbonyl-thieno[3,4-b]thiophene-4,6-diyl]} (PBDT-TT-F):PC61BM doped with C60 as active layer. Furthermore, the influence of C60 concentration on the optoelectronic performances is also discussed. With 1.6 wt.% C60 added, the P3HT:PC61BM:C60 OPD exhibits a 327.5% external quantum efficiency, a 1.21 A·W−1 responsivity, and a 4.22 × 1012 Jones normalized detectivity at −1 V under 460 nm (0.21 mW·cm−2) illumination. The experimental results show that the effective electron traps can be formed by doping a small weight of C60 into BHJ active layer. Thus the PM-type OPDs can be realized, which benefits from the cathode hole tunneling injection assisted by the trapped electrons in C60 near the Al side. The efficiency of PM is related to the C60 concentration. The present study provides theoretical basis and method for the design of highly sensitive OPDs with low operating voltage and facile fabrication.

1. Introduction

Organic photodetectors (OPDs) in the image sensing, communications, ultraviolet (UV)–visible–near infrared (NIR) region, and chemical/biological sensing show a wide range of commercial and scientific applications.[15] Conventionally, most of the OPDs are photodiodes with photoresponse originating from the collection of photogenerated charges.[6,7] Due to the limitation of photon harvesting efficiency, exciton dissociation efficiency, and photogenerated charge carrier collection efficiency, the external quantum efficiency (EQE) of photodiode-type OPDs is lower than 100%.[8,9] In spite of the photodiode-type OPDs possessing relatively low working voltage and noise current, the responsivity (R) is not too high due to the EQE values of <100%. In order to achieve highly sensitive optical detection, environmental noise should be generally reduced by complex equipment.[10,11] Therefore, the OPDs with facile fabrication, high responsivity R, and normalized detectivity (D*) are in urgent need.[1215] The critical steps for improving normalized detectivity of OPDs are to enhance responsivity.[16] In recent years, photomultiplication (PM) structure has been widely applied to the OPDs so as to further improve the optoelectronic performance. The efforts on trap-assisted PM have always been attempted in OPDs and become a hot topic recently.[1219]

Wang et al. proposed a solution processed PM-type OPD based on P3HT:PC61BM blend film with PEIE modified ITO electrode, which exhibits a R of 14.25 A·W−1 under 550 nm (3.09 μW/cm2) at −1.0 V bias.[16] However, the D* is not improved correspondingly. Luo et al. reported PM-type OPDs with a 5 nm C60 as hole block layer between ITO and active layer which show a R of 1.33 A·W−1 under 850 nm (0.782 mW/cm2) at −3.0 V bias.[17] However, a higher evaporation temperature of C60 will be complex to fabricate. Nie et al. utilized a photon-controllable “valve" formed by PBDTT-DPP: PC71BM/MoO3 to accumulate and release external circuit carriers, realizing EQE>100%.[18] However, this structure needs donor material with its electron trap levels lower than the lowest unoccupied molecular orbital (LUMO) of PC71BM, so the materials selection will be limited. Recently, Wang et al. reported that an EQE of 53500% was achieved by trap-assisted PM via a small weight of PC71BM doped in P3HT.[19] Whereas this large EQE value brought defect of too high operating voltage (−60 V). From the perspective of fabrication process, trap doping is an easy method to realize PM.

In order to solve the shortcomings of the above-mentioned methods, such as special materials and too high working voltage, the bulk heterojunction (BHJ) structure of high exciton dissociation efficiency combined with the method of trap-assisted PM is applied to the OPDs. In this paper, we investigate the PM-type OPDs prepared with P3HT:PC61BM (and PBDT-TT-F:PC61BM) blends doped with low LUMO level of C60 (4.5 eV, lower than PC71BM LUMO level) using the one-step solution process. The PM mechanism of this OPD is investigated systematically, especially for the varying relation between PM efficiency and C60 concentration. Additionally, the device performance of the two donor–acceptor combinations is also compared. Therefore, our works may provide an easy method for obtaining highly sensitive OPDs with low working voltage, high R and D* simultaneously.

