Fact-checked by Grok 2 weeks ago

Organic electronics

Organic electronics is a branch of that utilizes carbon-based organic materials, such as conjugated polymers and small molecules, as semiconductors to fabricate devices and components. These materials, which conduct through delocalized π-electrons in their molecular structures, enable the creation of flexible, , transparent, and low-cost devices compared to traditional silicon-based inorganic . The field leverages solution-processable thin-film deposition techniques, allowing large-area fabrication on flexible substrates and opening avenues for applications in displays, , and sensors. The origins of organic electronics trace back to the mid-20th century, with early investigations into the electrical properties of organic compounds beginning in the through studies on molecular crystals. Interest surged in the and following discoveries of in organic crystals and the conductivity of doped polymers, such as , which demonstrated metallic-like conduction upon doping. By the 1980s and 1990s, breakthroughs like the invention of efficient organic light-emitting diodes (OLEDs) by Ching W. Tang and Steven Van Slyke in 1987 and the development of polymer LEDs by Richard Friend's group in 1990 propelled the field forward, leading to commercial applications. These milestones established organic materials as viable alternatives for optoelectronic devices, with ongoing research emphasizing molecular design for tuned electronic properties. Key materials in organic electronics include p-type and n-type like pentacene, rubrene, and polythiophenes, which exhibit mobilities ranging from 0.1 to over 10 cm² V⁻¹ s⁻¹ in highly ordered films. Devices such as OLEDs power modern flexible displays and lighting, organic (OPVs) achieve power conversion efficiencies exceeding 20% (as of 2025) for , and organic thin-film transistors (OTFTs) enable bendable circuits and sensors. The and ionic conductivity of certain organic semiconductors have also extended the field into bioelectronics, interfacing with biological systems for neural probes and biosensors. Recent advances focus on improving device performance through techniques like molecular templating for ordered thin films and doping strategies to enhance up to 1 S cm⁻¹, enabling high-frequency operations in organic transistors reaching GHz speeds. efforts incorporate green solvents for processing, reducing environmental impact while maintaining high efficiencies in OPVs and . These developments underscore organic electronics' potential to complement in wearable, printed, and biomedical applications, with the global market alone surpassing tens of billions in value as of 2025.

Fundamentals

Organic Semiconductors

Organic semiconductors are carbon-based materials characterized by conjugated π-electron systems, which consist of alternating single and double bonds along the molecular backbone, enabling delocalization of electrons and semiconducting behavior. These materials exhibit electrical conductivity between that of insulators and metals, typically due to the presence of π-orbitals that facilitate generation upon excitation or doping. Unlike inorganic semiconductors, organic variants often feature molecular structures such as alternant hydrocarbons (e.g., polyacenes like pentacene, where carbon atoms form even-numbered rings with paired π-electrons) and donor-acceptor systems, where electron-donating and withdrawing groups are linked via conjugated bridges to enhance electron delocalization and tune optoelectronic properties. The semiconducting properties arise primarily from the energy difference between the highest occupied (HOMO) and the lowest unoccupied (LUMO), known as the HOMO-LUMO gap or bandgap (E_g), which determines the energy required to excite electrons from the to the conduction band. In , this bandgap typically ranges from 1 to 3 eV, corresponding to absorption in the visible to near-infrared spectrum and enabling applications in . The HOMO represents the energy level of the highest-energy occupied π-electrons, while the LUMO is the lowest-energy unoccupied π*-orbital; the spatial overlap and delocalization of these orbitals across the lower the bandgap compared to non-conjugated organics. Organic semiconductors are classified by their predominant charge carrier type: p-type (hole-transporting), n-type (electron-transporting), or ambipolar (both). P-type materials, such as pentacene, have a HOMO level closer to the vacuum level, facilitating hole injection and transport upon oxidation. N-type materials, exemplified by fullerene derivatives like PCBM ([6,6]-phenyl-C61-butyric acid methyl ester), feature a LUMO level well-aligned for acceptance and stability under reduction. Ambipolar organics combine structural motifs from both types, often through donor-acceptor architectures, allowing balanced transport of electrons and holes under appropriate bias. Doping in introduces impurities to increase concentration, primarily through chemical methods involving oxidation (p-type doping) or reduction (n-type doping), which generate free carriers by partial charge transfer from/to dopant molecules. For p-type doping, acceptors (e.g., F4-TCNQ) abstract electrons from the host, creating holes and shifting the (E_F) toward the HOMO, thereby enhancing . The effect on carrier concentration follows from charge neutrality and ; assuming complete ionization of dopants and non-degenerate statistics (Boltzmann approximation, valid when E_F is several kT away from band edges), the hole concentration p ≈ N_A, where N_A is the acceptor density. The density of states in the valence band (HOMO) is N_V, leading to p = N_V exp[(E_HOMO - E_F)/kT], or rearranged, E_F = E_HOMO - kT ln(N_V / N_A). Derivation starts from the Fermi-Dirac distribution approximated as f(E) ≈ exp[(E_F - E) /kT] for E < E_F (holes in valence band), integrated over the g(E) ≈ N_V δ(E - E_HOMO) for simplified molecular orbital model, yielding p = ∫ g(E) [1 - f(E)] dE ≈ N_V exp[(E_HOMO - E_F)/kT]. Assumptions include: (1) uniform doping without clustering, (2) negligible dopant ionization energy, (3) no trapping or compensation by unintentional impurities, and (4) temperature T such that kT << E_g (room temperature typical). For n-type, the symmetric reduction shifts E_F toward LUMO analogously. Characterization of organic semiconductors relies on spectroscopic techniques to probe energy levels and bandgap. Ultraviolet-visible (UV-Vis) absorption spectroscopy measures the optical bandgap by identifying the onset of π-π* transitions, corresponding to the HOMO-LUMO gap (E_g^opt ≈ 1240 / λ_onset in eV, where λ_onset is the absorption edge wavelength), providing insight into conjugation length and delocalization. Cyclic voltammetry (CV) determines HOMO and LUMO positions electrochemically; the oxidation onset potential (E_ox) relates to HOMO via E_HOMO = - (E_ox + 4.8) eV (vs. vacuum, assuming ferrocene reference), and reduction onset (E_red) to LUMO as E_LUMO = - (E_red + 4.8) eV, enabling precise alignment predictions for devices. These methods complement each other, as CV yields formal potentials while UV-Vis captures excitonic effects in the solid state.

