Chinese Physics Letters, 2023, Vol. 40, No. 3, Article code 038101 Large-Area Monolayer n-Type Molecular Semiconductors with Improved Thermal Stability and Charge Injection Sai Jiang1*, Lichao Peng1, Xiaosong Du1, Qinyong Dai2, Jianhang Guo2, Jianhui Gu1, Jian Su1, Ding Gu1, Qijing Wang2, Huafei Guo1, Jianhua Qiu1, and Yun Li2* Affiliations 1School of Microelectronics and Control Engineering, Changzhou University, Changzhou 213164, China 2National Laboratory of Solid-State Microstructures, School of Electronic Science and Engineering, Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China Received 14 December 2022; accepted manuscript online 14 February 2023; published online 8 March 2023 *Corresponding authors. Email: saijiang@cczu.edu.cn; yli@nju.edu.cn Citation Text: Jiang S, Peng L, Du X et al. 2023 Chin. Phys. Lett. 40 038101    Abstract We fabricated monolayer n-type two-dimensional crystalline semiconducting films with millimeter-sized areas and remarkable morphological uniformity using an antisolvent-confined spin-coating method. The antisolvent can cause a downstream Marangoni flow, which improves the film morphologies. The deposited crystalline monolayer films exhibit excellent thermal stabilities after annealing, which reveals the annealing-induced enhancement of crystallinity. The transistors based on the n-type monolayer crystalline films show linear output characteristics and superior electron mobilities. The improved charge injection between monolayer films and Au electrodes results from the energy level shift as the films decrease to the monolayer, which leads to a lower injection barrier. This work demonstrates a promising method for fabricating air-stable, low-cost, high-performance, and large-area organic electronics.
cpl-40-3-038101-fig1.png
cpl-40-3-038101-fig2.png
cpl-40-3-038101-fig3.png
cpl-40-3-038101-fig4.png
DOI:10.1088/0256-307X/40/3/038101 © 2023 Chinese Physics Society Article Text Organic crystalline semiconductors, which consist of conjugated molecules assembled via weak van der Waals (vdW) forces, have attracted increasing interest in low-cost large-area organic electronics and optoelectronics.[1-3] Among organic crystalline semiconductors, two-dimensional (2D) molecular films with high crystallinity long-range structural order and chemical purity are representative materials for high-performance organic field-effect transistors (OFETs) and high-sensitivity sensors, enabling in-depth research on film growth mechanisms, carrier transport dynamics, charge injection physics, and interfacial properties.[4-9] With continuous improvements in electrical performance of OFETs, more sophisticated applications, including organic complementary metal oxide semiconductor (CMOS) circuits and heterojunction devices, require both p-type and n-type semiconductors with solution processability, high mobility, and thermal stability.[10-12] Many p-type 2D organic semiconductors (OSCs), such as DTBDT-C9, C$_{8}$-BTBT, ph-BTBT-C10, and C$_{10}$-DNTT films, have recently been fabricated using dip-coating, bar-coating, solution-shearing, and the floating-coffee-ring-driven assembly methods, demonstrating inherent film crystallinity, high hole mobilities exceeding 10 cm$^{2}\cdot$V$^{-1}\cdot$s$^{-1}$, and excellent charge injection. These p-type 2D OSCs are fascinating for direct exploration of the relationship between charge accumulation and transport behaviors with semiconductor/insulator interface, energy disorder, and dielectric polarizations, which promoted the developments of the 2D organic semiconductors.[5,8,13-15] However, the development of high-quality n-type OSCs still lags behind their p-type counterparts due to the inferior stability under oxygen, moisture, and high-temperature exposure, hindering the study of interfacial transport physics in n-type semiconductors.[16,17] Therefore, developing an efficient strategy to fabricate high-quality n-type 2D OSCs in large areas with excellent thermal stability and electrical performance is desirable. The extended $\pi$-electron framework between molecules and energy levels of the lowest unoccupied molecular orbitals (LUMO) lower than $-$4.0 eV are characteristics of high-quality n-type 2D OSCs that can boost carrier transport and shield the material from oxidation by ambient O$_{2}$ and H$_{2}$O.[16,18] N,N$'$-1H,1H-perfluorobutyldicyanoperylene-carboxydi-imide (PDIF-CN$_{2}$) as the n-type OSC has demonstrated the single-crystal feature, thermal stability, and high electron mobility exceeding 1 cm$^{2}\cdot$V$^{-1}\cdot$s$^{-1}$ for solution-grown single-crystalline transistors, which will be a promising candidate for fabricating large-area thermally stable monolayer n-type molecular semiconductors.