Hiroshi
Imahori
*abc and
Midori
Akiyama
a
aDepartment of Molecular Engineering, Graduate School of Engineering, Kyoto University, Nishikyo-ku, Kyoto, 615-8510, Japan. E-mail: imahori@scl.kyoto-u.ac.jp
bInstitute for Integrated Cell-Material Sciences (WPI-iCeMS), Kyoto University, Sakyo-ku, Kyoto 606-8501, Japan
cInstitute for Liberal Arts and Sciences (ILAS), Kyoto University, Sakyo-ku, Kyoto 606-8501, Japan
First published on 3rd October 2024
p–n heterojunctions are fundamental components for electronics and optoelectronics, including diodes, transistors, sensors, and solar cells. Over the past few decades, organic–inorganic p–n heterojunctions have garnered significant interest due to the diverse properties they exhibit, which are a result of the limitless combinations of organic molecules and inorganic materials. This review article concentrates on photoinduced charge separation and photocurrent generation at heterojunctions between two-dimensional layered materials and structurally well-defined organic small molecules. We highlight representative examples, including our work, and critically discuss their potential and perspectives.
Wider impactp–n heterojuctions serve as crucial interfaces for electronics and optoelectronics. However, our understanding of these heterojunctions varies based on whether the materials involved are organic or inorganic, as viewed from a material science perspective. This review aritcle focuses on photoinduced charge separation occurring at hybrid heterojunctions, which are formed between inorganic two-dimensional layered materials and structurally well-defined organic small molecules. In order to gain a deeper understanding of the chemistry and physics at these heterojunctions for potential applications, it is necessary to develop well-defined interfaces that consist of monolayer two-dimensional semiconductors and small donor or acceptor molecules. |
In this context, two-dimensional (2D) layered materials have emerged as fascinating inorganic candidates for organic–inorganic p–n heterojunctions.7–10 For instance, transition metal dichalcogenides (TMDs) have diverse properties, including a wide range of electronic band structures that depend on crystal symmetry and stoichiometric identity. The metal cations are sandwiched between chalcogenide anions, forming a three-atom thick layer. Notably, the chalcogens do not exhibit high reactivity due to chemical saturation. These three-layer thick structures are stacked through multiple weak van der Waals forces.11 Accordingly, they can be separated easily using a variety of exfoliation techniques. In addition, TMDs possess different lattice structures, impacting their electronic character (e.g., 1T (metallic) and 2H (semiconductor) phases).12,13 Semiconducting molybdenum (Mo) and tungsten (W) TMDs, which have band gap energies that span the visible and near-infrared regions of the spectrum, are highly attractive for optoelectronic applications. Particularly, when transitioning from the bulk material, which is indirect semiconductor, to the corresponding few-layered structure, a shift to a direct band gap occurs. This is followed by a dramatic change in the bandgap, as well as enhanced absorption and photoluminescence. Thus, few layered TMDs, including MoS2 (bandgap of the monolayer (ML): 1.8 eV), WS2 (2.0 eV), MoSe2 (1.5 eV), WSe2 (1.7 eV), MoTe2 (1.1 eV), and WTe2 (1.1 eV), are promising as a platform for unprecedented electronic and optoelectronic function.14–16 Meanwhile, other few-layered single elemental 2D semiconducting materials exist, including phosphorene (P), antimonene (Sb), and bismuthene (Bi). However, these materials have been studied less extensively due to the difficulty in handling the materials.17–19
In this review article, our focus is on photoinduced CS and the resultant photocurrent generation at heterojunction between 2D layered materials and structurally well-defined organic small molecules. We highlight representative examples, including our works, and critically discuss their potentials and prospects.
