1. Introduction
Click to copy section linkSection link copied!
Optical technologies have become indispensable in modern scientific and industrial approaches for the analysis and monitoring of the physical and chemical properties of objects and environments. (1−3) In particular, the near-infrared (NIR) wavelength range, spanning approximately 750–1400 nm, has attracted increasing interest in the fields of medical diagnostics, (4) biometric sensing, (5) optical communication, (6) autonomous driving, (7) facial recognition, (8) and geographic remote sensing. (9) The unique penetrating power in this wavelength range is valuable in these contexts; potential applications of NIR technology continue to expand.
NIR detection has generally relied on sensors constructed from inorganic semiconductor materials. (10−12) These sensors have limited applications because they often require complex manufacturing processes, lack mechanical flexibility, and are very sensitive to temperature changes. Organic semiconducting materials may overcome these limitations; their main advantages include a broad spectral response, tunable energy levels, and cost-effective manufacturing. (13−15) Despite such potential, the development of organic semiconductors that absorb at wavelengths beyond 1000 nm remains a substantial ongoing challenge, primarily because of the difficulty in creating the narrow band gap required to effectively absorb long wavelengths. Operating beyond the 1000 nm threshold necessitates extremely narrow band gaps, which, if not carefully controlled, can lead to higher exciton recombination, increased turn-on voltages, and possible morphological instability. (16) Thus, research on NIR organic photodetectors (OPDs) that operate beyond 1000 nm has been relatively minimal. (17,18) New organic semiconductors that efficiently absorb wavelengths exceeding 1000 nm would be very useful.
Till date, most researchers have used the alternating donor–acceptor (D–A) approach to engineer the band gaps of conjugated polymers (CPs). (19−22) Strong acceptor units have been utilized to develop CPs with narrow band gaps. (23,24) The introduction of open-shells into conjugated polymers has recently been used to develop CPs with extremely narrow band gaps. (25,26) Open-shell CPs exhibit several novel magnetic properties attributable to their diradical characteristics. (27,28) Despite the excellent results of D–A CPs, most D–A CPs only absorb up to a relatively short NIR region, typically around 800 nm. (29,30) Although a few studies have demonstrated NIR detection around 1000 nm using conjugated polymers, such as an OPD with a detectivity of 9.86 × 1011 Jones at 900 nm by He et al. (31) and a detectivity of 1.03 × 1010 Jones at 1200 nm reported by Jacoutot et al., (32) such examples still remain rare. Despite the improvement in NIR detection beyond 1000 nm, open-shell CPs with extremely narrow band gaps developed until now still face several challenges in practical applications. The narrow band gap can lead to a decreased on/off ratio in devices, resulting in performance degradation. This is particularly critical in switching devices, such as transistors or photodetectors.
To solve these issues, we synthesized a novel open-shell conjugated terpolymer, poly{2,5-bis(2-decyltetradecyl)-3,6-di(thiophen-2-yl)-2,5-dihydropyrrolo[3,4-c]pyrrole-1,4-dione]-co-thiophene-co-benzo[1,2-c;4,5-c′]bis[1,2,5]thiadiazole} (PDPPTBBT). PDPPTBBT was designed to exhibit open-shell characteristics by incorporating three carefully selected monomers, diketopyrrolopyrrole (DPP), thiophene, and benzo[1,2-c;4,5-c′]bis[1,2,5]thiadiazole (BBT) and showed effective NIR absorption beyond 1000 nm. The combination of these three units enabled the formation of a random terpolymer with a suitable band gap for OPD applications. This polymer promoted strong interchain aggregation, which remarkably enhanced the stability of diradicals and broadened the absorption spectrum.
The repeating unit structure of the PDPPTBBT backbone plays a critical role in enhancing both the NIR absorption and electron paramagnetic resonance (EPR) signals. First, the integration of DPP and BBT into the backbone introduces open-shell characteristics, allowing the polymer to exhibit a diradical ground state. This state leads to stronger interchain π–π interactions after thermal annealing, which, in turn, results in an increase in NIR absorption, especially in the range beyond 1000 nm. The shoulder peak observed around 1200 nm after annealing indicates that these diradical interactions are further stabilized by enhanced molecular packing, thus extending the absorption capabilities into the longer wavelengths of the NIR region. In addition, the diradical nature of PDPPTBBT is closely linked to its EPR signals. The open-shell configuration facilitates the delocalization of unpaired electrons, which align with external magnetic fields to produce temperature-independent Pauli paramagnetic properties. Thermal annealing enhances the interchain aggregation, increasing the density of diradicals and, thus, amplifying the EPR signal intensity. This correlation among the backbone structure, molecular aggregation, and enhanced diradical stability is a key feature that distinguishes PDPPTBBT from other conjugated polymers.
Furthermore, the new open-shell conjugated polymer served as a donor when mixed with Y6 but as an acceptor when mixed with poly(3-hexylthiophene-2,5-diyl) (P3HT); the energy level alignments differed. Importantly, we present a rare polymer acceptor and demonstrate that the fabrication of an all-polymer-based OPD capable of detecting light beyond 1000 nm is feasible. Using PDPPTBBT, we fabricated the OPDs via blending of either P3HT (PDPPTBBT:P3HT) or Y6 (PDPPTBBT:Y6) and evaluated the performance of the OPDs. For devices based on PDPPTBBT:P3HT, the external quantum efficiency (EQE) was 153% and the specific detectivity (D*) was 2.0 × 1011 Jones at 850 nm. For devices based on PDPPTBBT:Y6, the EQE was 126% and the D* was 7.5 × 1011 Jones at 1050 nm. These performance metrics surpass those reported in previous studies of open-shell CPs, underscoring the potential of PDPPTBBT as a promising material for future advancements in photodetector technology. We thus open a new avenue toward the fabrication of NIR OPDs; our open-shell polymer effectively absorbs in the NIR spectrum.
2. Results and Discussion
Click to copy section linkSection link copied!
2.1. Synthesis and Thermal Properties
The synthesis pathway of the random terpolymer PDPPTBBT is depicted in Figure 1a. PDPPTBBT was synthesized via Stille polymerization in a microwave oven using Pd(PPh3)4 as a catalyst when combining the monomers 3,6-bis(5-bromothiophen-2-yl)-2,5-bis(2-decyltetradecyl)-2,5-dihydropyrrolo[3,4-c]pyrrole-1,4-dione (M1), 2,5-bis(trimethylstannyl)thiophene (M2), and 4,7-dibromobenzo[1,2-c:4,5-c′]bis([1,2,5]thiadiazole) (M3) in a 1:2:1 ratio. Polymer solubility was improved by the attachment of 2-decyltetradecyl alkyl chains to the DPP monomer; PDPPTBBT was well-dissolved in chloroform. The crude polymer product was purified via Soxhlet extraction and finally collected in chloroform to obtain the desired material in a yield of 41%. Gel permeation chromatography (GPC, 35 °C, tetrahydrofuran) revealed that the number-average molecular weight and dispersity of PDPPTBBT were 10.5 kDa and 3.13,kDa, respectively (Figure S1). Thermogravimetric measurements were performed to confirm the thermal stability. Decomposition commenced at 320 °C (Figure S2).