[19,20] Here, we successfully demonstrate the fabrication of monolayer PDIF-CN$_{2}$ films with millimeter-scale coverage and uniform morphologies using an antisolvent-confined spin-coating technique. The antisolvent not only speeds up the nucleation of the films but also induces a downstream Marangoni flow, improving the film morphologies. The monolayer PDIF-CN$_{2}$ films can tolerate annealing at a high temperature of 120 ℃, which enhances the crystallinity of the films. The OFETs with monolayer PDIF-CN$_{2}$ films exhibit average and maximum electron mobilities of 0.28 and 0.5 cm$^{2}\cdot$V$^{-1}\cdot$s$^{-1}$. Degradations induced by ambient O$_{2}$ and H$_{2}$O after prolonged exposure to air can be recovered through a post-annealing treatment. Furthermore, monolayer-based OFETs show linear output characteristics. The results of ultraviolet photoelectron spectroscopy (UPS) indicate that the reduction in film thickness causes a considerable change in energy levels, which lowers the injection barrier between the LUMO of monolayer PDIF-CN$_{2}$ films and Au electrodes. This work reveals the process of antisolvent-confined molecular assembly and the charge injection mechanism at the molecular level via stable and high-quality monolayer PDIF-CN$_{2}$ films, deepening the understanding of n-type organic semiconductors and advancing the development of organic electronics. Results and Discussion. PDIF-CN$_{2}$ is an n-type molecular semiconducting material with the $\pi$-conjugated core and N-fluorocarbon functionalization. PDIF-CN$_{2}$ films exhibit excellent electrical performance and thermal stability.[19,20] We first dissolved PDIF-CN$_{2}$ in the solvent of anisole with a concentration of 0.1 wt%. The solution (40 µL) was drop-cast onto a SiO$_{2}$/Si substrate. Then, the antisolvent of N,N-dimethylformamide (DMF) (40 µL) was further drop-cast onto the PDIF-CN$_{2}$ solution for the spin-coating process [Fig. 1(a) and see details in the Experimental Methods in the Supplementary Material]. During the spin-coating procedure, monolayer PDIF-CN$_{2}$ films with millimeter-sized coverage and good morphological uniformity can be deposited on the substrate as the solution was pulled by the centrifugal force [Fig. 1(c)]. Finally, the monolayer films were annealed at the temperature of 120 ℃ for 15 min in the N$_{2}$ glove box. The annealed films show no signs of film degradation and maintain their outstanding quality [Figs. 1(d) and 1(e)]. Atomic force microscopy (AFM) images of the monolayer films reveal that the film thicknesses are 1.1 and 1.5 nm before and after annealing, respectively. These are less than the $c$-axis unit cell parameter of PDIF-CN$_{2}$ single crystals (2.28 nm), validating the monolayer property of the depositing films [Figs. 2(a) and 2(c) and Fig. S1 in the Supplementary Material].[21] The smaller thickness of monolayer films results from the tilted molecular arrangement on the substrate, which is induced by the strong vdW interaction between the molecules and the substrate (Fig. S1). Additionally, the rms roughness of the AFM images before and after annealing is 2.6 and 2.0 Å, indicating the atomic smoothness of the monolayer PDIF-CN$_{2}$ films [Figs. 2(b) and 2(d)]. Furthermore, the crystallinity of the PDIF-CN$_{2}$ films before and after annealing was examined by the grazing incidence x-ray diffraction (GIXRD) (Fig. S2 and Note S1 in the Supplementary Material). The GIXRD results indicate that the PDIF-CN$_{2}$ films exhibit a crystalline property. As a result, the antisolvent-confined spin-coating method is efficient for the growth of high-quality monolayer PDIF-CN$_{2}$ films.
cpl-40-3-038101-fig1.png
Fig. 1. Large-area PDIF-CN$_{2}$ films using the antisolvent-confined spin-coating method. (a) Schematic diagram of the antisolvent-confined spin-coating deposition of PDIF-CN$_{2}$ thin films. For step 1, the solution of PDIF-CN$_{2}$ in anisole is drop-cast onto a SiO$_{2}$/Si substrate. For step 2, the upper solvent of DMF entirely covers the solvent of the anisole. For step 3, crystalline sheets assemble into monolayer films at the anisole/DMF interface during the spin-coating process. (b) Simulation of the Marangoni flow at the Meniscus line. (c) The optical microscopic image of the deposited large-area monolayer PDIF-CN$_{2}$ films. The inset is the molecular structure of PDIF-CN$_{2}$. Scale bar, 200 µm. (d) and (e) The optical microscopic images of monolayer PDIF-CN$_{2}$ films before and after annealing at 120 ℃. Scale bar, 100 µm (d), 50 µm (e).