Approach | Noncovalent | Covalent | |
---|---|---|---|
Flexible bridge | Rigid bridge | ||
Deposition | CVD, spin-coating, electrophoretic deposition | Edge functionalization | Surface functionalization |
State | Amorphous | Multi-conformation | Well-defined conformation |
Preparation | Facile and practical | Time-consuming and tedious | |
Analysis | Difficult | Difficult | Facile |
Photoinduced CS has been observed at noncovalent heterojunctions between organic small molecules and few-layer TMDs using time-resolved transient absorption (TA) and photoemission spectroscopies.20–25 Owing to the facile preparation of high-quality few-layer TMDs, most research has focused on semiconducting MoS2 and WS2. For instance, Hersam and coworkers reported ultrafast exciton dissociation and relatively long-lived CS in a heterojunction between pentacene and MoS2.21 In the experiment, a homogeneous ML-MoS2 film was grown directly on quartz wafers by CVD. Subsequently, a 30 nm thick pentacene film was deposited on the ML-MoS2. The TA spectra of the heterojunction were recorded at an excitation wavelength of 535 nm, where the excitation ratio of MoS2:pentance is 1:3. The four time constants for the decay of the MoS2 B-exciton, monitored at 612 nm, were attributed to the trapping of electrons and holes to surface defects26–28 (670 fs, 48%), hole transfer from photoexcited n-type MoS2 to p-type pentacene (6.7 ps, 28%), radiative recombination and electron trapping26,29–32 (431 ps, 9%), and charge recombination (CR) (5.1 ns, 15%) (Fig. 1 and Table 2). Interestingly, other heterojunctions also reveal similar CT (1–10 ps) and CR (1–10 ns).20–25
Hetero-junction | τ 1 (A1) | τ 2 (A2) | τ 3 (A3) | τ 4 (A4) | τ 5 (A5) | τ 6 | Ref. |
---|---|---|---|---|---|---|---|
a The quantities in parenthesis are the fractional amplitudes of each component. b Poly[[4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b′]dithiophene-2,6-diyl][3-fluoro-2-[(2-ethylhexyl)carbonyl]-thieno[3,4-b]thiophenediyl]]. | |||||||
Process | Carrier trapping | CT formation | Non-radiative decay | Radiative decay and e− trapping | CR | CD | |
ML-MoS2/pentacene | 670 fs (48%) | 6.7 ps (28%) | — | 431 ps (9%) | 5.1 ns (15%) | — | 21 |
ML-MoS2 | 670 fs (47%) | — | 16 ps (35%) | 431 ps (18%) | — | — | 21 |
ML-MoS2/PTB7b | — | 1–5 ps (–) | — | — | 3–4 ns | — | 23 |
ML-MoS2/ZnPc | — | 40 fs (∼100%), 120 fs | — | — | 100–1000 ps | 10 ps | 25 |
Bulk-MoS2/ZnPc | — | Limited ED and CT | — | — | T1 formation & CR (1–100 ps) | — | 22 |
ML-WS2/tetracene | — | 2–3 ps | — | — | 2 ns | — | 24 |
MoS2/P3HT:[60]PCBM | — | <100 fs (from MoS2 to P3HT) | — | — | — | — | 20 |
The overall hole transfer yield (charge-transfer (CT) efficiency) was estimated to be approximately 50%, which is limited by the fast trapping process. Therefore, trap sites caused by defects in MoS2 must be minimized during preparation. The CR lifetime is also shorter than that in polymer/fullerene bulk heterojunction solar cells (20 ns to 10 μs),33 but it is longer than the lifetime of indirect excitons in planar p–n 2D–2D heterojunctions.34 This fast CR eventually limits the ability of charge carriers to diffuse to respective electrodes, thereby limiting charge collection (CC) efficiency in a solar cell. Since ML-MoS2 exhibits fast carrier trapping occurring within less than 1 ps, high light-harvesting (LH) properties of organic small molecules with a long-lived S1 state relative to MoS2 are desirable for efficient CS. Namely, exclusive LH by organic molecules leads to better light energy conversion (vide infra).
Chan and coworkers presented photoinduced charge generation at the zinc phthalocyanine (ZnPc)/MoS2 interface.22,25 They selectively excited ZnPc at 700 nm to probe the formation and decay of exciton and CT states. It was found that a large band bending in ZnPc on bulk-MoS2 limits the extent of exciton delocalization and inhibits electron injection to the conduction band (CB) of bulk-MoS2, whereas efficient exciton diffusion (ED) and CT occurs at the ML-MoS2/ZnPc interface with the small band bending in ZnPc (Fig. 2 and Table 2). Namely, a hot CT state is formed with a high efficiency (∼100%) and a lifetime of 40 fs, which is transformed into a relaxed, localized CT state with a lifetime of 120 fs. They also proposed that large band bending in ZnPc on bulk MoS2 traps the hole of the CT excited state. The hole trapping with the faster spin conversion rate in bulk-MoS2 results in the T1 formation and subsequent CR from bulk-MoS2 to ZnPc. In contrast, the time for the dissociation of the CT excited state at the ML-MoS2/ZnP is relatively fast (∼10 ps), which is explained by a large number of loosely bound CT excited states with small binding energies and large spatial delocalization. In other words, ML-MoS2 is better than bulk-MoS2 in terms of ED, CT, CD, and CC. Therefore, a combination of ML-2D materials and highly LH organic molecules is promising for optoelectronics applications.