Figure 1
High Resolution Image
Download MS PowerPoint Slide
2.2. Photophysical and Electrochemcial Properties
Figure 1b shows the absorption spectra of PDPPTBBT, PDPPTBBT:Y6, and PDPPTBBT:P3HT thin films before and after annealing. PDPPTBBT and PDPPTBBT:P3HT films were annealed at 230 °C, whereas PDPPTBBT:Y6 films were annealed at 180 °C. As a typical D–A conjugated polymer, PDPPTBBT exhibits a shorter-wavelength band originating from π–π* transitions and a longer-wavelength band attributed to intramolecular charge transfer between monomers. The main absorption peak of the PDPPTBBT film is located at approximately 780 nm. Notably, the shoulder peak around 1200 nm intensifies after annealing, which suggests that interchain aggregations stabilize the diradical species generated in the open-shell polymer, thereby enhancing absorption near 1200 nm. A similar increase in the shoulder peak intensity is also observed from PDPPTBBT:Y6 and PDPPTBBT:P3HT films. In contrast, Figure 1c shows the absorption spectra of PDPPTBBT in toluene from 25 to 200 °C. As the solution temperature increases from room temperature to 200 °C, the absorption peak near 700 nm undergoes a blueshift and the shoulder peak at around 1200 nm gradually diminishes. This indicates that the polymer transitions from a disordered aggregate state at lower temperatures to a nonaggregated chain state at higher temperatures, reducing interchain interactions. Consequently, the stabilization of the diradical species in the open-shell polymer is weakened in solution at elevated temperatures, leading to decreased absorption near 1200 nm. This phenomenon will be addressed in detail below. Meanwhile, based on the thin-film absorption onsets in the thin films, the optical band gap (Egopt) of the PDPPTBBT-based films was estimated to be 0.75 eV.
To explore the electrochemical properties, cyclic voltammetry (CV) analysis was performed to estimate the frontier energy levels based on the onset potentials of oxidation (Eox) and reduction (Ere) (Figure 1d). The Eox/Ere of PDPPTBBT was 0.67/–0.76 V corresponding to highest occupied molecular orbital (HOMO)/ lowest unoccupied molecular orbital (LUMO) energy levels of–5.09/–3.66 eV, respectively. The electrochemical band gap (EgCV) of PDPPTBBT was 1.43 eV, differing from that of Egopt, due to the low dielectric constant and significant exciton-binding energy of the organic material. (33) The optical band gap is determined by the onset of photon absorption, while electrochemical measurements gauge intrinsic HOMO–LUMO gaps. Organic semiconductors often exhibit large exciton-binding energies; therefore, Egopt can be lower than EgCV. The optical and electrochemical properties are summarized in Table 1.
Table 1. Optical and Electrochemical Properties of PDPPTBBT
polymer | λmaxsol (nm) | λmaxfilm (nm) | Egopta (eV) | EHOMOb (eV) | ELUMOc (eV) |
---|---|---|---|---|---|
PDPPTBBT | 740 | 780 | 0.75 | –5.09 | –3.66 |
a
Determined using the absorption onset wavelength.
b
Calculated from the oxidation onset potential.
c
Calculated from the reduction onset potential.
2.3. Diradical Characteristics
To understand the polymer electronic structure and properties, density functional theory (DFT) calculations were performed at the unrestricted B3LYP/6-31G(d,p) level for the DPPTBBT dimer. In the calculations, the hexadecyl alkyl side chains were replaced with methyl groups. In Figure 2a (top and middle), the α and β singly occupied molecular orbitals (SOMOs) of BBT are apparent, which confirms that BBT played a crucial role in diradical production. BBT allows two electrons with opposite spins to correlate in separate spaces. This correlation further reduces the covalent character of the singlet and thus lowers electron repulsion; an open shell ultimately forms. The singlet–triplet energy gap (ΔEST) for the DPPTBBT dimer was −22.1 kcal mol–1. The transition between the closed- and open-shell states is facilitated by such a small transition energy. Additionally, the spin density distribution shown in Figure 2a (bottom) indicates that the BBT unit density is high across the entire unit, suggesting that the diradical form contributes to the ground state. Therefore, due to the two nonbonding electron pairs generated at the ends of BBT, the dimer loses its conjugation, causing thiophene and DPP to transform into quinoidal forms.
Figure 2
High Resolution Image
Download MS PowerPoint Slide
To investigate polymer spins and magnetic properties, zero-field-cooled (ZFC) and 10 kOe field-cooled (FC) magnetization measurements were conducted from 2 to 300 K using a superconducting quantum interference device. Amorphous solid polymer powders were analyzed; these resemble spin-coated thin films. (34) Figure 2c shows the mass magnetic susceptibility dependencies (χm values) of the powders. The ZFC-FC curves exhibit negligible splitting; no significant increase in susceptibility was observed even below 50 K. Between 45 and 80 K, the intriguing behavior suggests a potential role for oxygen in terms of magnetization.
Oxygen is easily adsorbed from air under ambient conditions, and the antiferromagnetic oxygen transition around 50 K is well-known. χm T is directly proportional to the temperature (Figure 2d). The magnetizations as functions of the magnetic field (M–H) curves of amorphous solid powders recorded at various temperatures (2, 3, 5, 10, 50, and 100 K) were temperature-independent (Figure 2e). Figure 2e also shows that the total magnetization was positive at a high magnetic field (+50 kOe) but negative at a low magnetic field (−50 kOe). Thus, the Pauli paramagnetic susceptibility (χPauli) was 3.2 μemu g–1 Oe1–, and PDPPTBBT exhibited temperature-independent Pauli paramagnetic properties that differ from the typical Curie–Weiss behavior. (35,36) Consequently, it is inferred that the diradicals within PDPPTBBT align with the external magnetic field, resulting in a Pauli paramagnet-like susceptibility, which is different from other open-shell CPs (Table S1).