The upper layer solvent of DMF strongly influences the morphology of the depositing monolayer films. Due to the immiscibility of anisole and DMF, the antisolvent of DMF shows vertical solvent separation with anisole.[22] The antisolvent can speed up the nucleation and aggregation of PDIF-CN$_{2}$ molecules at the interface of anisole and DMF, promoting the growth of monolayer crystalline sheets.[23] When the edge of the solution moves on the substrate by the centrifugal force during the spin-coating process, monolayer crystalline sheets can migrate at the anisole/DMF interface and assemble into large-area and high-quality monolayer films [Fig. 1(a)]. The fluid dynamics simulation was performed to further study the influence of the additive solvent of DMF on the film morphologies. When the solution was pulled by the centrifugal force, a meniscus line can be formed at the solution edge, controlling the solution flow and film deposition.[24] In particular, the Marangoni flow, which results from the surface-tension gradient, may act as a vital fluid process. The Marangoni flow causes the fluid to flow from regions of low surface tension to regions of high surface tension and attributes to the molecule migration along the meniscus line.[25] Improvement in film morphologies requires a downstream Marangoni flow to supply sufficient solutes to the meniscus line for monolayer film depositions using the spin-coating process. In the heterogeneous solvent system of DMF and anisole with vertical phase separation, DMF could be regarded as an additive solvent for the primary solvent of anisole. DMF can modify the surface tension gradient along the meniscus to adjust the Marangoni flow. As the solution was pulled by the centrifugal force, the upper solvent of DMF significantly would diffuse into the primary solvent of anisole, leading to a downward increase in the concentration gradient of DMF along the meniscus line (Fig. S3). Since the surface tension of DMF was larger than that of anisole (Table 1), a downstream Marangoni flow could be induced by the surface tension gradient, which would improve the film morphologies.[25] The CFD simulation using a confined geometry filled with 99% anisole and 1% additive DMF also demonstrates that DMF can induce Marangoni flow along the meniscus line in a downstream flow direction [Fig. 1(b), Fig. S3 and Note S2 in the Supplementary Material]. Therefore, the antisolvent-confined spin-coating method offers a downstream Marangoni flow to supply sufficient organic molecules to the contact line to achieve supersaturation, which yields highly uniform PDIF-CN$_{2}$ films.
Table 1. Properties of anisole and DMF.
Boiling point
(℃)
Density
(g/cm$^{3}$)
Surface tension
(mN/m)
Anisole 154 0.995 33.97
DMF 153 0.950 36.50
We further study the effects of spin-coating speed on the morphologies of the films. The spin-coating speeds are tuned from 1000 to 4000 rpm [Fig. 2(e)]. Figure S4 shows the optical images and the corresponding AFM images of PDIF-CN$_{2}$ films at different spin-coating speeds after annealing. When the spin coating process was performed at a low speed of 1000 rpm, discontinuous film morphologies were observed with the film coverage of only 61% on the substrate. The corresponding AFM images also exhibited defects and grain boundaries with a high rms roughness of 4.2 nm [Figs. S4(a) and S4(e)]. The morphological continuity of the films can be improved by raising the spin-coating speed to 2000 rpm. The film coverage and the surface roughness are 80% and 3.6 Å, respectively [Figs. S4(b) and S4(f)]. When the spin-coating process was at an optimized speed of 3000 rpm, we found that the substrate was almost entirely covered by a monolayer film with a low rms roughness of 2.1 Å, exhibiting atomic smoothness [Figs. S4(c) and S4(g)]. We further increased the spin-coating speed to 4000 rpm, and the films became discontinuous with obvious grain boundaries [Figs. S4(d) and S4(h)]. Therefore, the spin-coating speed of 3000 rpm was the optimized value to achieve a monolayer PDIF-CN$_{2}$ film with a large area and a uniform morphology. We believe that antisolvent-confined spin coating is a simple and efficient way to deposit high-quality monolayer molecular films. Note that thermal annealing can improve the crystalline properties of organic crystals. We then study the effect of thermal annealing on the crystal morphology of PDIF-CN$_{2}$ thin films at the monolayer limit by AFM and Raman spectroscopy characterizations. As shown in Figs. 2(a)–2(d), thermal annealing of monolayer PDIF-CN$_{2}$ films at 120 ℃ for 15 min in the N$_{2}$ glove box can increase the film thickness from 1.1 nm to 1.5 nm and reduce the rms roughness of films from 2.6 Å to 2.0 Å. The alterations in molecular packing show that thermal annealing can effectively reorganize the molecules to have more consistent morphologies and orient them upright [Figs. 2(c) and 2(d)]. The characteristic signals of the Raman spectrum of monolayer PDIF-CN$_{2}$ films can be measured at 1264 cm$^{-1}$, 1360 cm$^{-1}$, 1460 cm$^{-1}$, 1560 cm$^{-1}$, 1601 cm$^{-1}$, and 1701 cm$^{-1}$, which are close to the Raman spectrum of PDI [Fig. S5(b)].