High performance solution-processed infrared photodetector based on PbSe quantum dots doped with low carrier mobility polymer poly(N-vinylcarbazole)

Muhammad Sulamana, Shengyi Yang*ab, Arfan Bukhtiarc, Chunjie Fua, Taojian Songa, Haowei Wangc, Yishan Wangc, He Boa, Yi Tangd and Bingsuo Zoua
aBeijing Key Laboratory of Nanophotonics and Ultrafine Optoelectronic Systems, School of Physics, Beijing Institute of Technology, Beijing 100081, China. E-mail: syyang@bit.edu.cn; Tel: +86-10-68918188
bState Key Lab of Transducer Technology, Chinese Academy of Sciences, China
cBeijing Key Laboratory of Nanophotonics and Ultrafine Optoelectronic Systems, School of Materials Science and Engineering, Beijing Institute of Technology, Beijing 100081, China
dKey Lab of Photoelectronic Imaging Technology and System, Ministry of Education, School of Optoelectronics, Beijing Institute of Technology, Beijing 100081, China

Received 3rd December 2015 , Accepted 18th April 2016

First published on 20th April 2016


Abstract

Colloidal quantum dots (CQDs) are promising materials for flexible electronics, light sensing and energy conversion. In particular, as a narrow bandgap semiconductor, lead selenide (PbSe) CQDs have attracted considerable interest due to their potential applications in infrared (IR) optoelectronics such as IR light-emitting diodes (LEDs), photodetectors and solar cells. Solution-processed photodetectors are more attractive owing to their flexible, large-scale and low-cost fabrication, and their performance depends greatly on the film quality and surface morphology. In this study, a high performance solution-processed infrared photodetector based on PbSe CQDs blended with low hole mobility polymer poly(N-vinylcarbazole) (PVK) is presented. In order to obtain a higher device performance, different volume ratios (K = VPVK/VPbSe) of PVK (20 mg ml−1 in chloroform) in PbSe CQDs (15 mg ml−1 in chlorobenzene) were investigated, and a maximum responsivity and specific detectivity of 2.93 A W−1 and 1.24 × 1012 jones, respectively, were obtained at VG = −20 V under 30 mW cm−2 980 nm laser illumination for field-effect transistor (FET)-based photodetector Au(S&D)/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al(G), in which PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite with K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 acts as the active layer and poly (methyl methacrylate) (PMMA) as the dielectric layer. The reasons for the high device performance of PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite as an active layer are discussed, in which PbSe nanoparticles were blended with low hole mobility polymer PVK but showed comparable detectivity as that blended with regioregular P3HT. Moreover, all these types of photodetectors are very stable for reverse fabrication using PMMA dielectric layer to shield the active layer from the environment and by inorganic ligand exchange treatment on the active layer.


1 Introduction

Recently, colloidal quantum dots (CQDs), as a candidate material for a range of optoelectronic applications, have attracted substantial attention1,2 because they possess a tunable energy bandgap, a large absorption cross-section and a low-cost solution processibility.3,4 The appeal of these characteristics was emphasized in optoelectronic devices such as quantum dot solar cells,5 infrared photodetectors,6 and high-efficiency infrared light-emitting diodes (LEDs).7 In particular, solution processibility of CQDs is the main benefit for its ease and low-cost integration with almost all varieties of substrates for the fabrication of optoelectronic devices.8,9 Optical absorption and emission spectra of CQDs can be widely tuned through the effect of size-dependent quantum confinement.10

As a narrow bandgap semiconductor, lead selenide (PbSe) CQDs are a low-cost solution process, possessing a tunable energy bandgap and high luminescence quantum efficiency.11–13 Various types of nanomaterials, including PbSe CQDs and their nanocomposites, have shown responsive spectra ranging from the ultraviolet to near-infrared,6,14–17 The use of solution-processed thin films of colloidal organic/inorganic semiconductor nanoparticles (or CQDs) as photoconductor active layers has been a critical step for researchers to fabricate low-cost photodetectors.18,19 Being an IV–VI semiconductor with a large Bohr diameter of excitons (46 nm), the electronic structure of PbSe CQDs is strongly quantum confined, and the lowest energy transitions in these materials can be tuned into the near and mid-infrared region. On the other hand, organic semiconductors and/or polymers and CQDs and their nanocomposites have attracted considerable research interests over the past two decades for their use in solid-state optoelectronic devices. To avoid the drawback of low mobility of charge carriers and low photoconductivity that appears in pure polymer-based devices, charge carrier generation can be enhanced, which enables the broadening of photoresponse through their size-tunable optoelectronic properties by blending organic polymers into inorganic CQDs as the active layer. In this way, such a type of optoelectronic devices take advantage of the dependence of the bandgap of CQDs on particle size to absorb (or emit) light in tunable wavelengths.