EPR spectroscopy was used to explore the ground state of PDPPTBBT. Initially, spin-coated thin films were evaluated before and after annealing. Figure 2f shows that EPR intensity was enhanced after annealing (red line) compared to that before annealing (pink line), but both lines exhibited the same g-factor of 2.0048, suggesting the presence of unpaired or loosely coupled electrons within the polymer thin films. The relative densities of those films increased after annealing due to enhanced interchain polymer aggregation, in line with the rise in absorption around 1200 nm mentioned above. Next, the EPR spectra were fitted using both Gaussian (blue dashed line) and Lorentzian (black dashed line) functions (Figures 2f and S6). Either function may adequately fit an EPR spectrum due to unpaired electron charges. The Gaussian function indicates that the charges are in denser states and thus relatively immobile; the Lorentzian function implies that the charges are relatively mobile. (37) The observed EPR signals conform to the Lorentzian (black dashed line), indicating the presence of delocalized diradicals. (38) To further understand the intrinsic magnetic properties of PDPPTBBT, low-temperature EPR measurements were performed from 10 to 293 K (Figures 2g and S7). The EPR signal intensity (IEPR) is temperature (T)-dependent. In particular, as the temperature decreases, the IEPR of the triplet state rises, (39) indicating a paramagnetic ground state. (35,36) All EPR signal g-factors were 2.0047, and the line widths were relatively broad (approximately 4–5 G). Figure 2h shows the relationship between IEPR and the reciprocal temperature (1/T) fitted using the Bleaney–Bowers eq (eq 1); we then extracted the energy differences between singlet and triplet states (ΔEST values). The calculated ΔEST values were 14.3 × 10–3, 18.3 × 10–3, and 14.1 × 10–3 kcal mol–1 for PDPPTBBT, PDPPTBBT:P3HT, and PDPPTBBT:Y6, respectively; thus, the triplet ground state was more stable than the singlet ground state. The exchange coupling constants (eq 2) were 2.5, 3.2, and 2.5 cm–1 for PDPPTBBT, PDPPTBBT:P3HT, and PDPPTBBT:Y6, respectively. These results are consistent with the weak intramolecular ferromagnetic coupling apparent between spins (Figure 2b). (35,36)
2.4. Film Morphology, Miscibility, and Molecular Stacking
Molecular stacking transitions within PDPPTBBT induced by thermal annealing and/or blending with P3HT or Y6 molecules were investigated by using a grazing incidence wide-angle X-ray scattering (GIWAXS) platform. Figures S8 and S9 show that PDPPTBBT films exhibited the (100) peaks of lamellar packing in both in-plane and out-of-plane directions, with a lamellar distance of ∼24.2 Å. After thermal annealing, PDPPTBBT, PDPPTBBT:P3HT, and PDPPTBBT:Y6 thin films displayed more distinct lamellar stacking peaks [(100), (200), (300), and (400)] in the out-of-plane direction (Figures 3a–c and S8–S11). This observation implies that all polymers displayed increased crystallinity and preferred lamellar-like molecular orientations. When PDPPTBBT was blended with P3HT and Y6, peaks of P3HT and Y6 were observed (Figure 3d,e). The 2D GIWAXS data for P3HT and Y6 thin films are shown in Figures S11–S14. The coherence lengths (Lc values) and d-spacings associated with various film fabrication conditions were calculated by using the out-of-plane (100) lamellar stacking peaks (Figure 3f,g). For PDPPTBBT thin films, Lc values increased, but the d-spacing decreased after thermal annealing due to enhanced molecular packing after thermal rearrangement. Such packing also increased diradical stability, which is associated with the above-mentioned increase in the EPR signal. The d-spacings of annealed PDPPTBBT:P3HT and PDPPTBBT:Y6 thin films were higher than those of PDPPTBBT thin films; this difference is attributable to certain steric effects of the introduced donor or acceptor molecules. However, the Lc values of the PDPPTBBT:P3HT thin films did not change; crystallinity was maintained. Interestingly, the PDPPTBBT:Y6 thin film showed a much clearer (010) peak in the out-of-plane direction, which is attributed to the face-on arrangement preference of Y6 molecules and their intermolecular interactions with PDPPTBBT polymers (Figures S11 and S14). Therefore, enhanced π–π stacking and face-on orientation in PDPPTBBT:Y6 blends lead to the optimal molecular conformation of the bulk heterojunction system for NIR OPDs.
Figure 3
High Resolution Image
Download MS PowerPoint Slide
Next, PDPPTBBT:P3HT and PDPPTBBT:Y6 thin films prepared under various annealing conditions were subjected to atomic force microscopy (AFM) (Figure S15). The root-mean-square roughness (Rq) values of PDPPTBBT:P3HT and PDPPTBBT:Y6 thin films were measured. The Rq values of PDPPTBBT:P3HT thin films increased (4.38 nm), but those of PDPPTBBT:Y6 films decreased (1.98 nm) after annealing. Comparison of the Lc values of annealed PDPPTBBT:P3HT and PDPPTBBT:Y6 thin films showed that the Lc values of PDPPTBBT:P3HT thin films were greater than those of the PDPPTBBT:Y6 thin film due to polymer aggregation. Conversely, the small molecule Y6 exhibited not only a decrease in Lc value but also a reduction in Rq due to the molecular rearrangement with PDPPTBBT. This difference could be attributed to the local aggregation of polymers and small molecules. However, the morphological properties of the blended films do not impair device performance, ultimately resulting in improved device performance after annealing.
2.5. Device Characteristics
OPDs were fabricated to demonstrate that PDPPTBBT detected NIR (Figure 4a,b). After consideration of the energy levels (Figure 4c), PDPPTBBT was blended with either Y6 or P3HT (acceptor or donor, respectively) when fabricating two types of bulk heterojunction OPDs with the device configuration ITO/ZnO/active layer/MoO3/Au. In the two configurations, PDPPTBBT served as the donor or the acceptor, which affirmed the dual p-type and n-type functionalities of the polymer. The operating mechanisms under reverse bias of both PDPPTBBT:Y6- and PDPPTBBT:P3HT-based devices are illustrated in Figures 4c and S16a, respectively. Figure 1b confirms that the NIR absorption characteristics were maintained in both the PDPPTBBT:Y6 and PDPPTBBT:P3HT blends. After optimization of device fabrication, it was confirmed that NIR OPDs exhibited photoresponses up to a (long) wavelength of 1100 nm. The J–V curves of OPDs based on PDPPTBBT:Y6 and PDPPTBBT:P3HT in the dark and under light at 850 nm (150 μW cm–2) and 1,050 nm (87 μW cm–2) are shown in Figures 4d and S16b. Both devices exhibited low dark current densities (Jdark values) under reverse bias, and Jdark remained stable to −10 V. As the bias increased, the photocurrent density (Jph) rose due to enhanced charge extraction. At −10 V, the Jdark of the OPDs based on PDPPTBBT:Y6 was 6.25 μA cm–2, whereas that of the OPDs based on PDPPTBBT:P3HT was 82.5 μA cm–2. On OPD illumination at 850 and 1050 nm, devices based on PDPPTBBT:Y6 exhibited Jph values of 139 and 92.5 μA cm–2, respectively, which confirmed high NIR sensitivity. OPDs based on PDPPTBBT:P3HT displayed Jph values of 160 and 43.2 μA cm–2 at 850 and 1050 nm, respectively. Compared with devices based on PDPPTBBT:Y6, those fabricated from PDPPTBBT:P3HT exhibited a greater current density increase at 850 nm but a fall in sensitivity at the (long) wavelength of 1050 nm, which is attributable to the relatively low absorption of PDPPTBBT:P3HT thin films in the long-wavelength region.