Fig. 2 Schematic of (a) charge-transfer (CT) and (b) charge dissociation (CD) in the ML-MoS2/ZnPc heterojunction with small band bending in ZnPc. |
For photoelectrochemical devices with a macroscopic size of working electrodes (≥0.4 cm2) under the three electrodes configuration, there are limited examples of heterojunctions comprised of 2D layered materials and organic small molecules. In this regard, hybrids of fullerenes and 2D materials have attracted attention for the creation of novel functional materials through hybridization.35,36 Imahori and coworkers introduced composite films composed of C60 and well-exfoliated nanosheets of TMD, specifically MoS2 or WS2, obtained from the corresponding bulk materials.37 These films were conveniently fabricated onto a semiconducting SnO2 electrode using a two-step method. The first step involved the formation of aggregates by injecting a poor solvent (acetonitrile) into a good solvent (N-methyl-2-pyrrolidone (NMP)) that contained C60 and either MoS2 or WS2. The second step was the electrophoretic deposition, which was achieved by applying a direct current voltage to the aggregate solution.
The composite structures were meticulously characterized by dynamic light scattering and microscopic measurements, confirming the formation of a heterojunction between C60 and either MoS2 or WS2, with an average thickness of three layers. Upon photoexcitation of the composites on the SnO2 semiconducting electrode, they displayed significantly enhanced flash-photolysis time-resolved transient conductivities (FP-TRMCs) compared to the individual components of TMDs or C60. This indicates that the heterojunction nanostructure of TMD and C60 facilitated CS. Moreover, the decoration of TMD nanosheets with C60 molecules hindered the undesirable CR between an electron in the electrode and a hole in the TMD nanosheets. Thanks to the accelerated CS and suppressed CR, the photoelectrochemical devices with SnO2/(MoS2 + C60)m and SnO2/(WS2 + C60)m electrodes achieved incident photon-to-current efficiencies (IPCEs) of 35% and 23% at 400 nm, respectively (Table 3). These efficiencies are superior to those of the individual component films (7–13%), and the IPCE of 35% is the highest ever reported for photoelectrochemical devices under three electrodes configuration among composites of 2D layered materials and organic small molecules on semiconducting electrodes.37 As depicted in Fig. 3, the photoinduced CS and subsequent electron transfer (ET) at the heterojunctions contribute to the generation of photocurrent.
System | Layer number | IPCE (wavelength) | TRMC (lifetime, %) | Ref. |
---|---|---|---|---|
SnO2/(MoS2 + C60)m | 3 | 35% (400 nm) | 1.7 × 10−3 cm2 V−1 s−1 (0.2 μs, 59%; 7.0 μs, 16%; 270 μs, 25%) | 37 |
SnO2/(WS2 + C60)m | 3 | 23% (400 nm) | 1.8 × 10−3 cm2 V−1 s−1 (0.4 μs, 28%; 7.0 μs, 11%; 300 μs, 61%) | 37 |
SnO2/(Sb + C60)m | 8 | 11.4% (400 nm) | 3.3 × 10−4 cm2 V−1 s−1 (7.9 μs, 33%; 410 μs, 67%) | 38 |
(Bi + C60)m | 3 | — | No signal | 39 |
Fig. 3 Photoinduced ET pathways in SnO2/(MoS2 + C60)m. The number of parenthesis denotes the order of the pathways. |
The device featuring the MoS2/C60 heterojunction demonstrated a higher IPCE compared to the device with the WS2/C60 heterojunction. Given that the TRMC values of SnO2/(MoS2 + C60)m (1.7 × 10−3 cm2 V−1 s−1) and SnO2/(WS2 + C60)m (1.8 × 10−3 cm2 V−1 s−1) are comparable, the difference could be due to the inhomogeneous film structure of the latter, which results from the strong interaction between WS2 and C60. The insights gained from these findings will offer valuable information on the photodynamics of TMD–organic nanohybrids. Consequently, this will aid in the rational design of organic–inorganic hybrid devices.