Among semiconducting polymers, favorable properties of poly(N-vinylcarbazole) (PVK), such as high transparency, good film-forming, mechanical flexibility and its solution process, have attracted significant attention for applications in hybrid organic/inorganic devices.20 PVK is a hole-transporting organic semiconductor (hole mobility μh is much greater than electron mobility μe), which is widely used in organic LEDs (OLEDs), xerography and commonly used as an electro-optically energetic polymer due to its specific photo-physical properties.8 Poly(3-hexylthiophene) (P3HT) is another p-type organic polymer, and a hole mobility of about 0.1 cm2 V−1 s−1 (highest) has been reported21,22 and good device performance has been obtained in photodetectors.23 Therefore, it is reasonable for us to think about what the case would be if we were to use lower hole mobility (between 10−7 and 10−6 cm2 V−1 s−1)24–26 PVK to replace P3HT and to make FET-based IR photodetectors?

Therefore, based on our previous work,23 in this study, first we synthesize size-tailored PbSe CQDs and then incorporate them into the hole-transporting polymer PVK as the active layer to make a field-effect transistor (FET)-based infrared photodetector Au(S&D)/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al(G). Our experimental results showed that the photosensitive spectrum of the PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite extended into the IR region, and different volume ratios (K = VPVK/VPbSe) of PVK (20 mg ml−1 in chloroform) to PbSe CQDs (15 mg ml−1 in chlorobenzene) were investigated, and a maximum responsivity and high specific detectivity of 2.93 A W−1 and 1.24 × 1012 jones, respectively, were obtained at VG = −20 V under 30 mW cm−2 980 nm laser illumination for PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite with K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 as active layer. Our experimental data show that the large interfaces in the PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposites extend an incredible opportunity for efficient exciton dissociation after photogeneration in the PbSe QDs. Efficient exciton dissociation and charge transfer depend on the difference between the potential energy and the electron affinity of the photoactive species. Carrier transport in the nanocomposite layer can be improved under an electric field by the hopping of holes and electrons among the carbazole groups of PVK polymer chains.27 Therefore, in this way, charge transportation in active layer can be enhanced.

2 Experimental

PbSe CQDs were synthesized as by the previously reported technique28 with some modification. All syntheses were performed in a nitrogen-filled glove box. A solution of 1.28 g selenium (Se) in 12.8 ml of tri-n-octylphosphine (TOP) was prepared using ultrasonication. Another solution was prepared by mixing 1.784 g of lead oxide (PbO) in 32.6 ml of octadecene (ODE) and 5.2 ml of oleic acid (OA) in a three-neck flask using magnetic stirring by removing air for 20 minutes and introducing nitrogen for 15 minutes; the process was done 3 times each. The solution was then heated by gradually increasing the temperature of the solution, and it reached the desirable temperature (180 °C). After adjusting the solution to the desired temperature, the mixture of Se and TOP was rapidly injected into this hot solution. PbSe CQDs were grown at the desired temperature for an optimal time of 2 min, and then the reaction was stopped by rapidly injecting 15 ml of n-hexane into the solution and placing the 3-neck flask in a cold water bath. To make the quantum dots more conductive and stable, a solution of ammonium chloride (NH4Cl) in methyl alcohol was added to the solution at 60 °C for 10 min.29 The synthesized PbSe nanocrystals were washed twice with n-hexane/isopropyl alcohol/ethanol and once with n-hexane/acetone and then dried in a vacuum oven.

First, PVK (Sigma-Aldrich, Mw is 1[thin space (1/6-em)]100[thin space (1/6-em)]000, purity is 97%) and PbSe CQDs were dissolved in chloroform and chlorobenzene with a concentration of 20 mg ml−1 and 15 mg ml−1, respectively, and then both solutions were blended together at different volume ratios (K = VPVK[thin space (1/6-em)]:[thin space (1/6-em)]VPbSe) of K = 2[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]2 by continuous stirring at room temperature for 24 hours to produce the PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposites with uniform and continuous interfaces between PVK and PbSe CQDs.

The concept of a phototransistor, i.e. a photo-field-effect transistor,30 offers an attractive possibility: the gain, united with a lowered dark current as compared to the photoconductor, can be achieved if the thickness of the current-carrying channel can be chosen independently of the thickness of the light-absorbing layer. In the case of conventional FET-based photodetectors, the applied gate bias can modulate the lateral field across the active layer between source and drain electrodes; the electric field causes separation of the photogenerated carriers and greatly extends carrier recombination lifetime, leading to a much higher efficiency. Therefore, herein, we chose the FET configuration to fabricate the photodetectors; its cross-section diagram is shown in Fig. 1. First, an array of 60 nm thick Au electrodes (separated from each other by 100 μm) was deposited on a glass substrate as the source and drain electrodes by thermally evaporating through a shadow mask, which is further contacted with ITO for ease of electrical biasing. Then, the active layer of PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite and dielectric layer of poly(methyl methacrylate) (PMMA) were prepared by spin-coating PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite and PMMA (80 mg ml−1 in acetone solution) at 2000 rpm and baked at 100 °C and 120 °C, respectively, for 20 min each, forming ∼100 nm active layer and ∼930 nm dielectric layer. Finally, 100 nm Al was thermally evaporated as the gate electrode and annealed at 80 °C for 10 min.


image file: c5ra25761a-f1.tif
Fig. 1 Cross-section diagram of the FET-based photodetector and the electrical circuits for its characterization.