Figure 4
High Resolution Image
Download MS PowerPoint Slide
To quantitatively describe and compare the performances of the various NIR OPDs, the responsivity (R) and EQE values were calculated by using eqs 3–7). The EQE, R, and D* values of devices based on PDPPTBBT:P3HT and PDPPTBBT:Y6 were measured as a function of the wavelength (Figure 4e,f for PDPPTBBT:Y6 and Figure S16c,d for PDPPTBBT:P3HT). As the bias increased, Jdark remained relatively constant, but Jph significantly increased in association with a higher EQE and D* at −10 V. In particular, OPDs based on PDPPTBBT:Y6 exhibited EQEs of 135 and 126% at 850 nm and 1,050 nm and the corresponding D* values of 6.6 × 1011 and 7.5 × 1011 Jones, respectively. OPDs based on PDPPTBBT:P3HT exhibited EQE values of 153 and 59% and D* values of 2.0 × 1011 and 9.7 × 1010 Jones at 850 and 1050 nm, respectively. The results are summarized in Table 2. The novel open-shell conjugated polymer serves as either the acceptor or donor in OPDs, and it is noteworthy that it can act as one of the rare OPDs capable of detecting wavelengths beyond 1000 nm.
Table 2. OPD Performances at Certain Wavelengthsa
device | λ (nm) | R (A W1–) | EQE (%) | D* (Jones) |
---|---|---|---|---|
PDPPTBBT:Y6 | 850 | 0.903 ± 0.018 | 129.6±3.3 | (6.2±0.3)×1011 |
1050 | 1.002 ± 0.039 | 123.6 ± 2.5 | (6.8 ± 0.5) × 1011 | |
PDPPTBBT:P3HT | 850 | 0.996 ± 0.035 | 147.0 ± 4.5 | (1.8 ± 0.2) × 1011 |
1050 | 0.472±0.017 | 56.2 ± 1.9 | (9.1 ± 0.5) × 1010 |
a
All devices were biased at −10 V.
To investigate the linear dynamic range (LDR) of the OPDs, photocurrents were measured at −10 V over a wide range of incident intensities (Figure S17). Under these conditions, the OPDs based on PDPPTBBT:P3HT and PDPPTBBT:Y6 exhibited an LDR of 20 and 25 dB, respectively, indicating a more linear response in the Y6 blend. In addition, the dynamic performance was evaluated by measuring rise/fall times and the −3 dB cutoff frequency (f–3db) at −10 V (Figure S18). The rise (tr) and fall times (tf) are defined as the time for the photocurrent to increase from 10 to 90% of its maximum value and from 90 to 10% of its maximum value, respectively. The OPDs based on PDPPTBBT:P3HT showed a tr of 16 μs and a tf of 27 μs, while those based on PDPPTBBT:Y6 exhibited a tr of 14 μs and a tf of 17 μs (Figure S18a). The f–3db values were measured under 10 Hz-modulated LED illumination. The f–3db value for the PDPPTBBT:P3HT devices was 125 kHz, whereas that of the PDPPTBBT:Y6 devices was 236 kHz, confirming faster response speeds in the PDPPTBBT:Y6 based on the OPDs (Figure S18b).
To further investigate the mechanisms involved, the charge mobilities of PDPPTBBT:P3HT and PDPPTBBT:Y6 thin films were measured using the space charge limited current (SCLC) method. (40,41) To determine the mobilities of holes and electrons in the two types of thin films, hole-only and electron-only devices were fabricated. The thin films were coated, as were the OPD devices discussed above. To create a hole-only device, PEDOT:PSS was substituted for ZnO; to prepare an electron-only device, BCP was used instead of MoO3. As shown in Figure S19, when the slopes of the J1/2–V curve plots for each device are compared, and the mobility ratios between holes (μh) and electrons (μe) are obtained. In PDPPTBBT:Y6 thin films, the electron mobility was approximately 2.61-fold greater than the hole mobility; in PDPPTBBT:P3HT thin films, the hole mobility was about 8.32-fold greater than the electron mobility. These differences in the mobility between holes and electrons signify unbalanced charge-transport properties within the devices. Faster photogenerated carriers that are formed during device operation are directed toward the electrodes; slower carriers become trapped within the films. It has been previously reported that artificial creation of charge carrier traps within blended films triggered photomultiplication; trapped carriers induced band bending and influx of oppositely charged carriers from an electrode. (42,43) These devices thus exhibit photomultiplication, enhancing the performance at sufficiently high voltages.
3. Conclusions
Click to copy section linkSection link copied!
We synthesized an open-shell conjugated terpolymer from BBT, DPP, and thiophene units. By combining three monomers, we obtained an open-shell characteristic, enabling absorption of light beyond 1000 nm, and tuned the band gap for the OPDs. The diradical triplet ground state remained stable in the thin films. EPR and magnetic susceptibility analyses of PDPPTBBT revealed a dominant temperature-independent Pauli paramagnetism (χPauli) at 3.2 μemu g–1 Oe–1 of each repeating unit, indicating that the diradicals were delocalized in the solid state. Next, we utilized GIWAXS to examine the crystallinities of polymer and polymer-blended thin films after thermal annealing; the diradical stability was enhanced. Finally, we fabricated OPDs from PDPPTBBT:P3HT and PDPPTBBT:Y6 blends; the EQE values were 59 and 126% at 1050 nm; the D* values were 9.7 × 1010 and 7.5 × 1011 Jones, respectively. Moreover, we found that strong interchain interactions, enhanced by thermal annealing and material blending, support the stability of diradical species and enable photomultiplication effects over 1000 nm wavelength. Overall, we demonstrate the possibility of developing materials that detect wavelengths in the NIR region with a novel open-shell conjugated polymer that can be used to construct devices.
4. Experimental Section
Click to copy section linkSection link copied!
4.1. Materials
Anhydrous solvents, sodium metal, 2-methylbutan-2-ol, 2-thiophenecarbonitirile, diethyl succinate, potassium carbonate, 2-decyltetradecyl bromide, tetrakis(triphenylphosphine)palladium(0), potassium acetate, N-bromosuccinimide, and 2,5-bis(trimethylstannyl)thiophene were purchased from Sigma-Aldrich. 4,7-Dibromobenzo[1,2-c:4,5-c′]bis([1,2,5]thiadiazole) was obtained from J’s Science (South Korea). All chemicals were used without further purification.
4.2. Synthesis of PDPPTBBT
4.2.1. Synthesis of 3,6-Di(thiophen-2-yl)-2,5-dihydropyrrolo[3,4-c]pyrrole-1,4-dione (1)
Sodium metal (3.2 g, 138 mmol) was dissolved in 100 mL of 2-methylbutan-2-ol. The reaction mixture was refluxed at 110 °C until Na was completely consumed; the temperature was then lowered to 90 °C. 2-Thiophenecarbonitirile (8.0 mL, 85.8 mmol) was added to the reaction mixture in a single portion, and a solution of diethyl succinate (4.8 mL, 28.7 mmol) in 2-methylbutan-2-ol was added dropwise. After refluxing at 110 °C overnight, the reaction mixture was cooled, and a solution of glacial acetic acid in methanol was added. The resultant deep-purple precipitate was washed several times with methanol and dried under vacuum. (5 g, 60%) 1H NMR (400 MHz, CDCl3), δ (ppm): 11.20 (s, 2H), 8.18 (d, J, 2H), 7.90 (d, J, 2H), 7.27 (m, 2H).