Black phosphorous (BP) is a 2D material exhibiting both high carrier mobility and a tunable direct bandgap.40 However, its high sensitivity to light, water, and oxygen limits its applications. Sb, an exfoliated 2D material obtained from bulk antimony, is considered a potential alternative for BP because of its superior stability.41,42 Imahori and coworkers focused on several-layered Sb (approximately 8 layers), which was noncovalently functionalized with C60 molecules by swiftly adding a poor solvent (acetonitrile) into a mixed dispersion of several-layered Sb and C60 in a good solvent (toluene).38 The aggregation behavior of the mixed dispersion of several-layered Sb and C60 in toluene/acetonitrile was examined by UV-visible absorption, dynamic light scattering, and microscopic measurements. These measurements confirmed the formation of the several-layered Sb–C60 composite, (Sb + C60)m, where the Sb surface is covered with C60 molecules. In a FP-TRMC measurement, (Sb + C60)m exhibited a rapid increase in transient conductivity, whereas no signals were detected in FP-TRMC measurements of the individual components, several-layered Sb and C60. This demonstrates the occurrence of photoinduced CS between the several-layer Sb and C60 in (Sb + C60)m. Moreover, a photoelectrochemical device with a SnO2/(Sb + C60)m electrode displayed an enhanced IPCE of 11.4% at 400 nm, which is higher than those of the individual components, either several-layered Sb (8.1%) or C60 (6.9%). This result can be attributed to the photoinduced CS between several-layered Sb and C60 (Fig. 4). The photoelectrochemical properties of the several-layered (Sb + C60)m are lower than those of the corresponding SnO2/(MoS2 + C60)m and SnO2/(WS2 + C60)m probably due to the poor semiconducting properties of Sb with multi-layered structures, as seen in the TRMC data (3.3 × 10−4 cm2 V−1 s−1). Nevertheless, this work represented the initial step toward the fabrication of antimonene–organic molecule heterojunctions for their application in optoelectronics.
Fig. 4 Photoinduced ET pathways in SnO2/(Sb + C60)m. The number of parenthesis denotes the order of the pathways. |
Bi, an exfoliated 2D material obtained from bulk bismuth, has attracted great deal of attention because of its unique electronic and spintronic properties.43 Imahori and coworkers reported few-layered Bi with an average thickness of 1.0 nm (3 layers in average), which was prepared via a successive ball mill and sonication method.39 They also hybridized Bi with C60 molecules by injecting acetonitrile into the mixed dispersion of Bi and C60 in toluene. In the FP-TRMC measurement, the few-layered Bi–C60 composite, (Bi + C60)m, showed no transient conductivity. This is in sharp contrast to the significant transient conductivity observed in the composite of Sb and C60. The energy level of the Bi oxidative excited state (−0.6 V vs. NHE) is higher than the LUMO level of C60 (−0.2 V vs. NHE) and the energy level of the Bi oxidation state (0.8 V vs. NHE) is higher than those of the C60 excited triplet (1.4 V vs. NHE) and singlet (1.7 V vs. NHE) states. Despite the energetically favorable ET processes, they were found to be inefficient. The reason for no occurrence of ET is currently unclear, but the inferior semiconducting properties of few-layered Bi, as seen in the TRMC data, may be responsible for the absence of photoinduced CS at the heterojunction. So far there is no systematic investigation on the relationship between defect sites in 2D layered materials and photophysical and photoelectrochemical properties.
Regarding the use of TMD chalcogen vacancies at edge sites, thiol derivatives have been functionalized. In particular, 1,2-dithiolanes have proved to be useful for covalent functionalization. Tagmatarchis and coworkers created a number of TMDs that were covalently functionalized with organic electron donors or acceptors using 1,2-dithiolanes.44,45 For instance, they reported the covalent functionalization of exfoliated semiconducting MoS2 which on average had 6 layers, by a 1,2-dithiolane bearing an alkyl chain terminated with a 1-pyrenebutoxy ester (Fig. 5).46 The MoS2-based nanohybrids were characterized by spectroscopic, thermal and microscopic methods. Density functional theoretical studies, together with X-ray photoelectron spectroscopy (XPS) analysis, indicated preferential edge functionalization, primarily via sulfur addition along partially sulfur saturated zig-zag MoS2 molybdenum-edges. Before functionalization, the S 2p signal showed the major pair of peaks at 163.4 and 162.2 eV, assigned to the intact MoS2, with the minor one at 163.1 and 161.9 eV, derived from “vacancy” or damaged sites. After functionalization, these minor peaks disappeared as all the edges became fully sulfur-saturated. At the same time, there was no significant change in the Mo 3d signal. The functionalization ratio was estimated to be one per every 24 MoS2 units by thermogravimetric analysis (TGA). This rather low value is consistent with the selective edge-functionalization together with the multi-layered MoS2.