[26,27] Notably, the Raman peak intensity of the monolayer film at 1360 cm$^{-1}$ after annealing is three times higher than its pre-annealing value, and the full width at half maximum (FWHM) of Raman peaks is smaller than the pre-annealing films, both supporting the significant improvement of the film crystallinity and reduction of defects by annealing [Fig. 2(f)].[28] The GIXRD results also indicate that thermal annealing can effectively improve the crystallinity (Fig. S2). The insets in Figs. 2(b) and 2(d) show schematic depictions of molecular arrangement before and after annealing, illuminating how annealing may promote molecular reorganization and enhance the crystallinity of monolayer films. In addition, thermal annealing can also eliminate residual solvent, oxygen, and moisture in the films, improving film quality and stability. OFETs were fabricated utilizing monolayer annealed PDIF-CN$_{2}$ films with bottom-gate top-contact structures to examine the electrical properties [inset in Fig. 3(a) and Fig. S6]. Au pattern stripes were transferred onto PDIF-CN$_{2}$ films as the source and drain electrodes, which feature a physical contact interface between the semiconducting films and electrodes.[14,29] Figure 3(a) shows typical transfer characteristics of monolayer-based transistors with a pristine silicon oxide dielectric layer, yielding average and maximum field-effect mobilities ($\mu_{_{\scriptstyle \rm FET}}$) of 0.28 and 0.50 cm$^{2}\cdot$V$^{-1}\cdot$s$^{-1}$, respectively [Fig. 3(b)]. A slight decrease in carrier mobility can be observed at high gate voltage ($V_{\rm G}$), which could result from the interface polarization coupling of the dielectric or Coulomb interactions between electrons in the monolayer films (Fig. S6).[30,31] Furthermore, the output characteristics exhibit a linear increase at small $V_{\rm DS}$ from the gate voltage of 0 V to 30 V [Figs. 3(c) and 3(d)]. Near-zero threshold voltage ($V_{\rm TH}$) and negligible hysteresis of transfer curves are demonstrated at $V_{\rm DS}$ of 20 V, indicating that the interface between PDIF-CN$_{2}$ films and dielectric of SiO$_{2}$ has an ultralow density of trap states [Fig. 3(a)]. It is also shown that the $V_{\rm TH}$ is nearly unchanged after 15-cycle transfer scanning, further revealing the outstanding electrical stability of the monolayer-based transistor.[32] The mobilities of monolayer-based transistors are lower than those of transistors with few-layer crystalline PDIF-CN$_{2}$ films.[19] The absence of surface treatments is an important reason for the reduction of mobility in monolayer thin films. In addition, the molecules of monolayer (1 L) films are more tilted on the substrate than 2 L molecules due to the stronger molecule–substrate interactions, which lead to weaker $\pi$–$\pi$ stacking between molecules, resulting in the reduction of mobility in monolayer films [Fig. S1(b)].
cpl-40-3-038101-fig2.png
Fig. 2. The effect of the spin-coating speed and thermal annealing on the crystallinity of PDIF-CN$_{2}$ films. (a) and (b) AFM images of monolayer PDIF-CN$_{2}$ films before annealing, with scale bar 2 µm. The inset in (b) is the schematic diagram of molecular packing of PDIF-CN$_{2}$ films on SiO$_{2}$ before annealing. (c) and (d) AFM images of monolayer PDIF-CN$_{2}$ films after annealing, with scale bar 2 µm. The inset in (d) is the schematic diagram of molecular packing of PDIF-CN$_{2}$ films on SiO$_{2}$ after annealing. (e) Coverage and rms roughness of spin-coated monolayer PDIF-CN$_{2}$ films on SiO$_{2}$/Si substrate at various speeds of spin coating. (f) Raman spectrum of monolayer PDIF-CN$_{2}$ films on SiO$_{2}$ before and after annealing.
cpl-40-3-038101-fig3.png
Fig. 3. Electrical performance of transistors based on monolayer PDIF-CN$_{2}$ films. (a) Transfer characteristics at a drain voltage of 20 V of monolayer-based transistors with 15-cycle measurements. The inset is a typical diagram of a monolayer-based transistor. (b) The electron mobility histogram of monolayer-based OFETs. The green line is the Gaussian fitting curve. (c) Output characteristics at gate voltage from 0 to 30 V of the monolayer-based transistor. (d) Linear output curves at small $V_{\rm DS}$ corresponding to output characteristic of (c).