4.2.2. Synthesis of 2,5-Bis(2-decyltetradecyl)-3,6-di(thiophen-2-yl)-2,5-dihydropyrrolo[3,4-c]pyrrole-1,4-dione (2)
Sodium metal (3.2 g, 138 mmol) was dissolved in 100 mL of 2-methylbutan-2-ol. The reaction mixture was refluxed at 110 °C until Na was completely consumed; the temperature was then lowered to 90 °C. 2-Thiophenecarbonitirile (8.0 mL, 85.8 mmol) was added to the reaction mixture in a single portion, and a solution of diethyl succinate (4.8 mL, 28.7 mmol) in 2-methylbutan-2-ol was added dropwise. After refluxing at 110 °C overnight, the reaction mixture was cooled, and a solution of glacial acetic acid in methanol was added. The resultant deep-purple precipitate was washed several times with methanol and dried under vacuum. (5 g, 60%) 1H NMR (400 MHz, CDCl3), δ (ppm): 11.20 (s, 2H), 8.18 (d, J, 2H), 7.90 (d, J, 2H), 7.27 (m, 2H).
4.2.3. Synthesis of 3,6-Bis(5-bromothiophen-2-yl)-2,5-bis(2-decyltetradecyl)-2,5-dihydropyrrolo[3,4-c]pyrrole-1,4-dione (3)
N-Bromosuccinimide (0.156 g, 1.40 mmol) was added to a solution of 2,5-bis(2-decyltetradecyl)-3,6-di(thiophen-2-yl)-2,5-dihydropyrrolo[3,4-c]pyrrole-1,4-dione (0.5 g, 0.51 mmol) in chloroform (15 mL), and the reaction vessel was wrapped in aluminum foil to exclude light. After the reaction was complete, excess methanol was added for quenching. The mixture was filtered, and the solid was washed with methanol (2 × 200 mL) and dried under a vacuum. The crude product was recrystallized from chloroform and methanol to give the final product (0.43 g, 76%) as a dark red solid. 1H NMR (CDCl3, 400 MHz) δ (ppm): 8.66 (d, 2H), 7.25 (d, 2H), 3.96 (t, 4H), 1.88 (m, 2H), 1.56–1.24 (m, 80H), 0.88 (t, 12H).
4.2.4. Polymerization of PDPPTBBT
A microwave tube was charged with 3,6-bis(5-bromothiophen-2-yl)-2,5-bis(2-decyltetradecyl)-2,5-dihydropyrrolo[3,4-c]pyrrole-1,4-dione (339 mg, 0.30 mmol), 2,5-bis(trimethylstannyl)thiophene (245 mg, 0.60 mmol), 4,7-dibromobenzo[1,2-c:4,5-c′]bis([1,2,5]thiadiazole) (105 mg, 0.30 mmol), and tetrakis(triphenylphosphine)palladium(0) under nitrogen. Then, anhydrous toluene (6 mL) was added, and the mixture was irradiated in a microwave oven at 120 °C for 30 min, 140 °C for 60 min, and 160 °C for 30 min. The mixture was then poured into methanol, and the resultant precipitate was dissolved in chloroform and filtered through Celite 545 to remove the undissolved metal catalyst. The polymer fibers were subjected to Soxhlet extraction in methanol, acetone, hexane, and chloroform. The final polymer was obtained after reprecipitation from methanol (150 mg, 41%).
4.3. Characterization
The 1H NMR (400 MHz) spectra were recorded using an AvanceII HD instrument with deuterated chloroform (CDCl3) as the solvent and tetramethylsilane as the internal standard. GPC analysis was conducted at 35 °C on a Spectra System P1000 platform with polystyrene as the standard and tetrahydrofuran as the eluent. Thermogravimetric analysis (TGA) was performed from 60 to 800 °C using a Pyris 6 instrument at a heating rate of 10 °C min–1 under a nitrogen flow. CV was performed with an IviumStat electrochemical workstation and a dry acetonitrile solution containing n-Bu4NPF6 (0.1 M); the scan rate was 50 mV/s at room temperature under nitrogen. A Pt disk (2 mm diameter), a Pt wire, and an Ag/AgCl electrode served as the working, counter, and reference electrodes, respectively. UV–vis absorption spectra were recorded using a Jasco V-770 spectrophotometer. AFM studies were carried out with an NX-10 instrument (Bruker) operating at ambient conditions in the Research Institute of Advanced Materials (RIAM) of Seoul National University. Samples were evaluated using the PLS-II 9A U-SAXS beamline of the Pohang Accelerator Laboratory (Korea) with monochromated X-rays of 11.08 keV (λ = 1.119 Å); an incidence angle of 0.10° and an exposure time of 5–10 s were used. Each sample was placed on a seven-axis motorized stage to allow for fine alignment. Patterns were recorded by a 2D charge-coupled device detector (SX165; Rayonix). The sample-to-detector distance was ∼220.9 mm. Magnetic data were acquired using a magnetic property measurement system (Quantum Design, US) installed at the Institute of Applied Physics of Seoul National University. The sample consisted of 12.77 mg of PDPPTBBT in a polycarbonate capsule. Prior to measurement, the background of the empty capsule was obtained. ZFC-FC susceptibility curves were collected from 2 to 300 K at an applied field of 10 kOe. Magnetization over the temperature range 2–100 K was measured as a function of the field (range ± 50 kOe). EPR spectra of both solution and thin film samples (including measurements at various temperatures) were obtained using an EMXplus-9.5/12/P/L platform (Bruker, USA) operating in the X-band at the National Center for Inter-University Research Facilities of Seoul National University. The EPR signal intensity was fitted to the Bleaney–Bowers eq (eq 1 and Figure 2h) from 10 to 293 K to determine the energy difference between the singlet and triplet states (ΔEST) using the following equation: (44,45)
(1)
(2)
where C is a constant and Jab is the intramolecular exchange coupling constant correlated with ΔEST.
4.4. Computational Methodology
All of the electronic structural calculations were performed with the Gaussian 09 software package. The molecular geometries of the electronic ground state and the first triplet state of the model dimer were optimized using DFT that incorporated the Becke three-parameter B3LYP functional and the 6-31G(d,p) basis set. The alkyl chains in the polymer backbone were substituted with methyl (−CH3) groups. It is known that the alkyl functional groups do not significantly affect either the ground-state geometry or the electronic structural properties.
4.5. Device Fabrication and Characterization
The OPD devices were fabricated as ITO/ZnO/active layer/MoO3/Au preparations. ITO-coated glass substrates were cleaned by sonication in a detergent, deionized water, acetone, and isopropyl alcohol. The ZnO layers were spin-coated onto ITO-coated glass substrates at 3000 rpm and then thermally annealed at 110 °C for 10 min in air. The PDPPTBBT:P3HT (1:1 wt/wt)-based blends and PDPPTBBT:Y6 (1:2 wt/wt)-based blends were dissolved in chloroform to final concentrations of 20 and 30 mg mL–1, respectively, and the solutions were stirred at room temperature overnight in a nitrogen-filled glovebox. The solutions were then spin-coated onto the ZnO films. Blends based on PDPPTBBT:P3HT were annealed at 230 °C for 30 min; blends based on PDPPTBBT:Y6 were annealed at 180 °C for 30 min. Next, MoO3 and Au were thermally deposited to thicknesses of 5 and 20 nm, respectively.