Charge Transport Mechanisms

In organic semiconductors, charge transport mechanisms differ markedly from those in inorganic counterparts due to the predominance of weak van der Waals intermolecular forces and inherent structural disorder, which localize charge carriers on individual molecules or segments rather than allowing delocalized band states. In inorganic semiconductors such as , carriers propagate via coherent band transport within extended Bloch states, yielding high mobilities often exceeding 1000 cm²/V·s at room temperature. By contrast, organic systems exhibit disorder-dominated transport, where charges reside in localized states associated with the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) levels, and movement occurs primarily through thermally assisted hopping between these sites, resulting in mobilities typically ranging from 0.1 to 10 cm²/V·s. This hopping nature arises from the low dielectric constants (ε ≈ 2–4) and amorphous or polycrystalline morphologies common in organic films, which promote energetic and spatial disorder. The prevailing model for charge transport in disordered organic semiconductors is variable-range hopping (VRH), originally formulated by Mott for systems with localized states near the Fermi level. In VRH, carriers do not hop to the nearest neighbor but select optimal paths that balance the thermal energy required for activation (overcoming site energy differences) with the tunneling distance (shorter hops require higher energy barriers). The DC conductivity σ in three dimensions is described by Mott's equation: \sigma = \sigma_0 \exp\left[ -\left( \frac{T_0}{T} \right)^{1/4} \right] where σ₀ is a temperature-independent prefactor related to wavefunction overlap and attempt frequency (typically 10²–10⁵ S/cm), T is the absolute temperature, and T₀ is a characteristic temperature given by T₀ = 18 / [k_B N(E_F) ξ³], with k_B the Boltzmann constant, N(E_F) the density of states at the Fermi level (often ~10¹⁹–10²¹ states/cm³ eV for organics), and ξ the localization length (≈3–10 Å, reflecting the extent of molecular wavefunctions). This model predicts a sublinear temperature dependence, distinguishing it from nearest-neighbor hopping (exponential with T^{-1/2}), and is validated across a wide range of organic materials at low temperatures (T < 200 K), where disorder broadens the density of states into a Gaussian distribution. In highly ordered systems, such as molecular single crystals, band-like transport can emerge, approximating delocalized carrier motion akin to inorganic crystals but adapted to the molecular framework. Here, the Drude model provides a suitable description, with mobility μ expressed as μ = e τ / m^, where e is the elementary charge, τ the mean scattering time (limited by phonons or impurities), and m the effective mass (typically 1–10 m_e for organics, larger than in silicon due to narrow bandwidths from weak overlap integrals). For instance, in rubrene single crystals, room-temperature hole mobilities reach ~20 cm²/V·s, evidencing coherent transport along high-symmetry directions with minimal disorder. However, even in these cases, transport often transitions to activated hopping at higher temperatures or under strain, highlighting the fragility of band-like behavior in organics. A key feature influencing transport across both regimes is polaron formation, where an injected charge carrier couples strongly to lattice vibrations, causing self-localization and a surrounding distortion that increases effective mass and scattering. In the adapted to organics, the polaron binding energy E_p ≈ (1/2) k ω_ph, with k the dimensionless electron-phonon coupling constant (≈1–5 for typical organics) and ω_ph the characteristic phonon frequency (≈100–200 meV for intramolecular modes), yields activation barriers of 0.05–0.3 eV. This localization reduces mobility by factors of 10–100 compared to non-polaronic band transport in , as polarons hop as quasi-particles rather than free carriers. Several factors modulate these mechanisms: temperature dependence shows activated behavior in hopping regimes (positive dμ/dT) but can exhibit negative dependence in band-like cases due to enhanced phonon scattering; trap states, arising from impurities or structural defects, create deep energy wells (0.2–0.5 eV) that capture carriers and lower effective mobility by orders of magnitude; and anisotropy in thin films or crystals leads to directional variations, with transport enhanced along π-π stacking axes (up to 10× higher than perpendicular). Doping increases carrier density but can exacerbate trapping if not optimized. Charge transport parameters are commonly extracted using time-of-flight (TOF) photoconductivity, a transient technique that illuminates a thin film (typically 1–10 μm) with a short laser pulse to generate a sheet of carriers, which then drift across the sample under an applied electric field E. The transit time t_tr is measured from the photocurrent profile, yielding mobility via μ = L / (t_tr E), where L is the layer thickness; this method isolates bulk transport, avoiding contact effects, and reveals dispersive (disorder-broadened) vs. non-dispersive behavior. TOF measurements confirm VRH and polaronic signatures across diverse organic systems, providing benchmarks for theoretical models.