All fabrications were carried out in a nitrogen-filled glove box, and all characterizations were carried out in air at room temperature.

3 Results and discussions

Fig. 2a and b show the transmission electron microscopy (TEM) images of PbSe CQDs and PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite (K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2), respectively, in which the average size of PbSe nanoparticle is about 4–5 nm; Fig. 2c and d present TEM images of PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposites for K = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 and K = 2[thin space (1/6-em)]:[thin space (1/6-em)]1, respectively. From here, one can observe that a good quality PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite film can be obtained by increasing the volume ratio of PbSe; conversely, the self-assembly of nanocrystals was destroyed if the volume ratio of PVK was too high, resulting in a decrease in their photosensitivity. Normally, photosensitized CQDs having large and continuous interfaces with polymeric PVK in a certain volume ratio (see Fig. 2b) can show higher photoresponsivity and specific detectivity. Therefore, in our experiments, we chose the PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite with K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 as the optimal active layer to fabricate the IR photodetectors.
image file: c5ra25761a-f2.tif
Fig. 2 TEM image of PbSe CQDs (a) and PVK[thin space (1/6-em)]:[thin space (1/6-em)]PbSe nanocomposite of K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 (b), K = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 (c) and K = 2[thin space (1/6-em)]:[thin space (1/6-em)]1 (d), and SEM image of PVK[thin space (1/6-em)]:[thin space (1/6-em)]PbSe (K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2) thin film (e).

In order to confirm the nature of PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite, energy dispersive X-ray (EDX) spectroscopy of PbSe CQDs and PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK (K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2) nanocomposite films on Si substrate, respectively, was carried out. Their intensity peaks are shown in Fig. 3a and b, verifying the presence of all the elements in PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite. PbSe CQDs and PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposites (K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2) were redispersed in tetrachloroethylene and chloroform, respectively, and their absorption spectra are shown in Fig. 4, along with the absorption spectrum of PVK. From here, we can see that the absorption spectrum of PVK peaks at ∼400 nm and there is no absorption in near-infrared region, whereas the absorption spectrum of PbSe peaks at 1200 nm. Obviously, the absorption spectrum of the PVK[thin space (1/6-em)]:[thin space (1/6-em)]PbSe nanocomposite is the superposition of the two original constituents, indicating it is only a physical blending process and the presence of PbSe QDs can extend the absorption into the infrared region.


image file: c5ra25761a-f3.tif
Fig. 3 EDX spectrum for PbSe nanocrystals (a) and PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite (K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2) (b).

image file: c5ra25761a-f4.tif
Fig. 4 Absorption spectra of PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK (K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2); inset shows absorption spectra of PVK (blue) and PbSe CQDs (red).

An appropriate combination of the constituents in the blended nanocomposite contributes to the transport of charge carriers. Therefore, it is easy for one to know that PbSe CQDs play an essential role in terms of the effective absorption of photons in the infrared region. Fig. 5 illustrates the Raman spectrum of the synthesized CQDs, and all the Raman peaks exactly match with the previously reported Raman peaks of PbSe nanocrystals,31,32 confirming the purity of PbSe CQDs. The inset of Fig. 5 demonstrates the X-ray diffraction (XRD) pattern of PbSe CQDs and the PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite. PVK is an organic polymer having no sharp peaks, so the XRD pattern of PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite will alter with that of PbSe. The XRD pattern of the PbSe CQDs agrees perfectly with PbSe bulk,33,34 and its TEM is also consistent with its XRD pattern. Therefore, one can see that the PVK and PbSe are completely blended with each other. Furthermore, Fourier transform infrared (FTIR) spectra of PbSe nanocrystals, pure PVK and PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite are shown in Fig. 6. Therefore, all these results could confirm that the synthesized nanocrystals are really PbSe CQDs and the composite formed is really PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite.


image file: c5ra25761a-f5.tif
Fig. 5 Raman spectrum of PbSe nanocrystals; inset shows X-ray diffraction patterns for PbSe nanocrystals (blue) and PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite (red).

image file: c5ra25761a-f6.tif
Fig. 6 Fourier transform infrared (FTIR) spectra of PbSe CQDs (black), PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite (red) and pure PVK (blue).