The air stability of the monolayer-based devices is also investigated. As shown in Fig. S7(a), the carrier mobility of monolayer-based OFETs is stable, ranging from 0.32 to 0.20 cm$^{2}\cdot$V$^{-1}\cdot$s$^{-1}$ under a vacuum environment for 300 h. Furthermore, the effects of the aqueous solution on PDIF-CN$_{2}$ films are studied. We submerged the sample in deionized water for 10 s and then dried it by airflow. Obvious mobility reduction from 0.20 to 0.025 cm$^{2}\cdot$V$^{-1}\cdot$s$^{-1}$ can be observed. For the degraded device, a post-annealing treatment at 120 ℃ for 20 min under vacuum conditions could restore the electrical performance to approach the origin mobility values of 0.14 cm$^{2}\cdot$V$^{-1}\cdot$s$^{-1}$ [Fig. S7(b)]. Therefore, the high-temperature annealing can effectively eliminate attached water and oxygen molecules, contributing to the performance recovery of destroyed films. The air stability of the transistors with multilayer PDIF-CN$_{2}$ films ($\sim$ 3 L) was also investigated. The multilayer films were deposited by the inclined drop-cast method (Note S3 in the Supplementary Material). The carrier mobility of multilayer-based OFETs is more stable, ranging from 0.48 to 0.39 cm$^{2}\cdot$V$^{-1}\cdot$s$^{-1}$, than that of 1 L-based ones (Fig. S7). Since the organic molecules in the monolayer film would directly interact with the water and oxygen in the air as well as interface polarization of the dielectric layer, the monolayer films would be more susceptible to damage than thicker films. However, the monolayer PDIF-CN$_{2}$ films still exhibit good stability among the reported n-type organic monolayer semiconductors.[10,33] It is noted that the linear output characteristic is a key feature of transistors with monolayer PDIF-CN$_{2}$ films, which may signalize outstanding charge injection properties between the Au electrode and the monolayer annealed PDIF-CN$_{2}$ film [Figs. 3(c) and 3(d)]. To further study the carrier injection properties of the Au/PDIF-CN$_{2}$ interface, UPS characterizations were performed to examine the energy levels of the monolayer (1 L), $\sim$ bilayer (2 L), and $\sim$ trilayer (3 L) PDIF-CN$_{2}$ films (Figs. S8 and S9, and Note S3 in the Supplementary Material).[34] Figure 4(a) shows the UPS spectra of the PDIF-CN$_{2}$ films. The HOMO energy ($E_{\rm HOMO}$) can be obtained by[35,36] \begin{align} E_{\rm HOMO}=h\nu-(E_{\rm cutoff}-E_{\rm onset}), \tag {1} \end{align} where $h\nu$ is the incident photon energy of 21.2 eV, $E_{\rm cutoff}$ is the high binding energy cutoff, which is 15.97, 15.93, and 15.90 eV for 1 L, $\sim$ $2$ L, and $\sim$ $3$ L films, respectively, $E_{\rm onset}$ is the binding energy onset of the PDIF-CN$_{2}$ film relative to the $E_{\rm F}$, and $E_{\rm onset}$ takes 2.01, 1.60, and 1.55 eV for 1 L, $\sim$  $2$ L, and $\sim$ $3$ L films, respectively [Fig. 4(a)]. As a result, according to Eq. (1), the $E_{\rm HOMO}$ of the monolayer PDIF-CN$_{2}$ film is 7.24 eV, larger than that of $\sim$ $2$ L (6.87 eV) and $\sim$ $3$ L films (6.85 eV). In addition, the energy bandgap ($E_{\rm g}$) of PDIF-CN$_{2}$ films on the SiO$_{2}$ substrate can be analyzed by photoluminescence spectroscopy (PL).[37] The results show that $E_{\rm g}$ of 1 L, $\sim$  $2$ L, and $\sim$ $3$ L PDIF-CN$_{2}$ films takes 2.10, 2.07, and 2.06 eV, respectively, which are consistent with the value of 2.10 eV of spin-coated PDIF-CN$_{2}$ films [Fig. 4(b)].[21,38] Finally, the LUMO levels are calculated using the HOMO levels and $E_{\rm g}$, which are 5.14, 4.80, and 4.79 eV for 1 L, $\sim$ $2$ L, and $\sim$  $3$ L films, respectively [Fig. 4(c)].
cpl-40-3-038101-fig4.png
Fig. 4. Characteristics of the energy levels of PDIF-CN$_{2}$ films. (a) UPS spectra near the cutoff for secondary electron emission (left) and near the HOMO edge (right) for 1 L, $\sim$ $2$ L, and $\sim$ $3$ L PDIF-CN$_{2}$ films. (b) The photoluminescence spectra for 1 L, $\sim$ $2$ L, and $\sim$ $3$ L PDIF-CN$_{2}$ films. (c) The diagram of energy levels for 1 L, $\sim$ $2$ L, and $\sim$ $3$ L PDIF-CN$_{2}$ films as well as the work function of Au.