History

Discovery of Conductive Polymers

Prior to the 1970s, polymers were widely regarded as electrical insulators with conductivities typically below 10^{-10} S/cm, limiting their use in electronic applications. This perception shifted dramatically in 1977 when Hideki Shirakawa accidentally discovered a metallic form of polyacetylene during synthesis experiments at the University of Tsukuba. Using a Ziegler-Natta catalyst with a lower-than-intended concentration of aluminum alkyl, Shirakawa produced shiny, silvery films of cis-polyacetylene instead of the expected powder; heating these films to 150–200°C induced trans-cis isomerization, yielding stable trans-polyacetylene films with a metallic luster and conductivity around 10^{-5} S/cm, still semiconducting but far exceeding typical insulators. The breakthrough in achieving high conductivity came later that year through collaboration between Shirakawa, Alan J. Heeger, and Alan G. MacDiarmid at the University of Pennsylvania. Exposing trans-polyacetylene films to iodine vapor introduced p-type doping via oxidation, creating charge carriers and dramatically enhancing conductivity—up to approximately 10^3 S/cm initially, with optimized stretched films later reaching 10^5 S/cm, comparable to metals like copper. This doping process involved partial oxidation of the polymer backbone, forming a polyacetylene-iodine complex that enabled metallic-like conduction without altering the film's structure significantly. The underlying mechanism stems from the π-conjugation along the polyacetylene backbone, where alternating single and double bonds allow delocalized π-electron states, potentially forming a half-filled band suggestive of metallicity. However, in the undoped state, a Peierls distortion—a lattice instability—doubles the unit cell by alternating bond lengths, opening a band gap of about 1.5–1.8 eV and rendering the material semiconducting; doping pins the Fermi level within this gap, introducing mobile charge carriers for conduction. Early doped polyacetylene films faced significant challenges, including instability in air and moisture, where exposure led to rapid dedoping and oxidation, degrading conductivity within hours due to the sensitivity of the iodine-polyacetylene complex. Despite these limitations, the discovery sparked exploration of conductive polymers as lightweight alternatives to metals for applications like wiring and shielding, though practical devices remained elusive at the time. The pioneering work of , , and was recognized with the 2000 Nobel Prize in Chemistry for the discovery and development of conductive polymers.