In order to improve the performance of the photodetector, we fabricated a field-effect transistor (FET)-based photodetector Au(source & drain)/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al(gate) (see Fig. 1). The main limiting factor for the performance of CQDs-based photodetectors is the high dark current noise due to the narrow bandgap and the high density of the states of PbSe CQDs as the active layer. A disadvantage of PbSe CQDs is their poor stability in ambient air and that they can be destabilized easily; therefore, their applications are limited.35 In order to increase the interparticle coupling and carrier mobility, longer insulating oleate ligands should be exchanged by shorter and more conductive surfactant molecules such as ethanedithiol (EDT) and/or 3-mercaptopropionic acid (MPA).36,37 In our experiments, we performed inorganic ligand exchange on the active layer to enhance the stability of the PbSe CQDs.29 Therefore, unlike the conventional schematic of FET-based device fabrication, a reverse fabricated device was made in our experiments, i.e. the active layer first and then the dielectric layer.23 In this way, the dielectric layer can cover the active layer and prevent it from atmospheric air, and the active layer can be protected from oxidation. Consequentially, an enhanced stability of device can be achieved.23,29 The threshold voltage (Vth) of the FET-based photodetector was measured at VDS = 5 V, and Vth = 3.5 V was obtained, as shown in Fig. 7; other transfer curves at different VDS in dark are shown in the inset (Fig. 7).


image file: c5ra25761a-f7.tif
Fig. 7 VG vs. ID (black) at VDS = 5 V, and VG vs. SQRT of ID (blue); the straight line indicates threshold voltage (red); inset shows other transfer curves at different VDS.

The first feature of a photodetector is sensitivity, the capability of a photodetector to distinguish an incident optical signal from noise.38,39 Photosensitivity (P) of a device can be measured as P = Iphoto/Idark, where Iphoto = Iill.Idark, and Iphoto is the photocurrent, Iill. shows the drain current under illumination power intensity, and Idark is the drain current in the dark. By using the abovementioned formula for photosensitivity, the photosensitivity of the FET-based photodetector is measured as P = 67.5. The output characteristics of FET-based photodetector Au/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al in the dark and under different illumination intensities are shown in Fig. 8; the curve of photocurrent vs. illumination intensity is shown in the inset. From here, one can see that the photocurrent increases with increasing the illumination intensity up to a certain value, and then the photocurrent is saturated with further increasing of illumination intensity. In our experiments, therefore, we chose 980 nm laser illumination with an intensity of 30 mW cm−2 to illuminate the photodetector.


image file: c5ra25761a-f8.tif
Fig. 8 Output characteristics of FET-based photodetector Au/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al at VG = −20 V in the dark and under different illumination intensities; inset shows the curve of photocurrent vs. incident illumination intensity.

Photoresponsivity (R) is key factor for characterizing the performance of the photodetectors,27,40 and it is defined by R = Iph/Pill., where Iph = Iill.Idark, and Pill. is the incident illumination power intensity. The performance of the FET-based photodetectors in which PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite is composed of different K as active layer is listed in Table 1. From here, one can see that the device performance may be enhanced when we increase the volume ratio K to 1[thin space (1/6-em)]:[thin space (1/6-em)]2, and a high responsivity of 2.93 A W−1 was obtained for this case. Another important factor for evaluating the performance of a photodetector is its specific detectivity (D*), and it can be defined by the following equation:

 
image file: c5ra25761a-t1.tif(1)
where q is charge of electron, and Idark is dark current and A is effective area (2 mm × 2 mm) of the photodetector under illumination. By using eqn (1), a specific detectivity of 1.24 × 1012 jones can be obtained for FET-based photodetector Au(S&D)/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al(gate) in which PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK (K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2) nanocomposite acted as the active layer under 30 mW cm−2 980 nm irradiation at VG = −20 V. Obviously, from here one can see that this D* value is comparable to that of the previously reported PbSe-based IR-photodetectors in which PbSe was blended with high hole mobility polymer P3HT.23

Table 1 Performance of the FET-based photodetectors Au/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al in which the active layer is in different volume ratios K under same illumination conditions
Volume ratio of PVK to PbSe Responsivity (R, A W−1) Specific detectivity (D*, jones)
K = 2[thin space (1/6-em)]:[thin space (1/6-em)]1 0.0266 1.13 × 1010
K = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 0.0586 2.48 × 1010
K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 2.93 1.24 × 1012