2. Experiments
2.1. Device fabrication

In this paper, the two donor–acceptor combinations are P3HT:PC61BM and PBDT-TT-F:PC61BM respectively, and the trap dopant is C60. The C60 concentrations in the P3HT:PC61BM (1:1 wt/wt) active layer were 0, 0.4 wt.%, 1.6 wt.%, and 4.8 wt.%, corresponding to device A, B, C, and D, respectively. For C60-doped PBDT-TT-F:PC61BM (1:1.5 wt/wt) devices, the C60 concentrations were 0, 0.3 wt.%, 1.3 wt.%, and 3.8 wt.% and these represent device E, F, G, and H correspondingly. Then the P3HT (or PBDT-TT-F): PC61BM:C60 blends were dissolved in 1 ml chlorobenzene, respectively. These solutions were all stirred for 12 hours at room temperature with a magnetic stirrer. All devices were fabricated on indium tin oxide (ITO)-coated substrates (sheet resistance: 15 Ω/□). The ITO substrates were cleaned with sequential sonication in deionized water, acetone, ethanol, and then dried by a nitrogen gun. The PEDOT:PSS was firstly spin-coated on the top of ITO with 3500 rpm for 1 min and was subsequently dried in ambient at 100 °C for 10 min. Then, the active layers were spin-coated on top of ITO/PEDOT: PSS layer with 500 rpm for 1 min and were also annealed at 100 °C for 10 min. Finally, a 100 nm cathode Al was thermally deposited on top with 0.3 nm·s−1 deposition rate at a vacuum degree of 5 × 10−4 Pa. The deposition rates and thickness were monitored by oscillating quartz monitors. The effective area of the devices fabricated was 1 cm2. Device structure reported here is ITO/PEDOT: PSS(35 nm)/P3HT:PC61BM (and PBDT-TT-F:PC61BM):C60(160 nm)/Al(100 nm), as shown in Fig. 1(a). Herein, PBDT-TT-F was purchased from SunaTech Inc. and the molecular structure of PBDT-TT-F is shown in Fig. 1(b). P3HT, PC61BM, PEDOT:PSS, and C60 were obtained from Xi’an Polymer Light Technology Corp.

Fig. 1. (a) Device structure diagram. (b) Molecular structure of PBDT-TT-F. (c) Energy level diagram of the OPDs.
2.2. Measurements and characterizations

The current density–voltage (JV) characteristics and capacitance–voltage (CV) characteristics of devices were measured by Keithley 2636B Semiconducting System and Keithley PCT-CVU Multi-frequency CV meter (frequency: 10 kHz), respectively. The light source is a λ = 460 nm commercially available diode (Edson, 5050) powered by different bias so as to vary light intensity (Pin). The light intensity was calibrated by a Newport 818-UV power meter. The photoluminescence spectra were recorded using a Transient Steady-state Fluorescence Spectrometer (Edinburgh FLS9, laser wavelength: 510 nm). The absorption spectra and thickness of the active layer films were measured using a UV–Vis spectrophotometer (PerkinElmer, Lambda 950) and Ellipsometer (VB-400), respectively. An atomic force microscopy (AFM, Dimension Icon) was used to analyze the morphology of the films. All the measurements were carried out at room temperature in the ambient condition.

2.3. Calculations

In this paper, the responsivity (R), normalized detectivity (D*), and external quantum efficiency (EQE) of the OPDs were calculated by the following equations:[15,18]

where Jlight is the current density under light, Jdark is dark current density, Jph is the photocurrent density which equals the difference between Jlight and Jdark at the same voltage, Pin is the incident light intensity, q is absolute value of electron charge, and is the energy of incident photon.