Development of Charge Transfer Complexes

The development of charge transfer complexes marked an early milestone in organic electronics, predating the advent of conductive polymers and focusing on discrete molecular systems that exhibited enhanced conductivity through partial electron transfer between donor and acceptor molecules. In 1954, Hideo Akamatu, Hiroo Inokuchi, and Yoshio Matsunaga reported the discovery of semiconducting behavior in the perylene-iodine complex, achieving a conductivity of approximately 10^{-3} S/cm, which was remarkably high for organic materials at the time and attributed to intermolecular charge transfer facilitated by iodine acting as an electron acceptor. This finding stimulated further exploration of halogen-doped polycyclic aromatic hydrocarbons, revealing that such complexes could bridge the gap between insulators and metals by forming mixed-valence states that delocalize charge carriers along molecular stacks. A pivotal advancement occurred in 1973 when John P. Ferraris, Dwaine O. Cowan, and colleagues synthesized (TTF-TCNQ), the first organic material to display metallic conductivity at room temperature, reaching values around 10^3 S/cm along the molecular chains. In this segregated stack structure, TTF molecules form donor chains while TCNQ forms acceptor chains, enabling partial charge transfer (approximately 0.5 electrons per donor-acceptor pair) that partially fills the conduction band and promotes band-like transport. This metallic behavior persists above a temperature around 54 K, below which a charge density wave instability opens a band gap, transitioning the system to a semiconducting state; the partial filling and one-dimensional stacking were key to achieving quasi-metallic properties without full ionization. Building on these insights, the pursuit of even higher conductivities led to the synthesis of tetramethyltetraselenafulvalenium (TMTSF)-based salts, known as Bechgaard salts, in the late 1970s. In 1980, Denis Jérome, Alain Mazaud, Michel Ribault, and Klaus Bechgaard observed superconductivity in (TMTSF)_2PF_6 at 1.2 K under a modest pressure of 0.9 GPa, marking the first instance of superconductivity in an organic material and demonstrating that quasi-one-dimensional charge transfer systems could support Cooper pair formation despite strong electron correlations. These salts featured similar segregated stack architectures but with selenium atoms enhancing interchain coupling, which suppressed the and enabled a superconducting ground state. Despite these breakthroughs, charge transfer complexes faced significant limitations, including chemical instability due to reactivity of radical ions and conductivities orders of magnitude lower than inorganic metals (typically <10^5 S/cm versus 10^7 S/cm for copper), which restricted practical applications. Nonetheless, their ability to mimic metallic and even superconducting behavior inspired subsequent doping strategies in extended conjugated systems and facilitated early prototypes of organic photovoltaics, where donor-acceptor interfaces in complexes like perylene derivatives were explored for photoinduced charge separation.

Key Milestones in Device Innovation

The development of organic electronics transitioned from fundamental material discoveries to practical device innovations in the late 1980s, with the demonstration of the first efficient organic light-emitting diode (OLED) by Ching W. Tang and Steven Van Slyke at Eastman Kodak in 1987. This bilayer device utilized tris(8-hydroxyquinoline)aluminum (Alq3) as the emissive layer, achieving a luminance efficiency of approximately 1 lm/W and external quantum efficiency of 1%, marking a pivotal shift toward viable electroluminescent displays through vacuum-deposited small molecules. Building on this, the 1990s saw the introduction of solution-processable polymer OLEDs, exemplified by the work of J. H. Burroughes and colleagues at the University of Cambridge in 1990, who fabricated the first polymer-based electroluminescent device using poly(p-phenylene vinylene) (PPV). This single-layer structure emitted greenish-yellow light at brightness levels exceeding 100 cd/m² under low voltage, enabling large-area fabrication via spin-coating and inkjet printing, which laid the groundwork for scalable, flexible displays. Parallel advancements in photovoltaics occurred in 1995, when Gang Yu and coworkers at the University of California, Santa Barbara, reported the first efficient bulk heterojunction organic solar cell using a blend of poly[2-methoxy-5-(2'-ethylhexyloxy)-1,4-phenylene vinylene] () as the donor and [6,6]-phenyl-C61-butyric acid methyl ester () as the acceptor. This morphology, formed by phase separation in the active layer, achieved a power conversion efficiency () of around 1% under AM1.5 illumination, overcoming prior limitations of planar junctions by providing extended donor-acceptor interfaces for exciton dissociation. In the realm of transistors, the late 1990s brought breakthroughs in organic field-effect transistors (OFETs), highlighted by Henry Sirringhaus and team in 1999, who demonstrated high-mobility charge transport in regioregular poly(3-hexylthiophene) (P3HT) films on flexible substrates. These devices exhibited field-effect mobilities exceeding 0.1 cm²/V·s, enabling the first flexible OFETs suitable for low-cost, bendable electronics and integrated circuits. The 2000 Nobel Prize in Chemistry, awarded to , , and for the discovery of conductive polymers, profoundly influenced device innovation by accelerating research into scalable organic electronics, leading to enhanced charge transport models and the integration of polymers into commercial prototypes that improved efficiency and flexibility in subsequent decades. Commercial milestones emerged in the early 2000s, with Sony unveiling an 11-inch full-color television prototype in 2007, showcasing high-contrast imaging on a rigid substrate and paving the way for consumer adoption of organic emissive displays. The 2010s witnessed the rise of flexible displays, driven by advancements in polymer and barrier encapsulation, enabling foldable smartphones and wearable devices from companies like Samsung, which commercialized curved AMOLED screens by 2013. Recent progress through 2025 has focused on enhancing photovoltaic efficiencies, with non-fullerene acceptors like Y6—introduced by Jianhui Hou and colleagues in 2019—enabling organic solar cells to surpass 20% PCE when paired with donors such as PM6, due to improved charge separation and reduced voltage losses in bulk heterojunction architectures. Furthermore, hybrid perovskite-organic tandem solar cells achieved certified efficiencies of 24.5% by 2023, reaching 26.4% as of June 2025, leveraging wide-bandgap perovskites atop narrow-bandgap organic subcells for broader spectral absorption and higher overall performance.