Furthermore, it is natural for someone to ask how can we obtain such a high specific detectivity of a FET-based photodetector with low hole mobility PVK as dopant? To know the reason, we have to know the factors which determine the detectivity of the photodetector. As we know, the performance of the photodetector really depends on the charge carrier mobility, as well as the charge carrier injection barriers at the PVK/QDs' interface. Morphologically, PVK is an amorphous hole-conducting polymer, constituted of linear chains of repeated molecular groups (H2C–HC)n, with suspended carbazole side groups [(C6H4)2NH] randomly arranged around the same chain. Fig. 9 shows the schematic energy-level diagram of materials used in our experiments and the possible charge carrier transport under IR light illumination. The work function (Φ) of Au is about −5.1 eV. The highest occupied molecular orbital (HOMO) energy level and the lowest unoccupied molecular orbital (LUMO) energy level for PVK and PbSe QDs are −5.5 eV and −2.3 eV,41 −5.4 eV and −4.4 eV, respectively. It is worth mentioning that HOMO and LUMO of PbSe CQDs were determined by cyclic voltammetry (not shown here), and they are in agreement with the absorption spectra of PbSe nanocrystals (see Fig. 4) and the energy levels of PbSe QDs reported previously.42 Therefore, the injection barrier for electrons and holes photogenerated in PbSe QDs is about 2.1 eV and 0.1 eV. On the other hand, in the case of being with P3HT, the injection barrier for electrons from PbSe to P3HT is about 1.4 eV, and there is no barrier for holes from PbSe to P3HT. In principle, CQDs absorb photons and then photogenerated holes exit after excitons disassociate at the P3HT/QDs interface and then transfer to P3HT under VDS; however, photogenerated electrons will be trapped and accumulated in CQDs with increasing incident laser intensity,43 forming a built-in electric field at the P3HT/QDs interface, and will increase the potential of the channel, which can be regarded as a floating body effect of the transistor.44,45 Sequentially, the built-in electric field at the P3HT/QDs interface induced by the accumulated electrons will retard the charge transfer at the interface and induce a lower photosensitivity at higher incident light intensity. On the other hand, drift mobilities of both charge carriers are strongly electric field- and temperature-dependent. At room temperature and in an electric field of 2 × 105 V cm−1, the effective hole mobility46 is about 7.14 × 10−6 cm2 V−1 s−1. The electron mobility47 of P3HT is about 10−9 cm2 V−1 s−1. Basically, device performance depends on the sum of the mobility of electrons and holes, and their mobility in turn can be improved under an electric field; however, hole mobility plays the key role since electrons are minorities in a PVK- or P3HT-based photodetector. We found that the effective mobility of major charge carriers (i.e. holes) in P3HT[thin space (1/6-em)]:[thin space (1/6-em)]PbSe or PVK[thin space (1/6-em)]:[thin space (1/6-em)]PbSe as active layer are of the same order, i.e. ∼1.27 × 10−4 cm2 V−1 s−1 and ∼4.6 × 10−4 cm2 V−1 s−1 for PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK nanocomposite film and PbSe[thin space (1/6-em)]:[thin space (1/6-em)]P3HT nanocomposite film, respectively. On the other hand, we noted that the photosensitivity of FET-based photodetector Au/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]P3HT/PMMA/Al was less than that of photodetector Au/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al, and one possible reason may be the occurrence of exciton–exciton annihilation within the P3HT layer at higher light intensity, resulting effectively in a decrease in the exciton lifetime.48 Consequently, the specific photodetectivity will also be reduced, therefore, we got similar specific detectivities in the order of 1012 jones. Certainly, a more detailed theoretical simulation will be carried out for the nonlinear photosensitivity of the device and for the exciton–exciton annihilation within the PbSe[thin space (1/6-em)]:[thin space (1/6-em)]P3HT nanocomposite film.


image file: c5ra25761a-f9.tif
Fig. 9 Schematic energy-level diagram of each material used in our experiments (a) and charge carriers transport under IR light illumination (b).

Furthermore, higher D* could be obtained under even weaker incident illumination intensity since D* decreases with increasing incident illumination intensity. Moreover, higher D* could be obtained after further optimizing the device configuration. In this way, by taking advantage of their conductive properties and the size tunability of PbSe QDs, more applications for PbSe nanocrystals and their nanocomposites can be realized in novel optoelectronic devices such as high-sensitivity near/mid infrared photodetectors and high-efficiency solar cells.

4 Conclusion

In summary, we successfully demonstrated a high performance solution-processed FET-based IR-photodetector Au(S&D)/PVK[thin space (1/6-em)]:[thin space (1/6-em)]PbSe/PMMA/Al(G). By varying the volume ratios of PVK (20 mg ml−1 in chloroform) in PbSe CQDs (15 mg ml−1 in chlorobenzene), we found that the FET-based photodetector Au(S&D)/PVK[thin space (1/6-em)]:[thin space (1/6-em)]PbSe/PMMA/Al(gate), in which PVK[thin space (1/6-em)]:[thin space (1/6-em)]PbSe nanocomposite of K = 1[thin space (1/6-em)]:[thin space (1/6-em)]2, showed a high photosensitivity and photodetectivity of 67.5, 2.93 A W−1 and 1.24 × 1012 jones, respectively, at VG = −20 V under 30 mW cm−2 980 nm laser illumination. We discussed the reasons for such a high device performance of a FET-based photodetector Au(S&D)/PbSe[thin space (1/6-em)]:[thin space (1/6-em)]PVK/PMMA/Al(G), in which PbSe CQDs blended with low hole mobility PVK acts as the active layer, can be obtained as compared with PbSe CQDs doped with regioregular P3HT as the active layer. Our experimental data showed that the effective mobility of the major charge carriers (i.e. holes) in P3HT[thin space (1/6-em)]:[thin space (1/6-em)]PbSe or PVK[thin space (1/6-em)]:[thin space (1/6-em)]PbSe as active layer are of the same order of 10−4 cm2 V−1 s−1. Moreover, this type of FET-based photodetector is very stable because it was fabricated reversely using PMMA dielectric layer to shield the active layer from the environment, as well as with inorganic ligand-exchange treatment on the active layer. Therefore, it is promising to fabricate a high-sensitivity near-IR/mid-IR photodetector with good stability by this easy approach.