3. Results and discussion

Figure 1(c) depicts the energy level diagram of materials used here and we can know that the LUMO level of C60 is much lower than that of P3HT, PBDT-TT-F and PC61BM. So electron traps could be formed in the BHJ active layer after doping C60, as shown in Fig. 1(c), which will be confirmed by a series of tests later. In OPDs, the dark current density (Jdark) is one of the most important electrical parameters, which determines the detection ability of weak signals.[20] A main contribution to the Jdark is the charge carrier injection through the respective counter electrode, e.g., electrons being injected through the anode.[6] Due to the rather low electron mobility of PEDOT:PSS, the electrons are directly injected into the LUMO level of PC61BM from ITO anode.[8] Therefore, there exists a relatively high electron injection barrier of ∼ 1.1 eV and the hole injection barrier of ∼ 0.9 eV (1.0 eV) that prevents hole injection from Al cathode into the highest occupied molecular orbital (HOMO) level of P3HT (PBDT-TT-F) under reverse bias in darkness. Figures 2(a) and 2(b) show the JV characteristics of C60-doped P3HT:PC61BM and C60-doped PBDT-TT-F:PC61BM devices in dark condition and under 460 nm (Pin = 0.21 mW·cm−2) illumination, respectively. First of all, from the Jdark in the reverse bias regime, it is obvious that both of the two structure devices exhibit relatively low Jdark due to the high injection barrier. Additionally, Jdark is further decreased after doping C60, which may be attributed to the reduced electron injection from ITO into the LUMO level of PC61BM because some injected electrons are trapped in C60 near ITO. The decreased Jdark indicates that the incorporation of C60 has a significant influence on the charge injection and the transportation of charge carriers. In general, charge transport in organic semiconductor based diode follows a trap-limited space-charge-limited current (SCLC) model: JVk.[17] To directly investigate the difference of charge migration after doping C60, we further do SCLC analysis based on the dark JV curves of P3HT:PC61BM and P3HT:PC61BM:1.6 wt.% C60 devices under reverse bias, as shown in Fig. 2(c). In order to find the exponent k (Fig. 2(c)), we piecewise fit the measured data with the relation JVk. It can be seen that the slope k is close to 1 either for device A without C60 (k = 1.27) or device C doped with 1.6 wt.% C60 (k = 0.83) in the low voltage region, signifying a near Ohmic regime (k = 1) described by JΩ = qμNV/L.[21] Here q is the elementary charge, μ is the carrier mobility, N is the volume density of charge, V is the applied voltage, and L is the thickness of the active layer. It is generally accepted that Ohmic conduction originates from the low intrinsic conductivity of the organic semiconductor and the mobility is mainly influenced by the trapping of charge carrier.[22,23] Therefore, the slope k for device C is slightly lower than device A, which is caused by the decreased carrier mobility because that some injected electrons from ITO are trapped in C60. At moderate voltage, the injection carrier increases and space charge emerges inside the materials. Hence, JV relationship becomes nonlinear. When the trap states are not filled or do not affect the charge transport, referring to trap-free SCLC regime (TF-SCLC, k = 2), the JV relation is described by Mott–Gurney law[17]

where ε0 is the dielectric constant of vacuum, and εr is the dielectric constant of the organic active layer. Further, when the trap states are filled and then affect the charge transport, referring to trap-filling SCLC regime (TFLSCLC, k > 2). Here, the current density is defined by[17]
where Nv is the effective density of states, Nt is the total trap density, and l (l = k − 1) is an exponent with value larger than 1. Detailed interpretation about Eq. (5) can refer to Ref. [17]. A marked difference of the relation JVk for device A and device C from moderate voltage to high voltage should be noted. Device C doped with C60 undergoes a transition from TF-SCLC (k = 2.09) to TFLSCLC regime (k = 3.89), but device A without C60 is simply characterized by TFLSCLC model with k = 2.59 after low voltage. We ascribe the marked difference to the electron trapping ability of C60 doped in active layer, which will capture some electrons injected from ITO. So device C needs a process to fill the electron traps induced by doping C60 and thereby reduces the electron injection from ITO into PC61BM comparing with device A. By contrast, the conduction of device A after low-voltage region is governed by trap filling that stems from electron injection at moderate voltage without the influence of electron trapping. The results above indicate that C60 doped in the active layer could be used to effectively trap and accumulate photogenerated electrons under illumination.