The current–voltage characteristics and real-time photoswitching behaviors of the OPDs were measured in a vacuum chamber using a Keithley 4200-SCS semiconductor parametric analyzer.
The responsivity (R) and EQE were calculated using the following equations:
(3)
(4)
where Ilight is the current under light, Idark is the current in the dark, Pinc is the incident illumination power, Pint is the incident light intensity, h is Planck’s constant, c is the speed of light, e is the charge of an electron, A is the active area, and λ is the wavelength. The specific detectivity (D*) is a key performance parameter that typically indicates the sensitivity to weak light, as follows:
(5)
(6)
where A is the effective area of the detector in cm2, B is the bandwidth, NEP is the noise equivalent power, and is the measured noise current. If shot noise from the dark current is the major noise that limits D*, D* can be simplified as
(7)
Supporting Information
Click to copy section linkSection link copied!
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.5c03911.
GPC, TGA, and 1H NMR data of the synthesized materials, as well as additional EPR data; 2D GIWAXS, AFM, SCLC, LDR, and response speed of the device; and energy diagram, J–V curve, EQE, responsivity, and detectivity of the PDPPTBBT:P3HT device (PDF)
- am5c03911_si_001.pdf (1.52 MB)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Author Information
Click to copy section linkSection link copied!
- Corresponding Author
Joon Hak Oh - Schoolof Chemical and Biological Engineering, Institute of Chemical Processes, Seoul National University, Seoul 08826, Republic of Korea;
https://orcid.org/0000-0003-0481-6069; Email: [emailprotected]
- Authors
Moon-Ki Jeong - Schoolof Chemical and Biological Engineering, Institute of Chemical Processes, Seoul National University, Seoul 08826, Republic of Korea;
https://orcid.org/0009-0009-4199-9336
Sang Hyuk Lee - Schoolof Chemical and Biological Engineering, Institute of Chemical Processes, Seoul National University, Seoul 08826, Republic of Korea;
https://orcid.org/0009-0009-4829-5120
Yousang Won - Schoolof Chemical and Biological Engineering, Institute of Chemical Processes, Seoul National University, Seoul 08826, Republic of Korea;
https://orcid.org/0009-0003-7125-7858
Jaeyong Ahn - Schoolof Chemical and Biological Engineering, Institute of Chemical Processes, Seoul National University, Seoul 08826, Republic of Korea; Present Address: Department of Chemical Engineering, Stanford University, Stanford, California 94305, United States;
https://orcid.org/0000-0002-0036-0300
Myeong In Kim - Departmentof Organic and Nano Engineering and Human-Tech Convergence Program, Hanyang University, Seoul 04763, Republic of Korea;
https://orcid.org/0009-0008-5996-6308
- Author Contributions
M.-K.J. and S.H.L. contributed equally to this work.
- Notes
The authors declare no competing financial interest.
Acknowledgments
Click to copy section linkSection link copied!
This work was supported by the National Research Foundation of Korea (NRF) grant (2023R1A2C3007715, RS-2024-00398065) through the NRF by the Ministry of Science and ICT (MSIT), Korea. The Institute of Engineering Research at Seoul National University provided research facilities for this work. This work was supported by Samsung Electronics Co., Ltd.
References
Click to copy section linkSection link copied!
This article references 45 other publications.
1
Chen, Y.; Zheng, Y.; Jiang, Y.; Fan, H.; Zhu, X. Carbon-Bridged 1,2-Bis(2-thienyl)ethylene: An Extremely Electron Rich Dithiophene Building Block Enabling Electron Acceptors with Absorption above 1000 nm for Highly Sensitive NIR Photodetectors. J. Am. Chem. Soc. 2021, 143, 4281, DOI: 10.1021/jacs.0c12818
Google Scholar
There is no corresponding record for this reference.
2
Wei, Y.; Chen, H.; Liu, T.; Wang, S.; Jiang, Y.; Song, Y.; Zhang, J.; Zhang, X.; Lu, G.; Huang, F. Self-Powered Organic Photodetectors with High Detectivity for Near Infrared Light Detection Enabled by Dark Current Reduction. Adv. Funct. Mater. 2021, 31, 2106326 DOI: 10.1002/adfm.202106326
Google Scholar
There is no corresponding record for this reference.
3
Zhong, Z.; Peng, F.; Huang, Z.; Ying, L.; Yu, G.; Huang, F.; Cao, Y. High-Detectivity Non-Fullerene Organic Photodetectors Enabled by a Cross-Linkable Electron Blocking Layer. ACS Appl. Mater. Interfaces 2020, 12, 45092, DOI: 10.1021/acsami.0c13833
Google Scholar
There is no corresponding record for this reference.
4
Vitorino, R.; Barros, A. S.; Guedes, S.; Caixeta, D. C.; Silva, R. S. Diagnostic and monitoring applications using near infrared (NIR) spectroscopy in cancer and other diseases. Photodiagnosis Photodyn. Ther. 2023, 42, 103633 DOI: 10.1016/j.pdpdt.2023.103633
Google Scholar
There is no corresponding record for this reference.
5
Son, C.; Kim, J.; Kang, D.; Park, S.; Ryu, C.; Baek, D.; Jeong, G.; Jeong, S.; Ahn, S.; Lim, C.; Jeong, Y.; Eom, J.; Park, J.-H.; Lee, D. W.; Kim, D.; Kim, J.; Ko, H.; Lee, J. Behavioral biometric optical tactile sensor for instantaneous decoupling of dynamic touch signals in real time. Nat. Commun. 2024, 15, 8003, DOI: 10.1038/s41467-024-52331-4
Google Scholar
There is no corresponding record for this reference.
6
Li, L.; Wang, D.; Zhang, D.; Ran, W.; Yan, Y.; Li, Z.; Wang, L.; Shen, G. Near-Infrared Light Triggered Self-Powered Mechano-Optical Communication System using Wearable Photodetector Textile. Adv. Funct. Mater. 2021, 31, 2104782 DOI: 10.1002/adfm.202104782
Google Scholar
There is no corresponding record for this reference.
7
Jeong, M.-K.; Kang, J.; Park, D.; Yim, S.; Jung, I. H. A conjugated polyelectrolyte interfacial modifier for high performance near-infrared quantum-dot photodetectors. J. Mater. Chem. C 2020, 8, 2542, DOI: 10.1039/C9TC05541J
Google Scholar
There is no corresponding record for this reference.
8
Maity, P.; Singh, S. V.; Biring, S.; Pal, B. N.; Ghosh, A. K. Selective near-infrared (NIR) photodetectors fabricated with colloidal CdS:Co quantum dots. J. Mater. Chem. C 2019, 7, 7725, DOI: 10.1039/C9TC01871A
Google Scholar
There is no corresponding record for this reference.
9
Yao, Y.; Liang, Y.; Shrotriya, V.; Xiao, S.; Yu, L.; Yang, Y. Plastic Near-Infrared Photodetectors Utilizing Low Band Gap Polymer. Adv. Mater. 2007, 19, 3979, DOI: 10.1002/adma.200602670
Google Scholar
There is no corresponding record for this reference.