Materials

Conducting Polymers

Conducting polymers represent a vital class of materials in organic electronics, characterized by their conjugated backbone structures that enable electrical conductivity while retaining the mechanical advantages of polymers. These materials are broadly classified into intrinsic and extrinsic types. Intrinsic conducting polymers derive their conductivity directly from the delocalized π-electrons along the polymer chain without requiring additional dopants, exemplified by undoped , which exhibits semiconductor-like behavior with conductivities around 10^{-5} S/cm. Extrinsic conducting polymers, in contrast, achieve enhanced conductivity through doping, as seen in , , and , where charge carriers are introduced to bridge the band gap. The seminal discovery of conductivity in by , , and in 1977, recognized with the , laid the foundation for this class by demonstrating that doping could transform insulating polymers into metallic conductors. Synthesis of conducting polymers employs several methods to tailor their structure and properties. Chemical oxidation polymerization, often using oxidants like FeCl_3 for , allows bulk production of doped polymers in solution, yielding materials with controlled oxidation states during formation. Electrochemical polymerization, applied to monomers such as or , deposits thin films directly onto electrodes, enabling precise control over thickness and doping level through applied potential. For advanced applications requiring narrow molecular weight distributions, living polymerization techniques, such as controlled radical methods, facilitate the synthesis of well-defined chains with targeted molecular weights, enhancing reproducibility in device fabrication. Key properties of conducting polymers include tunable electrical conductivity, mechanical flexibility, and improved processability. Doping enables conductivity to span from 10^{-5} S/cm in undoped states to over 10^3 S/cm in heavily doped forms, such as iodized reaching metallic levels. Mechanically, these polymers exhibit flexibility with Young's moduli around 1 GPa for materials like regioregular (P3HT), far lower than inorganic semiconductors, supporting bendable electronics. Solubility is enhanced by incorporating alkyl side chains, as in P3HT, which promotes solution processing and self-assembly into ordered films without compromising conjugation. Doping in conducting polymers involves fractional charge transfer from the dopant to the polymer backbone, generating polarons or bipolarons that serve as charge carriers. The resulting conductivity follows the relation \sigma = ne\mu, where \sigma is conductivity, n is the dopant-derived carrier concentration, e is the elementary charge, and \mu is charge mobility; for , typical hole mobilities reach approximately 0.1 cm²/Vs under optimized doping. This mechanism allows fine-tuning of electronic properties, with dopant levels determining the transition from semiconducting to conducting regimes. Despite their advantages, conducting polymers face stability challenges, primarily degradation through oxidation in ambient conditions, which disrupts conjugation and reduces conductivity over time. In PEDOT:PSS, the polystyrenesulfonate (PSS) counterion stabilizes the positively charged PEDOT chains by preventing aggregation and improving water dispersibility, though residual acidity from PSS can accelerate device degradation via corrosion or defect formation. In organic electronic devices, conducting polymers primarily serve as hole-transport layers, facilitating efficient charge injection and extraction while leveraging their processability for large-area fabrication, with detailed implementations covered in subsequent sections.