Fig. 2. (a), (b) JV curves in darkness and and under 460 nm (0.21 mW·cm−2) illumination for the C60-doped P3HT:PC61BM and C60-doped PBDT-TT-F:PC61BM devices. (c) Experimental measured dark JV curves (symbols) and the simulated plots (red line, JVk) of the devices under reverse bias. (d) EQE–V curves of the two kinds of the active layer devices under 460 nm (0.21 mW·cm−2) illumination.

It can be clearly seen that the light current density (Jlight) of C60-doped devices exhibits an obvious improvement compared with their respective undoped devices. For example, device A shows only Jlight = 5.01 × 10−5 A·cm−2 under −1.0 V, however device B, C, and D exhibit larger Jlight (1.99 × 10−4, 2.55 × 10−4, and 1.68 × 10−4 A·cm−2 correspondingly). To give a more intuitive comparison, we calculate the EQE values based on light JV curves according to Eq. (3), and the EQE–V curves are shown in Fig. 2(d). It is apparent that all the EQE values of C60-doped devices have improved significantly. Especially, the EQE values of device B, C, and D are all greater than 100% at −1.0 V bias (255.2%, 327.5%, 215.9%, respectively). Whereas the EQE value of device A is no more than 100% even at −3.0 V bias. Furthermore, these EQE values increase first and then decrease as the increase of C60 concentration. It means that the electron traps are mainly induced by adding C60 rather than intrinsic defect in active layer films.[8] Noted that these EQE values of C60-doped devices increases with the applied voltage, but the EQE values of undoped device A are nearly unchanged as the applied bias increases. A similar phenomenon happens to C60-doped PBDT-TT-F:PC61BM devices. Thus, we can conclude that the improvement of device optoelectronic performances is attributed to the external carrier injection after doping C60. The detailed explanation about the EQE value of >100% will be discussed later.

The dispersed C60 in the active layer can be considered as electron traps, which has been discussed in the SCLC analysis above. Given that the electron traps will also influence the transportation of charge carriers generated by incident photons, we further do photoluminescence (PL) study and light-assisted capacitance measurements subsequently. PL is also a strong mean to prove the electron trapping ability of C60 doped in active layer,[13] which can trap electrons generated in P3HT and PC61BM based on its energy level, shown in Fig. 1(c). Figure 3(a) depicts the PL spectra of P3HT, P3HT:C60, P3HT:PC61BM, and P3HT:PC61BM:C60 films respectively. Emission from P3HT is obviously quenched by C60, demonstrating the electron transfer from P3HT to C60. The P3HT emission from the P3HT:PC61BM mixture is further quenched by adding C60, which confirms the electron transfer from P3HT:PC61BM mixture to C60 and can be predicted from their energy levels as well.

Fig. 3. (a) PL spectra of P3HT, P3HT:C60 (6:0.2 wt/wt), P3HT:PC61BM, and P3HT:PC61BM:C60 (6:6:0.2 wt/wt/wt) films. (b) CV curves and (c) MS plots of P3HT:PC61BM and P3HT:PC61BM:1.6 wt.% C60 devices in darkness and under 460 nm (5.54 mW·cm−2) illumination, respectively.