Fig. 5 Schematic of several-layered MS2 (M = Mo or W) sheets covalently functionalized with pyrene molecules. Dotted circles denote defect sites at the edge. |
The UV-Vis spectrum of the MoS2–pyrene hybrid in DMF exhibited the characteristic features arising from the semiconducting MoS2 and pyrene moieties. The fluorescence spectrum of the MoS2–pyrene hybrid showed moderate 50% quenching of the pyrene fluorescence and the appearance of a broad emission at 490 nm. The authors suggested the occurrence of photoinduced energy transfer (EnT) or ET from the pyrene excited singlet state to the MoS2, but no convincing experimental evidence was given. Later, they also observed a similar photoluminescence spectrum of WS2–pyrene hybrid using the same pyrene derivative.47 At that time, the authors indicated the formation of a pyrene excimer that emits at 490 nm. Their calculations also predicted that the pyrene is stacked preferentially parallel to the MoS2 basal plane with the help of intermolecular interactions. Therefore, the emission at 490 nm might be interpreted as the CT emission arising from the CT excited state between the TMD and pyrene, as reported very recently by Imahori and coworkers (vide infra).48 In such a case, some fractions of the pyrene moieties would not interact with the MoS2 basal plane, leading to no quenching of the pyrene excited singlet state by the MoS2 because of the edge functionalization of the MoS2. This is consistent with the moderate quenching of the pyrene fluorescence and the appearance of a broad CT emission.
Tagmatarchis and coworkers extended this approach to develop a MoS2–ZnPc hybrid. Specifically, ZnPc with a 1,2-dithiolane oxide linker was used to functionalize MoS2 at defect sites located at the edges (Fig. 6).49 The structure of MoS2–ZnPc was assessed by spectroscopic, thermal, and microscopic methods, but XPS analysis was not conducted to support the covalent functionalization. The functionalization ratio was estimated to be very low (one per every 103 MoS2 units by TGA). There is no information on the number of the exfoliated MoS2, but the transmission electron microscope (TEM) image indicates the multi-layered structures. An energy-level diagram showing different photochemical events in MoS2–ZnPc was presented, indicating the occurrence of photoinduced ET to generate the MoS2˙−–ZnPc˙+ state. The photodynamic behaviors obtained by femtosecond TA spectroscopy were complicated by the involvement of continuously relaxing MoS2 excitons, making it difficult to give reliable assignment and their time constants for transient species. Nevertheless, the authors claimed to detect fingerprint of the ZnPc radical cation at 840 nm, indicating the occurrence of photoinduced CS.
Fig. 6 Schematic of multi-layered MoS2 sheets covalently functionalized with zincphthalocyanine (ZnPc) molecules. Dotted circles denote defect sites. |
This result is in sharp contrast with the covalently linked MoS2-free base porphyrin (H2P) hybrid, where the first EnT occurs from the H2P excited singlet state to the MoS2, followed by the second EnT from the MoS2 excited state to the H2P, without generating the charge-separated state.50 It is noteworthy that photoinduced CS with a time constant of 80 fs and subsequent CR with a time constant of 1–100 ps to generate the ZnPc triplet excited state was also observed for the heterojunction between bulk, single crystal MoS2 and face-on oriented, few-layered ZnPc (vide supra).22,25 The corresponding monolayer MoS2/ZnPc interface exhibited sub-100 fs CT from the ZnPc excited singlet state to the MoS2, followed by the dissociation to an electron–hole pair. The difference in the single crystal MoS2 and the monolayer MoS2 was rationalized by the different amount of band bending. The small band bending in ZnPc on monolayer MoS2 can widen the spatial range of the coherent CT and facilitate the dissociation. This emphasizes the importance of the interfacial energy level alignment at the heterojunctions between 2D layered materials and small organic molecules.