Small-Molecule Organics

Small-molecule organics refer to discrete, low-molecular-weight compounds, typically with molecular weights below 1000 Da, that serve as active materials in organic electronics due to their well-defined structures and ability to form ordered films via vacuum deposition techniques. These materials enable precise control over layer thickness and interfaces in devices such as organic light-emitting diodes () and organic field-effect transistors (), facilitating high-performance optoelectronics. Unlike polymeric counterparts, small molecules can achieve superior purity and crystallinity, which are critical for efficient charge transport and light emission. Prominent examples include anthracene, a prototype polycyclic aromatic hydrocarbon studied for its photoconductivity since the early 20th century; pentacene, a linear acene widely used as a benchmark p-type semiconductor in thin-film transistors; and tris(8-hydroxyquinolinato)aluminum (Alq3), an archetypal electron-transporting and emissive material in OLEDs. Discotic liquid crystals, such as metal-free phthalocyanines, form columnar stacks that promote one-dimensional charge pathways, making them suitable for self-assembled semiconducting wires in field-effect devices. A key advantage of small-molecule organics is their high purity, achieved through thermal sublimation, which minimizes defects and impurities that impede charge mobility. This process allows for the growth of highly ordered crystalline films, as exemplified by pentacene single crystals exhibiting hole mobilities up to 35 cm²/V·s at room temperature, enabling band-like transport in high-quality devices. Electronically, small-molecule organics benefit from strong π-π stacking interactions between planar conjugated cores, which facilitate intermolecular charge transfer and delocalization essential for efficient hopping or band conduction. Their excitons exhibit binding energies around 0.5 eV, typically higher than in conjugated polymers due to reduced dielectric screening and more localized wavefunctions, influencing dissociation efficiency in optoelectronic applications. Fullerene derivatives like buckminsterfullerene (C60) and [6,6]-phenyl-C61-butyric acid methyl ester (PCBM) function as robust n-type electron acceptors, with LUMO levels around -4.5 eV for C60 and approximately -4.2 eV for PCBM, enabling efficient electron injection and extraction in devices. These molecules' spherical symmetry and high electron affinity make them ideal for bulk heterojunctions in photovoltaics. Non-fullerene acceptors (NFAs), such as Y6 (a small-molecule A-D-A type with a dicyano-substituted core) and its derivatives, represent a major advance in small-molecule organics for organic photovoltaics, featuring tunable LUMO levels around -4.1 eV, broader absorption, and improved morphological stability. As of 2023, NFAs have enabled single-junction OPVs with power conversion efficiencies exceeding 18%, surpassing fullerene-based devices in performance and longevity. Synthesis of small-molecule organics often involves multi-step organic reactions to construct extended π-conjugated systems, such as the for oligothiophenes, which links thiophene units with aryl halides under palladium catalysis to yield soluble, processable semiconductors. In hybrid systems, small molecules are blended with polymers to balance electron and hole transport, leveraging the crystallinity of small molecules to enhance overall mobility while retaining the solution-processability of polymers, as demonstrated in improved charge separation in mixed films.