Acknowledgements

This project was partially funded by the Foundation of Distinguished Teacher at Beijing Institute of Technology (BIT) (BIT-JC-201005), the project of State Key Laboratory of Transducer Technology (SKT1404), the project of the Key Laboratory of Photoelectronic Imaging Technology and System (2015OEIOF02), Beijing Institute of Technology, Ministry of Education of China, and the Key Project of Chinese National Programs for Fundamental Research and Development (2013CB329202).

References

  1. E. H. Sargent, Infrared photovoltaics made by solution processing, Nat. Photonics, 2009, 3, 325–331 CrossRef CAS.
  2. M. H. Zarghami, Y. Liu, M. Gibbs, E. Gebremichael, C. Webster and M. Law, P-Type PbSe and PbS Quantum Dot Solids Prepared with Short-Chain Acids and Diacids, ACS Nano, 2010, 4, 2475–2485 CrossRef CAS PubMed.
  3. L. E. Bru, Electron-electron and electron–hole interactions in small semiconductor crystallites: the size dependence of the lowest excited electronic state, J. Chem. Phys., 1984, 80, 4403–4409 CrossRef.
  4. X. Peng, An essay on synthetic chemistry of colloidal nanocrystals, Nano Res., 2009, 2, 425–447 CrossRef CAS.
  5. Y. Shirasaki, G. J. Supran, M. G. Bawendi and V. Bulovic, Emergence of colloidal quantum-dot light-emitting technologies, Nat. Photonics, 2013, 7, 13–23 CrossRef CAS.
  6. G. Konstantatos, I. Howard, A. Fischer, S. Hoogland, J. Clifford, E. Klem, L. Levina and E. H. Sargent, Ultrasensitive solution-cast quantum dot photodetectors, Nature, 2006, 442, 180–183 CrossRef CAS PubMed.
  7. L. Sun, J. J. Choi, D. Stachnik, A. C. Bartnik, B. R. Hyun, G. G. Malliaras, T. Hanrath and F. W. Wise, Bright infrared quantum-dot light-emitting diodes through inter-dot spacing control, Nat. Nanotechnol., 2012, 7, 369–373 CrossRef CAS PubMed.
  8. F. Guo, B. Yang, Y. Yuan, Z. Xiao, Q. Dong, Y. Bi and J. Huang, A nanocomposite ultraviolet photodetector based on interfacial trap-controlled charge injection, Nat. Nanotechnol., 2012, 7, 798–802 CrossRef CAS PubMed.
  9. K. S. Yook and Y. J. Lee, Small molecule host materials for solution processed phosphorescent organic light-emitting diodes, Adv. Mater., 2014, 26, 4218–4233 CrossRef CAS PubMed.
  10. D. Zhitomirsky, M. Furukawa, J. Tang, P. Stadler, S. Hoogland, O. Voznyy, H. Liu and E. H. Sargent, N-type colloidal-quantum-dot solids for photovoltaics, Adv. Mater., 2012, 24, 6181–6185 CrossRef CAS PubMed.
  11. X. Wang, G. I. Koleilat, J. Tang, H. Liu, I. J. Kramer, R. Debnath, L. Brzozowski, D. Aaron, R. Barkhouse, L. Levina, S. Hoogland and E. H. Sargent, Tandem colloidal quantum dot solar cells employing a graded recombination layer, Nat. Photonics, 2011, 5, 480–484 CrossRef CAS.
  12. W. Ma, S. L. Swisher, T. Ewers, J. Engel, V. E. Ferry, H. A. Atwater and A. P. Alivisa, Photovoltaic performance of ultrasmall PbSe quantum dots, ACS Nano, 2011, 5, 8140–8147 CrossRef CAS PubMed.
  13. D. V. Talapin and C. B. Murray, PbSe nanocrystal solids for n- and p-channel thin film field-effect transistors, Science, 2005, 310, 86–89 CrossRef CAS PubMed.
  14. Y. Z. Jin, J. P. Wang, B. Q. Sun, J. C. Blakesley and N. C. Greenham, Solution-processed ultraviolet photodetectors based on colloidal ZnO nanoparticles, Nano Lett., 2008, 8, 1649–1653 CrossRef CAS PubMed.
  15. H. Chen, M. Lo, G. Yang, H. Monbouquette and Y. Yang, Nanoparticle-assisted high photoconductive gain in composites of polymer and fullerene, Nat. Nanotechnol., 2008, 3, 543–547 CrossRef CAS PubMed.
  16. G. Konstantatos and G. H. Sargent, Nanostructured materials for photon detection, Nat. Nanotechnol., 2010, 5, 391–400 CrossRef CAS PubMed.