Remarkable energy level shifts can be found when the films decrease to the monolayer limit. A small Schottky barrier of 0.04 eV exists between the Au electrode and the LUMO of monolayer PDIF-CN$_{2}$ films, while a large Schottky barrier of 0.31 eV is found between Au and $\sim$  $3$ L films. When the film thickness is varied from $\sim$ $3$ L to $\sim$ $2$ L, the energy levels are almost unchanged. Hence, the decrease of the film thickness is not the reason for the energy level shifts. Such an abrupt transition for monolayer films may be attributed to strong modulation of the molecular packing by interfacial vdW interactions. The molecules of 1 L films are more tilted on the substrate than 2 L molecules due to the stronger molecule–substrate interactions, which leads to weaker $\pi$–$\pi$ stacking between molecules in monolayer films.[5] Different molecular packing has important influences on the molecular orbitals in 1 L and 2 L films, which could result in energy level shifts.[5,14] The low injection Schottky barrier can enhance the carrier injection, leading to linear output characteristics. Additionally, a nonlinear increase of output curves can be found for transistors with multilayer PDIF-CN$_{2}$ films, which is consistent with the larger Schottky barrier of 0.31 eV for charge injections (Fig. S10). Therefore, the high-quality monolayer PDIF-CN$_{2}$ films not only enhance the electrical performance of OFETs but also offer unique physical characteristics of energy level changes for improved charge injections, which have great potential for high-performance heterojunction devices. In summary, we have successfully demonstrated an antisolvent-confined spin-coating method for depositing large-area and high-quality n-type monolayer molecular films, which shows excellent thermal stability at an annealing temperature of 120 ℃. Furthermore, the monolayer-based OFETs exhibit linear output characteristics and stabilities after 15 days of vacuum storage. Particularly, we find that the reduction in film thickness leads to a significant shift in energy levels, which lowers the injection barrier between the LUMO of monolayer PDIF-CN$_{2}$ films and Au electrodes. Therefore, this work can help to develop low-cost, large-area, and high-performance n-type monolayer molecular films, which are important for more complicated devices, such as heterojunctions and organic circuits. Acknowledgements. This work was supported by the National Natural Science Foundation of China (Grant No. 62206030), the Natural Science Foundation of Jiangsu (Grant Nos. BK20220624 and BK20220620), the Scientific Research Foundation of Jiangsu Provincial Education Department (Grant No. 21KJB510010), the Changzhou Sci & Tech Program (Grant No. CJ20220085), and the Leading Innovative Talents Introduction and Cultivation Project of Changzhou (Grant No. CQ20210084).
References Organic semiconductor crystalsVertical organic synapse expandable to 3D crossbar arrayOrganic electrochemical transistorsStable organic thin-film transistorsProbing Carrier Transport and Structure-Property Relationship of Highly Ordered Organic Semiconductors at the Two-Dimensional LimitIn-situ/operando characterization techniques for organic semiconductors and devicesElectronic functionalization of the surface of organic semiconductors with self-assembled monolayersControllable Growth and Field-Effect Property of Monolayer to Multilayer Microstripes of an Organic SemiconductorWeak epitaxy growth of organic semiconductor thin filmsBottom-up growth of n-type monolayer molecular crystals on polymeric substrate for optoelectronic device applicationsMixed