As is well known, light-assisted capacitance measurements have been used to study the device capacitance, which is determined by the charges at the electrode interfaces.[16,24] Figure 3(b) shows the CV curves of the P3HT:PC61BM:C60 devices in dark condition and under 460 nm (5.54 mW·cm−2) illumination, respectively. It can be clearly seen that the capacitance at zero bias increases from 41.7 nF to 45.9 nF (increments are 4.2 nF) and increases from 44.0 nF to 88.1 nF (increments are 44.1 nF) for undoped device A and 1.6 wt.% C60-doped device C, respectively. It implies that the photogenerated electrons do accumulate at the interface of active layer/Al electrode for the C60-doped devices under illumination. It is generally accepted that the capacitance of diode under small reverse and low forward voltage obeys Mott–Schottky equation[25]

where ε = ε0 εr, ε0 is the dielectric constant of vacuum, εr is the relative dielectric constant of the semiconductor (∼ 3.8), Nq is the carrier density, Vbi is the built-in voltage, and A is the area. Thus, the so-called Mott–Schottky plot of C−2 versus V yields a straight line. Then we can calculate the Nq according to the slope and the Vbi based on the extrapolated intersection with the voltage axis of the straight line respectively. In order to further understand the device physics of the PM-type OPDs here, we fit a straight line to the inclination point of the C−2 vs V relation according to Eq. (6), as shown in Fig. 3(c). For device C doped with C60, the calculated Nq is 5.20 × 1016 cm−3 in darkness, and increases up to 2.04 × 1017 cm−3 under illumination. By contrast, the Nq only increases from 5.72 × 1016 cm−3 up to 7.68 × 1016 cm−3 for device A. Noted that the Nq increment of device C (1.52 × 1017 cm−3) is much larger than device A (1.95 × 1016 cm−3), which is consistent in the capacitance increment. Hence, the light-induced electrons could be effectively captured by C60. When the electron traps in the active layer capture photogenerated electrons, the cathode side would be induced to accumulate holes, whereas the anode side might not accumulate charges (electron-rich layer). Therefore, the charge accumulation at the interface of the cathode/active layer causes the band bending of P3HT, which reduces the hole injection barrier at the cathode Al side. In fact, the barrier height (q(VbiV)) for device C does reduce to 0.71 eV, which is lower than hole injection barrier in darkness (0.9 eV). On the other hand, the high density of accumulated electrons will form a high local electric field near Al electrode (Fig. 4(b)), which narrows the space charge region (SCR) at active layer/Al interface inside device, thereby triggering trap-assisted hole tunneling injection from the Al electrode under reverse bias. Further, we calculate the width (w) of the SCR according to[25]
As expected, the w for device C decreases to 3.82 nm (under illumination) from 7.65 nm (in darkness), which is attributed to the electron accumulation in C60 and is beneficial for tunneling effect.[12,18] Unlike the photovoltaic process in photodiode type OPDs, the detailed operating mechanism of this PM-type OPDs can be summarized below.

Fig. 4. Schematic diagram of the operating mechanism of typical photovoltaic OPDs (a) and PM-type OPDs (b).

Incident photons absorbed in P3HT (PBDT-TT-F) and PC61BM firstly generate electron–hole pairs, which separate at the interface of P3HT (PBDT-TT-F)/PC61BM into free electrons and holes, for undoped devices, holes eventually end up on P3HT (PBDT-TT-F), and electrons end up on PC61BM, finally, these free charges transport in their respective channels and are collected by anode/cathode, which is sketched in Fig. 4(a). However, for C60-doped devices, photogenerated electrons are captured by C60 traps due to the lowest LUMO energy level among all these components. The accumulated electrons near the cathode side will also induce hole accumulation of Al side (Fig. 4(b)), resulting in the band bending of P3HT (PBDT-TT-F) and the increase of Al work function.[18,19,26] As a result, the barrier width (SCR of active layer/Al interface) will be reduced, which is sketched in Fig. 4(b). When the band bending reduces hole wavelength to less than or equal to the barrier width under enough illumination, the cathode holes can tunnel through the thin barrier region into the HOMO level of P3HT (PBDT-TT-F) under reverse bias. Now, these external hole injection current plus photogenerated current together contributes to the EQE value of >100%.[1219] It should be noted that the electron transfer from P3HT (PBDT-TT-F) to C60, either direct transfer or indirect transfer through PC61BM, is paramount for the observed improvement of device performance.