Tagmatarchis and coworkers reported the covalent functionalization of MoS2 with an electron-acceptor, perylenediimide (PDI), which possesses a 1,2-dithiolane oxidized at each sulfur atom with a 60:40 ratio (Fig. 7).51 The MoS2–PDI hybrid structure underwent evaluation using spectroscopic, thermal, and microscopic methods. XPS analysis was not conducted to verify the covalent functionalization. The functionalization ratio was estimated to be one per every 97 MoS2 units by TGA. The average number of the exfoliated MoS2 layer was calculated to be 3–5 by analysing average heights using AFM. The photophysical relaxation of MoS2–PDI was visualized in an energy-level diagram, revealing the possibility of photoinduced CS leading to formation of the MoS2˙+–PDI˙− state. Additionally, femtosecond TA studies were performed on the MoS2–PDI hybrid in DMF. The spectral features of 1PDI*, MoS2*, and PDI are similar, and there is no selective excitation for the two moieties, making the assignment of the transient species difficult. However, the decay associated spectra exhibited three components: τ1 = 1.69 ps, assigned to the PDI and MoS2 excited states, τ2 = 2.36 ns, characterized by a peak of PDI˙− at 730 nm, and τ3 = 5.37 ns, arising from the unmodified MoS2 excited state (Table 4). These results highlight the potential of electron acceptor-tethered MoS2 and its application in optoelectronic devices.
Fig. 7 Schematic of several-layered MoS2 sheets covalently functionalized with perylenediimide (PDI) molecules. Dotted circles denote defect sites. |
The covalently-linked hybrids described so far possess a flexible linker between organic molecules and few-layered 2D materials, making it difficult to analyze the photodynamics due to multiple conformers arising from the flexibility and continuous relaxation of the MoS2 excited states. Furthermore, the use of defect sites in TMDs also limits the degree of functionalization on TMDs significantly, inhibiting selective excitation of organic molecules against MoS2.20–25 However, the poor reactivity at the surface of 2D layered materials has inhibited the covalent modification by organic molecules with a well-defined structure. Imahori and colleagues rationally designed a heterointerface between a photoactive molecule (i.e., pyrene, Py) and a semiconducting 2D inorganic monolayer (MoS2) with a well-defined bridge at the nanoscale (MoS2–Py). Py and MoS2 are linked by a short, semi-rigid bridge, N-benzylsuccinimide (BSI) (Fig. 8).48
Fig. 8 Schematic of MoS2Py hybrid where a pyrene (Py) is linked to a semiconducting 2D inorganic monolayer (MoS2) with a well-defined bridge N-benzylsuccinimide (BSI). |
The BSI groups can be covalently bonded onto the MoS2 surface at high densities by the reaction with N-benzylmaleimide (BMI) according to the report by Pérez and coworkers.52,53 Namely, the entire MoS2 surface is completely covered with the BSI moieties covalently. Because of the densely filled BSI moieties on MoS2, the attached Py moieties are significantly inclined toward the BSI-modified MoS2 and are in direct contact with the BSI moieties at a separation distance of approximately 1 nm. This is different from the sparse edge functionalization by Tagmatarchis and coworkers.44–47,49–51 The Py moiety is tethered to the para-position of the phenyl group by Suzuki coupling after the first-step modification of the BSI moiety on MoS2. Since the modification ratio by the Suzuki coupling is low, only 30% of the BSI on MoS2 is modified with the Py moiety. The moderate efficiency of the Suzuki coupling results from the detachment of halogen atoms from the para-position of the phenyl group during the reaction.54,55 Thus, the excited singlet state of Py at the single molecular level is expected to interact solely with MoS2 to form the CT excited state (i.e., excited state with CT character) and/or the charge-separated state at the well-defined heterointerface of MoS2–Py.