Processing and Fabrication Methods

Organic electronics rely on a variety of processing and fabrication methods to deposit and pattern thin films of organic materials, enabling the creation of functional devices on both rigid and flexible substrates. These techniques must account for the sensitivity of organic semiconductors to heat, solvents, and environmental contaminants, while prioritizing scalability for large-area applications such as displays and solar cells. Vacuum-based methods are particularly suited for small-molecule organics, which sublime under controlled conditions to form high-purity films. Thermal evaporation, a cornerstone vacuum technique, involves heating small-molecule materials in an ultra-high vacuum chamber (typically 10^{-7} to 10^{-10} Torr) to evaporate them onto a substrate, achieving deposition rates of 0.1-1 Å/s to minimize defects and ensure uniform coverage. This method is widely used for multilayer structures in devices like (OLEDs), where precise control over film thickness (often 10-100 nm) is critical to optimize charge injection and recombination. Ultra-high vacuum conditions prevent contamination from residual gases, which could degrade carrier mobility to below 1 cm²/V·s in films like . However, thermal evaporation requires line-of-sight deposition, limiting its use for complex topographies without additional tooling. In contrast, solution-based processing leverages the solubility of conducting polymers and some small molecules, enabling cost-effective, large-area fabrication compatible with roll-to-roll (R2R) production. Spin-coating disperses the material in orthogonal solvents (e.g., chlorobenzene for polymers like , avoiding dissolution of underlying layers) onto a spinning substrate, yielding smooth films of 50-200 nm thickness with high reproducibility for lab-scale prototyping. For scalable manufacturing, inkjet printing deposits picoliter droplets of ink formulations (e.g., dispersions with conductivities up to 1000 S/cm), achieving resolutions down to 20 μm while minimizing material waste compared to spin-coating, which discards over 90% of the solution. R2R techniques, such as slot-die or gravure printing, further enhance throughput for polymer-based films, with solvent orthogonality ensuring layer integrity in stacked architectures like organic photovoltaics (). Advanced deposition methods bridge the gap between and solution processing for improved uniformity and . Organic vapor phase deposition (OVPD) transports vaporized organics via a hot inert carrier gas (e.g., nitrogen at 200-400°C) to a cooler substrate, enabling particle-free films with doping control and growth rates up to 10 nm/s, outperforming thermal evaporation in uniformity over large areas (>1 m²). This technique has demonstrated OLED efficiencies comparable to vacuum thermal evaporation, with external quantum efficiencies exceeding 10% in phosphorescent devices. Blade-coating, a solution-based variant, shears inks across substrates at speeds of 1-10 m/min, ideal for OPV , producing active layers with power conversion efficiencies around 10% on flexible foils while reducing use by 50% relative to spin-coating. Patterning techniques are essential for defining active areas and interconnects without damaging sensitive organics, often adapting methods from inorganic electronics while avoiding aggressive chemicals. Shadow masking, a vacuum-compatible approach, uses metal stencils to selectively deposit evaporated materials, achieving alignments down to 50 μm for pixelated arrays, though it requires precise mechanical alignment to prevent shadows in multilayer stacks. Adaptations of employ lift-off processes with solvent-resistant photoresists (e.g., novolac-based) to pattern electrodes and dielectrics, circumventing harsh developers that could etch organics; resolutions as fine as 5 μm have been reported for transistors. , including with elastomeric stamps, enables non-contact patterning of self-assembled monolayers for surface modification, supporting feature sizes below 1 μm in flexible circuits without vacuum needs. Encapsulation protects organic layers from oxygen and moisture permeation, which can reduce device lifetimes from hours to years by preventing oxidative degradation. (ALD) of Al₂O₃ barriers (10-100 nm thick) at low temperatures (<100°C) forms dense, pinhole-free films with water vapor transmission rates below 10^{-6} g/m²/day, enabling OLED operational lifetimes exceeding 10,000 hours under accelerated testing. Multilayer stacks, such as Al₂O₃/ZrO₂ nanolaminates, further enhance barrier performance by mitigating stress-induced cracks on flexible substrates, maintaining >90% efficiency retention after 1000 bending cycles. These encapsulations are often applied via R2R ALD for commercial viability in OPVs and displays. As of , printable electronics have advanced with aerosol jet printing, which atomizes inks into a carrier gas stream for non-contact deposition on flexible substrates like , achieving resolutions below 50 μm (down to 10 μm) for high-density interconnects in wearable sensors and . This method supports multimaterial printing at speeds up to 1 m/s, with silver nanoparticle inks yielding conductivities >10^6 S/m post-sintering, enabling scalable fabrication of sensors with factors >100 on curved surfaces. Such innovations address challenges, paving the way for conformable in biomedical applications.