Conduction in an N‐Type Organic Semiconductor in the Absence of Hydrophilic Side‐ChainsCharge-Trapping-Induced Non-Ideal Behaviors in Organic Field-Effect TransistorsFew‐Layer Organic Crystalline van der Waals Heterojunctions for Ultrafast UV PhototransistorsUltrahigh mobility and efficient charge injection in monolayer organic thin-film transistors on boron nitrideRetina‐Inspired Self‐Powered Artificial Optoelectronic Synapses with Selective Detection in Organic Asymmetric HeterojunctionsRobust, high-performance n-type organic semiconductorsN-Type 2D Organic Single Crystals for High-Performance Organic Field-Effect Transistors and Near-Infrared Phototransistorsn -Channel Semiconductor Materials Design for Organic Complementary CircuitsHigh Electron Mobility in Air for N,N′-1H,1H-Perfluorobutyldicyanoperylene Carboxydi-imide Solution-Crystallized Thin-Film Transistors on Hydrophobic SurfacesHigh-Performance Phototransistors Based on PDIF-CN2 Solution-Processed Single Fiber and Multifiber AssemblySupramolecular Order of Solution-Processed Perylenediimide Thin Films: High-Performance Small-Channel n-Type Organic TransistorsSpin-Coated Crystalline Molecular Monolayers for Performance Enhancement in Organic Field-Effect TransistorsA General Method for Growing Two-Dimensional Crystals of Organic Semiconductors by “Solution Epitaxy”The meniscus-guided deposition of semiconducting polymersSolutal‐Marangoni‐Flow‐Mediated Growth of Patterned Highly Crystalline Organic Semiconductor Thin Film Via Gap‐Controlled Bar CoatingNoncovalent Functionalization of Black PhosphorusVibrational and Electronic Properties of Perylenediimide Linked to Fullerene and TetrathiafulvaleneAFM-IR and IR-SNOM for the Characterization of Small Molecule Organic SemiconductorsApproaching the Schottky–Mott limit in van der Waals metal–semiconductor junctionsTunable Fröhlich polarons in organic single-crystal transistorsProbing Coulomb Interactions on Charge Transport in Few‐Layer Organic Crystalline Semiconductors by the Gated van der Pauw MethodSolution-Processed Crystalline n-Type Organic Transistors Stable against Electrical Stress and PhotooxidationN-Type Self-Assembled Monolayer Field-Effect Transistors and Complementary InvertersAnalysis of External and Internal Disorder to Understand Band‐Like Transport in n‐Type Organic SemiconductorsEnergy level alignment at interfaces of organic semiconductor heterostructuresIn-Depth Investigation of the Correlation between Organic Semiconductor Orientation and Energy-Level Alignment Using In Situ Photoelectron SpectroscopyBand gap characterization and photoluminescence properties of SiC nanowiresN-Channel Zinc Oxide Nanowire:Perylene Diimide Blend Organic Thin Film Transistors
[1] Wang C L, Dong H L, Jiang L, and Hu W P 2018 Chem. Soc. Rev. 47 422
[2] Choi Y, Oh S, Qian C, Park J H, and Cho J H 2020 Nat. Commun. 11 4595
[3] Rivnay J, Inal S, Salleo A, Owens R M, Berggren M, and Malliaras G G 2018 Nat. Rev. Mater. 3 17086
[4] Jia X J, Fuentes-Hernandez C, Wang C Y, Park Y, and Kippelen B 2018 Sci. Adv. 4 eaao1705
[5] Zhang Y H, Qiao J S, Gao S, Hu F R, He D W, Wu B, Yang Z Y, Xu B C, Li Y, Shi Y, Ji W, Wang P, Wang X, Xiao M, Xu H, Xu J B, and Wang X 2016 Phys. Rev. Lett. 116 016602
[6] Jiang S, Dai Q, Guo J, and Li Y 2022 J. Semicond. 43 041101
[7] Calhoun M F, Sanchez J, Olaya D, Gershenson M E, and Podzorov V 2008 Nat. Mater. 7 84
[8] Li L G, Gao P, Schuermann K C, Ostendorp S, Wang W C, Du C, Lei Y, Fuchs H, De Cola L, Müllen K, and Chi L 2010 J. Am. Chem. Soc. 132 8807
[9] Yang J L and Yan D H 2009 Chem. Soc. Rev. 38 2634