  17. M. S. Arnold, J. D. Zimmerman, C. K. Renshaw, X. Xu, R. Lunt, C. M. Austin and S. R. Forrest, Broad spectral response using carbon nanotube/organic semiconductor/C60 photodetectors, Nano Lett., 2009, 9, 3354–3358 CrossRef CAS PubMed.
  18. L. Y. Chang, R. R. Lunt, P. R. Brown, V. Bulovic and M. G. Bawendi, Low-temperature solution-processed solar cells based on PbS colloidal quantum dot/CdS heterojunctions, Nano Lett., 2013, 13, 994–999 CrossRef CAS PubMed.
  19. P. Wei, J. H. Oh, G. F. Dong and Z. Bao, Use of a 1H-benzoimidazole derivative as an n-type dopant and to enable air-stable solution-processed n-channel organic thin-film transistors, J. Am. Chem. Soc., 2010, 132, 8852–8853 CrossRef CAS PubMed.
  20. P. D. Angelo, M. Barra, A. Cassinese, M. G. Maglione, P. Vacca, C. Minarini and A. Rubino, Electrical transport properties characterization of PVK (poly N-vinylcarbazole) for electroluminescent devices applications, Solid-State Electron., 2007, 51, 123–129 CrossRef.
  21. H. Sirringhaus, N. Tessler and R. H. Friend, Integrated optoelectronic devices based on conjugated polymers, Science, 1998, 280, 17741–17744 CrossRef.
  22. A. Dodabalapur, Z. Bao, A. Makhija, J. G. Laquindanum, V. R. Raju, Y. Feng, H. E. Katz and J. A. Rogers, Organic smart pixels, Appl. Phys. Lett., 1998, 73, 142–144 CrossRef CAS.
  23. H. W. Wang, Z. X. Li, C. J. Fu, D. Yang, L. Zhang, S. Y. Yang and B. S. Zou, Solution-Processed PbSe Colloidal Quantum Dot-Based Near-Infrared Photodetector, IEEE Photonics Technol. Lett., 2015, 27, 612–617 CrossRef CAS.
  24. M. B. Khalifa, D. Vaufrey, A. Bouazizi, J. Tardy and H. Maaref, Hole injection and transport in ITO/PEDOT/PVK/Al diodes, Mater. Sci. Eng., 2002, 21, 277–282 CrossRef.
  25. L. Qian, S. Y. Yang, Z. S. Jin, Z. J. Zhang, T. Zhang, F. Teng and X. R. Xu, Enhanced performance of light-emitting diodes based on a nanocomposite of dehydrated nanotube titanic acid and poly(vinylcarbazole) (PVK), Phys. Lett. A, 2005, 335, 56–60 CrossRef CAS.
  26. M. B. Khalifa, D. Vaufrey and J. Tardy, Opposing influence of hole blocking layer and a doped transport layer on the performance of heterostructure OLEDs, Org. Electron., 2004, 5, 187–198 CrossRef.
  27. P. Y. Lai and J. S. Chen, Influence of electrical field dependent depletion at metal–polymer junctions on resistive switching of poly(N-vinylcarbazole) (PVK)-based memory devices, Org. Electron., 2009, 10, 1590–1595 CrossRef CAS.
  28. W. W. Yu, J. C. Falkner, B. S. Shih and V. L. Colvin, Preparation and characterization of monodisperse PbSe semiconductor nanocrystals in a noncoordinating solvent, Chem. Mater., 2004, 16, 3318–3322 CrossRef CAS.
  29. C. J. Fu, H. W. Wang, T. J. Song, L. Zhang, W. L. Li, B. He, M. Sulaman, S. Y. Yang and B. S. Zou, Stability enhancement of PbSe quantum dots via post-synthetic ammonium chloride treatment for high-performance infrared photodetector, Nanotechnology, 2015, 27, 065201 CrossRef PubMed.
  30. T. P. Osedach, S. M. Geyer, J. C. Ho, A. C. Arango, M. G. Bawendi and V. Bulovic, Lateral heterojunction photodetector consisting of molecular organic and colloidal quantum dot thin films, Appl. Phys. Lett., 2009, 94, 043307 CrossRef.
  31. F. S. Manciu, Y. Sahoo, F. Carreto and P. N. Prasad, Size-dependent Raman and infrared studies of PbSe nanoparticles, J. Raman Spectrosc., 2008, 39, 1135–1140 CrossRef CAS.
  32. K. Ahmad, M. Afzaal, P. O'Brien, G. Hua and J. D. Woollins, Morphological Evolution of PbSe Crystals via the CVD Route, Chem. Mater., 2010, 22, 4619–4624 CrossRef CAS.