In order to exclude the other factors that may influence the device performance to some extent, such as the morphology and the light absorption of active layer film,[2729] next we investigate the difference of films after doping C60 based on their AFM images and absorption spectra. Figures 5(a) and 5(b) show the AFM images of P3HT:PC61BM:C60 films with different C60 concentrations. The neat P3HT:PC61BM film exhibits a surface morphology with a root mean square roughness (RMS) of 1.26 nm, whereas the P3HT:PC61BM:1.6 wt.% C60 film exhibits an RMS of 1.01 nm. It is believed that there are not significant changes of morphology in these two films. Meanwhile, from the absorption spectra of the films shown in Fig. 5(c), we can observe that no apparent change of absorbance occurs between the undoped-films and C60-doped films. Thus, the film morphology and light absorption cannot be the real reason for improved device performance.

Fig. 5. AFM images of films for (a) P3HT:PC61BM and (b) P3HT:PC61BM:1.6 wt.% C60. (c) Absorption spectra of P3HT:PC61BM:C60 (0 and 1.6 wt.%) and PBDT-TT-F:PC61BM:C60 (0 and 1.3 wt.%), respectively.

It is known that the responsivity (R) and normalized detectivity (D*) of the OPDs strongly depend on the voltage and photon wavelength.[16] Tables 1 summarizes the optoelectronic performances for P3HT:PC61BM:C60 and PBDT-TT-F:PC61BM:C60 OPDs under −1.0 V bias, in which the C60 doping concentration is changed. It should be noted that the PM happening inside devices doped with C60 is the key reason for the observed improvement of device performance. Specifically, the device performance improves first and then degrades complying with the increase of C60 concentration. We attribute the main factor of the variation of device performance to the cathode hole tunneling injection. If C60 concentration is lower, the trap numbers of the active layer near Al side will decrease as well. Hence, the trapped electrons in C60 near the Al side will reduce, which induces weaker band bending and triggers less cathode hole tunneling injection. If C60 concentration is higher, small molecule C60 would like aggregating to form some isolated clusters or even continuous electron transport channels.[8] A part of photogenerated electrons may be collected by cathode through these channels. Therefore the trapped electrons in the active layer near Al side will be also decreased, thereby resulting in the less cathode hole tunneling injection. Only when the C60 concentration is moderate, the cathode hole tunneling injection can be effectively enhanced, which is supported by the optimal performance for device C and G in their respective active layer system.

Table 1.

Optoelectronic performances for P3HT:PC61BM:C60 (A, B, C, and D) and PBDT-TT-F:PC61BM:C60 (E, F, G, and H) OPDs under −1.0 V and comparison with other reported PM-type OPDs.

.