The photoluminescence spectra of MoS2–Py and the pyrene reference were taken in DMF. The pyrene reference, when excited at 345 nm, showed intense emission peaks at 382 and 402 nm with an emission quantum yield of 56%. When the photoluminescence spectrum of MoS2–Py was recorded with an excitation wavelength at 345 nm, the excited singlet state of pyrene was strongly quenched, and a broad, red-shifted intense emission was observed at 453 nm. The emission decay profiles exhibited the photoluminescence lifetime of MoS2–Py (2.7 ns), which is much shorter than that of the pyrene reference (77 ns). As solvent polarity increased from toluene to DMF, MoS2–Py also revealed a gradual red-shift and decrease in emission, which can be assigned to CT emission. Surprisingly, the emission quantum yield of MoS2–Py (68% in DMF) is high, considering the emission quantum yield of typical intermolecular CT excited states between pyrene derivatives and N,N-dimethylaniline (approximately 20% in nonpolar solvents). This may originate from the short, fixed orientation between the pyrene and MoS2 through the BSI moieties on MoS2. In other words, this geometry stabilizes the CT excited state and suppresses the nonradiative decay of the CT excited state.
The picosecond TA spectroscopic measurement of MoS2–Py was conducted with an excitation wavelength at 345 nm, where the excitation ratio of MoS2 : Py is 1 : 2. Three components with lifetimes of τ1 = 0.9 ps, τ2 = 117 ps, and τ3 = 2540 ps were deduced from the global analysis. The first and second decay components, characterized by bleach arising from the excited state of MoS2, were assigned to the decay of the hot and cooled MoS2 excitons. The third component, which shows broad positive signals in the visible and NIR regions with a peak around 860 nm, is assignable to the decay of the CT excited state, which is in good agreement with the emission lifetime of MoS2–Py (2.7 ns). From the rise component at 850 nm, the CT excited state is generated with a time constant of 1.2 ps (Fig. 9 and Table 4). Considering the lifetime of the pyrene reference (77 ns), the CT efficiency from the pyrene excited singlet state to MoS2 is estimated to be ∼100%.
Fig. 9 Relaxation pathways of MoS2–Py in DMF. The energy level of the CT state was determined by the peak position of the emission. The plausible pathways are shown as a black dotted arrow. |
Theoretical studies were also conducted for the MoS2–Py model with a Mo28S63 nanosheet, in which one Py–BSI and two BSI units are covalently attached to sulfur atoms. The geometry optimization of the Py–MoS2 model revealed the geometry of the Py moiety on MoS2 that is similar to Fig. 8, which aligns well with the small thickness (2.6 nm) of the experimentally obtained monolayer MoS2–Py using AFM. Furthermore, the time-dependent density functional theory (TDDFT) and the three-dimensional reference interaction site model (3D-RISM) were used to investigate the emission behavior of MoS2–Py. The LUMO of the Py moiety is hybridized with the vacant orbitals of MoS2, signifying the formation of the CT excited state by interaction between the vacant orbitals of MoS2 and the excited state of pyrene. Specifically, photoexcitation of the Py moiety in the MoS2–Py covalently linked system effectively formed the long-lived, highly emissive CT excited state, delocalizing over the heterointerface of pyrene and monolayered MoS2. The highly emissive CT excited state may provide insight into how to reduce the voltage loss stemming from nonradiative CR in organic solar cells (OSCs). As has been suggested by several groups,56–60 if the CT excited state becomes more emissive after hybridizing between the locally-excited (LE) and CT excited states due to the small offset for the CT formation, nonradiative vibrational CR would be inhibited, reducing the voltage loss. Such CT states would be transformed into charge-separated states by either attaching an additional donor moiety to the pyrene or weakening electronic coupling between the pyrene and MoS2. Accordingly, this study provides a fundamental understanding of interactions in the excited states at heterojunctions between 2D monolayered materials and single organic molecules. This will be useful for the applications of organic–inorganic 2D heterointerfaces in optoelectronic devices and sensors. However, exfoliation and chemical modification may increase defect sites in TMDs, deteriorating their intrinsic photophysical properties. The effects of defect sites in 2D layered materials need to be examined systematically for better understanding of photodynamics at the heterojunctions.