10
Chen, L. L.; Zhao, L.; Wang, Z. G.; Liu, S. L.; Pang, D. W. Near-Infrared-II Quantum Dots for In Vivo Imaging and Cancer Therapy. Small 2022, 18, 2104567 DOI: 10.1002/smll.202104567
Google Scholar
There is no corresponding record for this reference.
11
Vasilopoulou, M.; Kim, H. P.; Kim, B. S.; Papadakis, M.; Ximim Gavim, A. E.; Macedo, A. G.; da Silva, W. J.; Schneider, F. K.; Mat Teridi, M. A.; Coutsolelos, A. G. Efficient colloidal quantum dot light-emitting diodes operating in the second near-infrared biological window. Nat. Photonics 2020, 14, 50, DOI: 10.1038/s41566-019-0526-z
Google Scholar
There is no corresponding record for this reference.
12
Xu, K.; Zhou, W.; Ning, Z. Integrated Structure and Device Engineering for High Performance and Scalable Quantum Dot Infrared Photodetectors. Small 2020, 16, 2003397 DOI: 10.1002/smll.202003397
Google Scholar
There is no corresponding record for this reference.
13
Chow, P. C.; Someya, T. Organic Photodetectors for Next-Generation Wearable Electronics. Adv. Mater. 2020, 32, 1902045 DOI: 10.1002/adma.201902045
Google Scholar
There is no corresponding record for this reference.
14
Ren, H.; Chen, J. D.; Li, Y. Q.; Tang, J. X. Recent Progress in Organic Photodetectors and their Applications. Adv. Sci. 2021, 8, 2002418 DOI: 10.1002/advs.202002418
Google Scholar
There is no corresponding record for this reference.
15
Zhao, Z.; Xu, C.; Niu, L.; Zhang, X.; Zhang, F. Recent Progress on Broadband Organic Photodetectors and their Applications. Laser Photonics Rev. 2020, 14, 2000262 DOI: 10.1002/lpor.202000262
Google Scholar
There is no corresponding record for this reference.
16
Cheng, P.; Zhan, X. Stability of Organic Solar Cells: Challenges and Strategies. Chem. Soc. Rev. 2016, 45, 2544– 2582, DOI: 10.1039/C5CS00593K
Google Scholar
There is no corresponding record for this reference.
17
Xu, J.; Zhang, Y.; Liu, J.; Wang, L. NIR-II Absorbing Monodispersed Oligomers Based on N–B←N Unit. Angew. Chem., Int. Ed. 2023, 62, e202310838 DOI: 10.1002/anie.202310838
Google Scholar
There is no corresponding record for this reference.
18
Zhang, Y.; Yu, Y.; Liu, X.; Miao, J.; Han, Y.; Liu, J.; Wang, L. An n-Type All-Fused-Ring Molecule with Photoresponse to 1000 nm for Highly Sensitive Near-Infrared Photodetector. Adv. Mater. 2023, 35, 2211714 DOI: 10.1002/adma.202211714
Google Scholar
There is no corresponding record for this reference.
19
Wu, J.-S.; Cheng, S.-W.; Cheng, Y.-J.; Hsu, C.-S. Donor–acceptor conjugated polymers based on multifused ladder-type arenes for organic solar cells. Chem. Soc. Rev. 2015, 44, 1113, DOI: 10.1039/C4CS00250D
Google Scholar
There is no corresponding record for this reference.
20
Scharber, M. C.; Sariciftci, N. S. Low Band Gap Conjugated Semiconducting Polymers. Adv. Mater. Technol. 2021, 6, 2000857 DOI: 10.1002/admt.202000857
Google Scholar
There is no corresponding record for this reference.
21
Kim, M.; Ryu, S. U.; Park, S. A.; Choi, K.; Kim, T.; Chung, D.; Park, T. Donor–Acceptor-Conjugated Polymer for High-Performance Organic Field-Effect Transistors: A Progress Report. Adv. Funct. Mater. 2020, 30, 1904545 DOI: 10.1002/adfm.201904545
Google Scholar
There is no corresponding record for this reference.
22
Dou, L.; Liu, Y.; Hong, Z.; Li, G.; Yang, Y. Low-Bandgap Near-IR Conjugated Polymers/Molecules for Organic Electronics. Chem. Rev. 2015, 115, 12633, DOI: 10.1021/acs.chemrev.5b00165
Google Scholar
There is no corresponding record for this reference.
23
Hashemi, D.; Ma, X.; Ansari, R.; Kim, J.; Kieffer, J. Design principles for the energy level tuning in donor/acceptor conjugated polymers. Phys. Chem. Chem. Phys. 2019, 21, 789, DOI: 10.1039/C8CP03341B
Google Scholar
There is no corresponding record for this reference.
24
Adegoke, O. O.; Jung, I. H.; Orr, M.; Yu, L.; Goodson, T., III Effect of Acceptor Strength on Optical and Electronic Properties in Conjugated Polymers for Solar Applications. J. Am. Chem. Soc. 2015, 137, 5759, DOI: 10.1021/ja513002h
Google Scholar
There is no corresponding record for this reference.
25
Cai, H.; Tang, H.; Wang, T.; Xu, C.; Xie, J.; Fu, M.; Luo, X.; Hu, Z.; Zhang, Y.; Deng, Y.; Li, G.; Liu, C.; Huang, F.; Cao, Y. An n-Type Open-Shell Conjugated Polymer with High-Spin Ground-State and High Intrinsic Electrical Conductivity. Angew. Chem., Int. Ed. 2024, 63, e202402375 DOI: 10.1002/anie.202402375
Google Scholar
There is no corresponding record for this reference.
26
Mayer, K. S.; Adams, D. J.; Eedugurala, N.; Lockart, M. M.; Mahalingavelar, P.; Huang, L.; Galuska, L. A.; King, E. R.; Gu, X.; Bowman, M. K. Topology and ground state control in open-shell donor-acceptor conjugated polymers. Cell Rep. Phys. Sci. 2021, 2, 100467 DOI: 10.1016/j.xcrp.2021.100467
Google Scholar
There is no corresponding record for this reference.
27
Yuen, J. D.; Fan, J.; Seifter, J.; Lim, B.; Hufschmid, R.; Heeger, A. J.; Wudl, F. High Performance Weak Donor–Acceptor Polymers in Thin Film Transistors: Effect of the Acceptor on Electronic Properties, Ambipolar Conductivity, Mobility, and Thermal Stability. J. Am. Chem. Soc. 2011, 133, 20799– 20807, DOI: 10.1021/ja205566w
Google Scholar
There is no corresponding record for this reference.
28
Liu, C.-T.; Vella, J.; Eedugurala, N.; Mahalingavelar, P.; Bills, T.; Salcido-Santacruz, B.; Sfeir, M. Y.; Azoulay, J. D. Ultrasensitive Room Temperature Infrared Photodetection Using a Narrow Bandgap Conjugated Polymer. Adv. Sci. 2023, 10, 2304077 DOI: 10.1002/advs.202304077
Google Scholar
There is no corresponding record for this reference.