  33. A. Lipovskii, E. Kolobkova, V. Petrikov, I. Kang, A. Olkhovets, T. Krauss, M. Thomas, J. Silcox, F. Wise, Q. Shen and S. Kycia, Synthesis and characterization of PbSe quantum dots in phosphate glass, Appl. Phys. Lett., 1997, 71, 3406–3408 CrossRef CAS.
  34. O. M. Primera-Pedrozo, Z. Arslan, B. Rasulevb and J. Leszczynskib, Room temperature synthesis of PbSe quantum dots in aqueous solution: stabilization by interactions with ligands, Nanoscale, 2012, 4, 1312–1320 RSC.
  35. M. Sykora, A. Y. Koposov, J. A. McGuire, R. K. Schulze, O. Tretiak and J. M. Pietryga, Effect of air exposure on surface properties, electronic structure, and carrier relaxation in PbSe nanocrystals, ACS Nano, 2010, 4, 2021–2034 CrossRef CAS PubMed.
  36. M. Law, J. M. Luther, Q. Song, B. K. Hughes, C. L. Perkins and A. J. Nozik, Structural, optical, and electrical properties of PbSe nanocrystal solids treated thermally or with simple amines, J. Am. Chem. Soc., 2008, 130, 5974–5985 CrossRef CAS PubMed.
  37. J. M. Luther, M. Law, Q. Song, C. L. Perkins, M. C. Beard and A. J. Nozik, Structural, Optical, and Electrical Properties of Self-Assembled Films of PbSe Nanocrystals Treated with 1,2-Ethanedithiol, ACS Nano, 2008, 2, 271–280 CrossRef CAS PubMed.
  38. T. Kawazoe and M. Ohtsu, Increasing Si photodetector photosensitivity in near-infrared region and manifestation of optical amplification by dressed photons, Appl. Phys. B, 2012, 108, 51–56 CrossRef.
  39. M. Fontana, N. Matias, F. Fragoso, C. Aiese, C. Barros and A. Oliveira, Photosensitivity Characterization of Nanostructured Tin Oxide Films and Alternative Photodetector Application, IEEE Sens. J., 2011, 11, 869–874 CrossRef CAS.
  40. Y. Zhang, T. Liu, B. Meng, X. Li, G. Liang, X. Hu and Q. J. Wang, Broadband high photoresponse from pure monolayer graphene photodetector, Nat. Commun., 2013, 4, 1811–1815 CrossRef PubMed.
  41. S. Y. Yang, N. Zhao, L. Zhang, H. Zhong, R. B. Liu and B. S. Zou, Field-effect transistor-based solution-processed colloidal quantum dot photodetector with broad bandwidth into near-infrared region, Nanotechnology, 2012, 23, 255203 CrossRef PubMed.
  42. N. Cho, K. R. Choudhury, Y. Sahoo and K. S. Lee, High density coupling of quantum dots to carbon nanotube surface for efficient photodetection, US Pat., US8003979, 2007.
  43. M. M. Sheung, Y. Feng and H. L. W. Chan, Organic phototransistor based on poly(3-hexylthiophene)/TiO2 nanoparticle composite, Appl. Phys. Lett., 2008, 93, 023310–023313 CrossRef.
  44. H. C. Shin, I. S. Lim, M. Racanelli, W. L. M. Huang, J. Foerster and B. Y. Hwang, Analysis of floating body induced transient behaviors in partially depleted thin film SOI devices, IEEE Trans. Electron Devices, 1996, 43, 318–325 CrossRef.
  45. F. Yan, P. Migliorato and R. Ishihara, Influence of trap states on dynamic properties of single grain silicon thin film transistors, Appl. Phys. Lett., 2006, 88, 153507–153509 CrossRef.
  46. B. J. Chen and S. Y. Liu, Measurement of electron/hole mobility in organic/polymeric thin films using modified time-of-flight apparatus, Synth. Met., 1997, 91, 169–171 CrossRef CAS.
  47. M. F. Ahmed and I. A. Hümmelgen, Determination of electron and hole mobility in poly(3-hexylthiophene) using space-charge-limited-current measurements, International Journal of Electroactive Materials, 2013, 1, 60–63 Search PubMed.
  48. J. E. Kroeze, T. J. Savenije, M. J. W. Vermeulen and J. M. Warman, Contactless Determination of the Photoconductivity Action Spectrum, Exciton Diffusion Length, and Charge Separation Efficiency in Polythiophene-Sensitized TiO2 Bilayers, J. Phys. Chem. B, 2003, 107, 7696–7705 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2016