Several researchers have observed photocurrent generation at covalently-linked heterojunctions between 2D materials and organic molecules, but the detailed mechanism has not been addressed because of the use of flexible bridges.61–63
Heteojunction | Type | Detection range | Responsivity | Response speed | Ref. |
---|---|---|---|---|---|
a BTBT: (12-(benzo[b]benzo[4,5]thieno[2,3-d]thiophen-2-yl)dodecyl)-phosphonic acid. b PDVT-10: poly{3,6-dithiophen-2-yl-2,5-di(2-decyltetradecyl)}-pyrrolo[3,4-c]pyrrole-1,4-dione-alt-thienylenevin. c PDPP3T: poly[[2,5-bis(2-hexyldecyl)-2,3,5,6-tetrahydro-3,6-dioxo-pyrrolo[3,4-c]pyrrole-1,4-diyl][2,2′:5′,2′′-terthiophene]-5,5′′-diyl]. d PTCDA: perylenetetracarboxylic dianhydride. | |||||
MoS2/BTBTa | Transistor | 406 nm | 475 A W−1 | 5.5–17.8 s | 64 |
MoS2/ZnPc | Transistor | 532 nm | 430 A W−1 | 8–72 ms | 65 |
MoS2/CuPc | Transistor | 520 nm | 1.98 A W−1 | 0.3 s | 66 |
MoS2/PDVT-10b | Transistor | White light | 1.38 A W−1 | — | 67 |
MoS2/rhodamine 6G | Transistor | 520 nm | 1.17 A W−1 | — | 68 |
MoS2/rubrene | Diode | 532 nm | 0.5 A W−1 | <5 ms | 69 |
MoS2/PDPP3Tc | Transistor | 660 nm | 0.45 A W−1 | 4–40 ms | 70 |
MoS2/C8-BTBT | Diode | White light | 22 mA W−1 | — | 71 |
MoS2/rubrene | Transistor | 532 nm | 0.8 mA W−1 | — | 72 |
MoS2/pentacene | Diode | 532 nm | — | 1.8–2 ms | 73 |
MoSe2/NiPC | Diode | 658 nm | 5.95 A W−1 | — | 74 |
WS2/PTCDAd | Diode | 430 nm | — | — | 75 |
InSe/C8-BTBT | Diode | 532 nm | — | — | 76 |
Bi2Se3/rubrene | Transistor | 532 nm | 124 A W−1 | 10–59 ms | 77 |
Bi2Te3/pentacene | Diode | 450–3500 nm | 14.89 A W−1 | 2.98–3.57 ms | 78 |
Another issue to be addressed is their layered-number dependent properties at the heterojunctions. However, conventional approaches need multiple complicated preparation procedures to fabricate the heterojunctions. Choi and coworkers developed a series of heterojunctions from eosin Y (EY) molecules stacked on mechanically exfoliated WSe2 flakes of ML through 6-layer (6L) via one-step solution phase chemistry.79 As the layer number of WSe2 increases, its CB level decreases from –3.70 eV and VB level increases from −5.35 eV. EY's LUMO level (−4.05 eV) is lower than the CB of 6L WSe2 and its HOMO level (−4.75 eV) is higher than the VB of ML WSe2, forming a type-II band alignment. When the WSe2 was excited, the WSe2/EY heterojunction showed significantly enhanced photoconductivity compared to pristine WSe2 with a lower barrier height and a longer effective screening length. As the layer-number increased, the photoresponse decreased gradually and finally increased slightly. Overall, the staggered bands between the EY layer and ML-WSe2 flake can facilitate the exciton dissociation efficiently, leading to enhanced photoconductivity. This is consistent with the results on ML-MoS2/ZnP.22,25
Unexplored potential applications of p–n heterojunctions between small organic molecules and 2D layered materials include photocatalysts and solar cells. So far, a combination of 2D layered materials with inorganic materials has been investigated for the applications. For photocatalysts, improving the CS efficiency and lifetime is essential. The charge-separated state must be quenched efficiently by diffusing reagents in solutions.81,82 For solar cells, such p–n heterojunctions have not been used at this stage because of insufficient photovoltaic parameters for achieving high IPCE values. ML-2D layered materials are potential candidates, as seen in ML-MoS2/ZnP heterojunction.22,25 We need to develop organic molecules with an excellent light-harvesting property in visible and near-infra red regions, long-lived excited state (e.g., long exciton diffusion length),83,84 and self-CT and CD without bulk heterojuctions,59 as demonstrated in non-fullerene acceptors.
HOMO and LUMO of organic small molecules can be tuned to make a desirable heterojunction with 2D layered materials. If we can modulate the electronic coupling at the p–n heterointerface intentionally, efficient formation of CT excited states and subsequent charge dissociation at the interface for optoelectronics and solar cells will be achieved.
This journal is © The Royal Society of Chemistry 2025 |