29
Kim, M. I.; Kang, J.; Park, J.; Jeong, W.; Kim, J.; Yim, S.; Jung, I. H. Improvement of Dynamic Performance and Detectivity in Near-Infrared Colloidal Quantum Dot Photodetectors by Incorporating Conjugated Polymers. Molecules 2022, 27, 7660, DOI: 10.3390/molecules27217660
Google Scholar
There is no corresponding record for this reference.
30
Tu, X.; Yao, X.; Li, Y.; Zhang, S.; He, Z.; Zhang, C.; Liu, Z.; Zhong, H.; Fei, Z. Fluorinated Conjugated Polymers Enabled Enhanced Detectivity in Organic Near-Infrared Photodetectors. ACS Appl. Polym. Mater. 2024, 6, 5970– 5979, DOI: 10.1021/acsapm.4c00663
Google Scholar
There is no corresponding record for this reference.
31
He, J.; Wang, Z.; Gao, Y.; Xiong, B.; Shao, M.; Li, Z. Ladder-Type π-Conjugated Polysquaraines for Near-Infrared Organic Photodetectors. Macromolecules 2024, 57, 6123– 6132, DOI: 10.1021/acs.macromol.4c00791
Google Scholar
There is no corresponding record for this reference.
32
Jacoutot, P.; Scaccabarozzi, A. D.; Nodari, D.; Panidi, J.; Qiao, Z.; Schiza, A.; Nega, A. D.; Dimitrakopoulou-Strauss, A.; Gregoriou, V. G.; Heeney, M.; Chochos, C. L.; Bakulin, A. A.; Gasparini, N. Enhanced sub–1 eV detection in organic photodetectors through tuning polymer energetics and microstructure. Sci. Adv. 2023, 9, eadh2694 DOI: 10.1126/sciadv.adh2694
Google Scholar
There is no corresponding record for this reference.
33
Joo, Y.; Huang, L.; Eedugurala, N.; London, A. E.; Kumar, A.; Wong, B. M.; Boudouris, B. W.; Azoulay, J. D. Thermoelectric Performance of an Open-Shell Donor–Acceptor Conjugated Polymer Doped with a Radical-Containing Small Molecule. Macromolecules 2018, 51, 3886, DOI: 10.1021/acs.macromol.8b00582
Google Scholar
There is no corresponding record for this reference.
34
Steelman, M. E.; Adams, D. J.; Mayer, K. S.; Mahalingavelar, P.; Liu, C.-T.; Eedugurala, N.; Lockart, M.; Wang, Y.; Gu, X.; Bowman, M. K.; Azoulay, J. D. Magnetic Ordering in a High-Spin Donor–Acceptor Conjugated Polymer. Adv. Mater. 2022, 34, 2206161 DOI: 10.1002/adma.202206161
Google Scholar
There is no corresponding record for this reference.
35
Vrankić, M.; Šarić, A.; Bosnar, S.; Pajić, D.; Dragović, J.; Altomare, A.; Falcicchio, A.; Popović, J.; Jurić, M.; Petravić, M. Magnetic oxygen stored in quasi-1D form within BaAl2O4 lattice. Sci. Rep. 2019, 9, 15158, DOI: 10.1038/s41598-019-51653-4
Google Scholar
There is no corresponding record for this reference.
36
Mugiraneza, S.; Hallas, A. M. Tutorial: a beginner’s guide to interpreting magnetic susceptibility data with the Curie-Weiss law. Commun. Phys. 2022, 5, 95, DOI: 10.1038/s42005-022-00853-y
Google Scholar
There is no corresponding record for this reference.
37
Yumoto, S.; Katsumata, J.; Osawa, F.; Wada, Y.; Suzuki, K.; Kaji, H.; Marumoto, K. Operando ESR observation in thermally activated delayed fluorescent organic light-emitting diodes. Sci. Rep. 2023, 13, 11109, DOI: 10.1038/s41598-023-38063-3
Google Scholar
There is no corresponding record for this reference.
38
Chen, Z.; Li, W.; Zhang, Y.; Wang, Z.; Zhu, W.; Zeng, M.; Li, Y. Aggregation-Induced Radical of Donor–Acceptor Organic Semiconductors. J. Phys. Chem. Lett. 2021, 12, 9783, DOI: 10.1021/acs.jpclett.1c02463
Google Scholar
There is no corresponding record for this reference.
39
Wang, X.-Q.; Song, C.; Lei, T. Open-Shell Oligomers and Polymers: Theory, Characterization Methods, Molecular Design, and Applications. Chin. J. Polym. Sci. 2024, 42, 417– 436, DOI: 10.1007/s10118-024-3087-7
Google Scholar
There is no corresponding record for this reference.
40
Wang, S.; Chen, H.; Liu, T.; Wei, Y.; Yao, G.; Lin, Q.; Han, X.; Zhang, C.; Huang, H. Retina-Inspired Organic Photonic Synapses for Selective Detection of SWIR Light. Angew. Chem., Int. Ed. 2023, 62, e202213733 DOI: 10.1002/anie.202213733
Google Scholar
There is no corresponding record for this reference.
41
SajediAlvar, M.; Blom, P. W. M.; Wetzelaer, G.-J. A. H. Space-charge-limited electron and hole currents in hybrid organic-inorganic perovskites. Nat. Commun. 2020, 11, 4023, DOI: 10.1038/s41467-020-17868-0
Google Scholar
There is no corresponding record for this reference.
42
Xing, S.; Kublitski, J.; Hänisch, C.; Winkler, L. C.; Li, T.-Y.; Kleemann, H.; Benduhn, J.; Leo, K. Photomultiplication-Type Organic Photodetectors for Near-Infrared Sensing with High and Bias-Independent Specific Detectivity. Adv. Sci. 2022, 9, 2105113 DOI: 10.1002/advs.202105113
Google Scholar
There is no corresponding record for this reference.
43
Miao, J.; Zhang, F. Recent Progress on Photomultiplication Type Organic Photodetectors. Laser Photonics Rev. 2019, 13, 1800204 DOI: 10.1002/lpor.201800204
Google Scholar
There is no corresponding record for this reference.
44
Bleaney, B.; Bowers, K. D. Anomalous paramagnetism of copper acetate. Proc. R. Soc. London A Math. Phys. Sci. 1952, 214, 451, DOI: 10.1098/rspa.1952.0181
Google Scholar
There is no corresponding record for this reference.
45
London, A. E.; Chen, H.; Sabuj, M. A.; Tropp, J.; Saghayezhian, M.; Eedugurala, N.; Zhang, B. A.; Liu, Y.; Gu, X.; Wong, B. M.; Rai, N.; Bowman, M. K.; Azoulay, J. D. A high-spin ground-state donor-acceptor conjugated polymer. Sci. Adv. 2019, 5, eaav2336 DOI: 10.1126/sciadv.aav2336
Google Scholar
There is no corresponding record for this reference.
Cited By
Click to copy section linkSection link copied!
This article has not yet been cited by other publications.
Download PDF