It is noteworthy that the C60-doped P3HT:PC61BM devices show better optoelectronic performance comparing with C60-doped PBDT-TT-F:PC61BM devices according to Table 1. We consider there are two main factors contributing to the better performance of P3HT:PC61BM:C60 devices. The one factor is the higher exciton dissociation efficiency of P3HT:PC61BM active layer due to larger energy offset of the LUMO energy levels of the donor and acceptor materials (P3HT/PC61BM:0.7 eV, PBDT-TT-F/PC61BM:0.3 eV).[30] Hence, more photogenerated electrons are trapped in C60-doped P3HT:PC61BM active layer and higher efficiency of PM can be obtained. Another factor of performance improvement for P3HT:PC61BM:C60 devices is the increased hole injection from Al cathode into the HOMO level of donor. In order to compare the hole injection to different donor materials, we measured the JV curves of the PM-type OPDs based on ITO/PEDOT:PSS/P3HT:C60 (and PBDT-TT-F:C60)/Al with all the weight ratio of 10:0.3 in darkness and under 460 nm (Pin = 5.54 mW·cm−2) illumination, respectively, as shown in Fig. 6(a). Because of the smaller hole injection barrier between the Al work function and the HOMO level of P3HT (0.9 eV, PBDT-TT-F/Al: 1.0 eV), the Jdark of P3HT:C60 device is slightly larger than PBDT-TT-F:C60 device. In fact, the higher Jlight of P3HT:C60 device at larger reverse bias regime (larger than −4.5 V) suggests more cathode hole injection into the HOMO level of P3HT. Figure 6(b) further depicts the EQE–V curves of P3HT:C60 and PBDT-TT-F:C60 devices based on their light JV curves. It is clearly observed that when the EQE value is equal to 100%, the voltage of P3HT:C60 device is about −5.1 V, which is lower than that of PBDT-TT-F:C60 device (−5.9 V). The curve for P3HT:C60 device increases much faster at large reverse bias, as shown in Fig. 6(b). Consequently, we can conclude that more cathode hole can be injected into the donor material with shallower HOMO level.

Fig. 6. (a) JV curves in darkness and under 460 nm (5.54 mW·cm−2) illumination for the P3HT:C60 (10:0.3, wt/wt) and PBDT-TT-F:C60 (10:0.3, wt/wt) devices. (b) EQE–V curves of the corresponding P3HT:C60 and PBDT-TT-F:C60 devices under 460 nm (5.54 mW·cm−2) illumination.

Table 1 also gives the comparison of optoelectronic performance of the OPDs fabricated in this paper with other PM-type OPDs. In short, for 1.6 wt.% C60-doped P3HT:PC61BM device C, the high R of 1.21 A·W−1 is ∼5-fold higher than that of photodiode-type OPDs (λ = 500 nm, U = −1 V, R = 0.273 A·W−1),[6] the high D* of 4.22 × 1012 Jones is ∼ 4-fold higher than that of PM-type OPDs under the same bias of −1.0 V (λ = 550 nm, Pin = 3.09 μW·cm−2, D* = 1.23 × 1012 Jones),[16] which is comparable or even superior to some inorganic photodetectors.[7,11,31] Moreover, under the premise of the same EQE, device C shows a much lower working voltage compared with the PM-type OPDs reported in Ref. [8] (−6 V). Although the EQE values are not greater than 1000%, the PM-type OPDs reported here with facile fabrication and relatively high device performance can solve the problem of too high working voltage and the limitation of special materials.[18,19]

4. Conclusion

In summary, we have demonstrated a one-step solution processed PM-type OPD based on C60-doped bulk heterojunction structure at low bias. Using the advantages of BHJ with high exciton dissociation efficiency as active layer, large PM under low bias can be achieved by doping a small weight of C60 as traps. A schematic energy-level model combined with experimental measurements is capable of explaining the operating mechanism of the PM-type OPDs proposed here. The PM originates from cathode hole tunneling injection assisted by the trapped electrons in C60 near the Al side. The efficiency of PM is related to the C60 concentration in the active layer. In order to further improve the device performance, the donor material should have higher LUMO level and shallower HOMO level. On the basis, the 1.6 wt.% C60-doped P3HT:PC61BM device exhibits EQE = 327.5%, R = 1.21 A·W−1, and D* = 4.22 × 1012 Jones at −1 V under 460 nm (0.21 mW·cm−2) illumination. This paper provides a universal method of designing high PM OPDs with low operating voltage and facile fabrication.