[10] Shi Y J, Jiang L, Liu J, Tu Z, Hu Y, Wu Q, Yi Y, Gann E, McNeill C R, Li H, Hu W, Zhu D, and Sirringhaus H 2018 Nat. Commun. 9 2933
[11] Surgailis J, Savva A, Druet V, Paulsen B D, Wu R, Hamidi-Sakr A, Ohayon D, Nikiforidis G, Chen X, McCulloch I, Rivnay J, and Inal S 2021 Adv. Funct. Mater. 31 2010165
[12] Un H I, Cheng P, Lei T, Yang C Y, Wang J Y, and Pei J 2018 Adv. Mater. 30 1800017
[13] Guo J H, Jiang S, Pei M J, Xiao Y L, Zhang B W, Wang Q J, Zhu Y, Wang H Y, Jie J S, Wang X R, Shi Y, and Li Y 2020 Adv. Electron. Mater. 6 2000062
[14] He D W, Qiao J S, Zhang L L, Wang J Y, Lan T, Qian J, Li Y, Shi Y, Chai Y, Lan W, Ono L K, Qi Y B, Xu J B, Ji W, and Wang X R 2017 Sci. Adv. 3 e1701186
[15] Hao Z Q, Wang H Y, Jiang S, Qian J, Xu X, Li Y T, Pei M, Zhang B, Guo J, Zhao H, Chen J, Tong Y, Wang J, Wang X, Shi Y, and Li Y 2022 Adv. Sci. 9 2103494
[16] Okamoto T, Kumagai S, Fukuzaki E, Ishii H, Watanabe G, Niitsu N, Annaka T, Yamagishi M, Tani Y, Sugiura H, Watanabe T, Watanabe S, and Takeya J 2020 Sci. Adv. 6 eaaz0632
[17] Wang C, Ren X C, Xu C H, Fu B, Wang R, Zhang X, Li R, Li H, Dong H, Zhen Y, Lei S, Jiang L, and Hu W 2018 Adv. Mater. 30 1706260
[18] Usta H, Facchetti A, and Marks T J 2011 Acc. Chem. Res. 44 501
[19] Soeda J, Uemura T, Mizuno Y, Nakao A, Nakazawa Y, Facchetti A, and Takeya J 2011 Adv. Mater. 23 3681
[20] Rekab W, Stoeckel M A, Gemayel M E, Gobbi M, Orgiu E, and Samorì P 2016 ACS Appl. Mater. & Interfaces 8 9829
[21] Fabiano S, Wang H, Piliego C, Jaye C, Fischer D A, Chen Z, Pignataro B, Facchetti A, Loo Y L, and Loi M A 2011 Adv. Funct. Mater. 21 4479
[22] Wang Q J, Juarez-Perez E J, Jiang S, Qiu L, Ono L K, Sasaki T, Wang X, Shi Y, Zheng Y, Qi Y, and Li Y 2018 J. Phys. Chem. Lett. 9 1318
[23] Xu C H, He P, Liu J, Cui A J, Dong H L, Zhen Y G, Chen W, and Hu W P 2016 Angew. Chem. Int. Ed. 55 9519
[24] Gu X D, Shaw L, Gu K, Toney M F, and Bao Z N 2018 Nat. Commun. 9 534
[25] Lee S B, Lee S, Kim D G, Kim S H, Kang B, and Cho K 2021 Adv. Funct. Mater. 31 2100196
[26] Abellán G, Lloret V, Mundloch U, Marcia M, Neiss C, Görling A, Varela M, Hauke F, and Hirsch A 2016 Angew. Chem. 128 14777
[27] Łapiński A, Graja A, Olejniczak I, Bogucki A, Połomska M, Baffreau J, Perrin L, Leroy-Lhez S, and Hudhomme P 2006 Mol. Cryst. Liq. Cryst. 447 87
[28] Rao V J, Matthiesen M, Goetz K P, Huck C, Yim C, Siris R, Han J, Hahn S, Bunz U H F, Dreuw A, Duesberg G S, Pucci A, and Zaumseil J 2020 J. Phys. Chem. C 124 5331
[29] Liu Y, Guo J, Zhu E, Liao L, Lee S J, Ding M, Shakir I, Gambin V, Huang Y, and Duan X 2018 Nature 557 696
[30] Hulea I N, Fratini S, Xie H, Mulder C L, Iossad N N, Rastelli G, Ciuchi S, and Morpurgo A F 2006 Nat. Mater. 5 982
[31] Jiang S, Qian J, Wang Q, Duan Y, Guo J, Zhang B, Sun H, Wang X, Liu C, Shi Y, and Li Y 2020 Adv. Electron. Mater. 6 2000136
[32] Yi H T, Chen Z H, Facchetti A, and Podzorov V 2016 Adv. Funct. Mater. 26 2365
[33] Ringk A, Li X, Gholamrezaie F, Smits E C P, Neuhold A, Moser A, Van Der Marel C, Gelinck G H, Resel R, De Leeuw D M, and Strohriegl P 2013 Adv. Funct. Mater. 23 2016
[34] Stoeckel M, Olivier Y, Gobbi M, Dudenko D, Lemaur V, Zbiri M, Guilbert A A Y, D'Avino G, Liscio F, Migliori A, Ortolani L, Demitri N, Jin X, Jeong Y, Liscio A, Nardi M, Pasquali L, Razzari L, Beljonne D, Samorì P, and Orgiu E 2021 Adv. Mater. 33 2007870
[35] Hill I G and Kahn A 1998 J. Appl. Phys. 84 5583
[36] Yun D J, Yun Y, Lee J, Kim J Y, Chung J G, Kim S H, Kim Y S, Heo S, Park J I, Kim K H, Kwon Y N, and Chung J W 2020 ACS Appl. Mater. & Interfaces 12 50628
[37] Chen J J, Tang W H, Xin L P, and Shi Q 2011 Appl. Phys. A 102 213
[38] Chen S P, Chen Y S, and Hsieh G W 2017 IEEE J. Electron Devices Soc. 5 367