Reference
[1] Ross D Ardalan A Ajay K P Paul L B Paul M 2016 Adv. Mater. 28 4766
[2] Baierl D Pancheri L Schmidt M Stoppa D Betta G F D Scarpa G Lugli P 2012 Nat. Commun. 3 1175
[3] Lyons D M Armin A Stolterfoht M Nagiri R Vuuren R D Pal B N Burn P L Lo S C Meredith P 2014 Org. Electron. 15 2903
[4] Li W H Li S B Duan L Chen H J Wang L D Dong G F Xu Z Y 2016 Org. Electron. 37 346
[5] Du L L Luo X Zhao F Y Lv W L Zhang J P Peng Y Q Tang Y Wang Y 2016 Carbon 96 685
[6] Valouch S Hönes C Kettlitz S W Christ N Do H Klein M Kalt H Colsmann A Lemmer U 2012 Org. Electron. 13 2727
[7] Xiao B Zhang M L Wang H B Liu J Y 2017 Acta Phys. Sin. 66 228501 in Chinese
[8] Li L L Zhang F J Wang W B An Q S Wang J Sun Q Q Zhang M 2015 ACS Appl. Mater. & Interfaces 7 5890
[9] Shen L Fang Y J Wei H T Yuan Y B Huang J S 2016 Adv. Mater. 28 2043
[10] Nie R M Zhao Z J Deng X Y 2017 Synth. Met. 227 163
[11] Zhang Z Z Fu Z L Guo X G Cao J C 2018 Chin. Phys. 27 030701
[12] Gao M L Wang W B Li L L Miao J L Zhang F J 2017 Chin. Phys. 26 018201
[13] Dong R Bi C Dong Q F Guo F W Yuan Y B Fang Y J Xiao Z G Huang J S 2014 Adv. Opt. Mater. 2 549
[14] Fang Y J Guo F W Xiao Z G Huang J S 2014 Adv. Opt. Mater. 2 348
[15] Li L L Zhang F J Wang J An Q S Sun Q Q Wang W B Zhang J Teng F 2015 Sci. Rep. 5 9181
[16] Wang Y Zhu L J Hu Y F Deng Z B Lou Z D Hou Y B Teng F 2017 Opt. Express 25 7719
[17] Luo X Lv W L Du L L Zhao F Y Peng Y Q Wang Y Tang Y 2016 Phys. Status Solidi RRL 10 485
[18] Nie R M Deng X Y Feng L Hu G G Wang Y Y Yu G Xu J B 2017 Small 13 1603260
[19] Wang W B Zhang F J Du M D Li L L Zhang M Wang K Wang Y S Hu B Fang Y Huang J S 2017 Nano Lett. 17 1995
[20] Binda M Iacchetti A Natali D Beverina L Sassi M Sampietro M 2011 Appl. Phys. Lett. 98 073303
[21] Nismy N A Jayawardena K I Adikaari A D T Silva S R P 2015 Org. Electron. 22 35
[22] Goodman A M Rose A 1971 J. Appl. Phys. 42 2823
[23] Li A Y Miao X R Deng X Y 2013 Synth. Met. 168 43
[24] Gong W Xu Z Zhao S L Liu X D Yang Q Q Fan X 2014 Acta Phys. Sin. 63 078801 in Chinese
[25] Kirchartz T Gong W Hawks S A Agostinelli T MacKenzie R Yang Y Nelson J 2012 J. Phys. Chem. 116 7672
[26] Gao Y L 2010 Mater. Sci. Eng. 68 39
[27] Wei G Wang S Renshaw K Thompson M E Forrest S R 2010 ACS Nano 4 1927
[28] Liu R Xu Z Zhao S L Zhang F J Gao X N Kong C Cao W Z Gong W 2011 Acta Phys. Sin. 60 058801 in Chinese
[29] Deng L J Zhao S L Xu Z Zhao L Wang L 2016 Acta Phys. Sin. 65 078801 in Chinese
[30] Zafar Q Fatima N Karimov K S Ahmed M M Sulaiman K 2017 Opt. Mater. 64 131
[31] Sun L Wang L Lu J L Liu J Fang J Xie L L Hao Z B Jia H Q Wang W X Chen H 2018 Chin. Phys. 27 047209