Devices and Applications

Organic Light-Emitting Diodes

Organic light-emitting diodes (s) are emissive devices that generate light through in organic thin films, enabling applications in displays and lighting due to their , wide viewing angles, and potential for flexibility. In an , electrical current drives the recombination of injected charge carriers to form excitons, which decay radiatively to produce photons. This process contrasts with inorganic LEDs by relying on molecular orbitals in organic materials for charge transport and emission, allowing for solution-processable fabrication and tunable emission colors. The technology originated from pioneering work demonstrating efficient double-layer structures, achieving over 1,000 cd/m² with external quantum efficiencies around 1%. A typical OLED structure consists of an (ITO) anode on a transparent , a hole-transport layer such as N,N'-bis(3-methylphenyl)-N,N'-diphenylbenzidine (TPD), an emissive layer like tris(8-hydroxyquinolinato)aluminum (Alq₃), an electron-transport layer, and a low-work-function cathode such as calcium/aluminum (Ca/Al). This multilayer configuration facilitates balanced charge injection: holes from the anode traverse the hole-transport layer to the emissive region, while electrons from the cathode reach it via the electron-transport layer, minimizing leakage and enhancing recombination . The organic layers are typically 50–200 nm thick, deposited via or solution methods, with the emissive layer serving dual roles in light emission and charge transport in simpler designs. Operation begins with charge injection at the electrodes, overcoming barriers via applied voltage (typically 2–10 V), followed by drift and of carriers to form excitons in the emissive layer. Excitons form as bound electron-hole pairs, with spin statistics yielding a singlet-to-triplet ratio of 1:3 due to the conservation of total spin during recombination. Only singlet excitons (25%) typically decay radiatively in fluorescent OLEDs, while triplets (75%) are lost to non-radiative processes unless harvested. The overall efficiency is governed by the equation: \eta = \gamma \times \chi \times \eta_{\text{IQE}} where \gamma is the charge balance factor (ratio of electrons to holes recombining, ideally ~1), \chi is the singlet fraction (0.25 for fluorescence), and \eta_{\text{IQE}} is the internal quantum efficiency accounting for radiative decay yield. This limits fluorescent OLEDs to theoretical maximum external quantum efficiencies (EQE) of ~5–7.5% after light outcoupling losses. Phosphorescent OLEDs overcome the triplet limitation by using heavy-metal complexes, such as fac-tris(2-phenylpyridinato)(III) (Ir(ppy)₃), which enable strong spin-orbit coupling for and radiative decay from both singlet and triplet states, approaching 100% internal . These complexes are doped into a host matrix to prevent concentration , with early devices demonstrating EQE exceeding 20% through efficient triplet harvesting via energy transfer from the host. This breakthrough, building on foundational demonstrations, has enabled commercial high-efficiency green and red emitters. Color tuning in OLEDs is achieved through host-guest systems, where emissive dopants (guests) are incorporated at low concentrations (1–10 %) into a wide-bandgap matrix to control emission wavelength and suppress self-quenching. The facilitates charge transport and formation, transferring energy to the guest for tuned emission, such as blue from derivatives or red from complexes, while dopant levels optimize Förster and energy transfer without aggregation. This approach allows precise spectral control for full-color displays. Operational stability is a key metric for OLEDs, with lifetime defined as T₅₀, the time to reach 50% initial . Commercial phosphorescent OLEDs achieve T₅₀ >100,000 hours at 100 cd/m², limited by degradation mechanisms like exciton-induced molecular dissociation and charge accumulation, but mitigated through material encapsulation and balanced injection. This durability supports applications in televisions and lighting panels. Variants include flexible OLEDs fabricated on plastic substrates like (), enabling bendable displays with radii down to 1 mm while maintaining efficiency and lifetime comparable to rigid counterparts. These devices require thin-film barriers against and oxygen , often using multilayer stacks, to preserve performance under mechanical stress.

Organic Photovoltaics

Organic photovoltaics (OPVs) represent a class of organic electronic devices designed for harvesting, converting sunlight into electrical power through photo-induced charge generation in . Unlike traditional inorganic solar cells, OPVs rely on the unique properties of organic materials, such as solution processability and mechanical flexibility, to achieve lightweight and potentially low-cost photovoltaic modules. The active layer in OPVs typically consists of a blend of electron-donor and electron-acceptor materials, enabling efficient light absorption and charge transport. The core architecture of modern OPVs is the bulk heterojunction (BHJ), where donor and acceptor materials are intermixed at the nanoscale to form a bicontinuous network of phase-separated domains. A representative example is the blend of poly(3-hexylthiophene) (P3HT) as the donor polymer and [6,6]-phenyl-C61-butyric acid methyl ester (PCBM) as the acceptor, spin-coated from solution to create interpenetrating domains on the order of 10-20 nm. This morphology is optimized to match the typical diffusion length in , which is approximately 10 nm, ensuring that photogenerated excitons can reach a donor-acceptor (D-A) interface before recombining. The BHJ concept was first demonstrated in , markedly improving charge collection compared to planar heterojunctions. The operational processes in OPVs begin with photoexcitation, where absorbed photons create bound electron-hole pairs (s) in the donor material due to the relatively high in organics (around 0.3-0.5 ). These must diffuse to a D-A within their short lifetime (picoseconds to nanoseconds) for into free charge carriers via ultrafast to the acceptor. The separated electrons and holes then migrate through their respective percolating pathways to the electrodes for collection, driven by the built-in potential and external load. This sequence contrasts with inorganic , where free carriers are generated directly. The theoretical efficiency limit for OPVs, adapted from the Shockley-Queisser model for low-mobility organics with bandgaps of 1.2-1.6 , is approximately 30%, accounting for radiative and non-radiative recombination losses. Device performance is characterized by current-voltage (J-V) curves measured under standard AM1.5 global illumination (100 mW/cm²), revealing key parameters: (V_oc), short-circuit (J_sc), and fill factor (). The power conversion efficiency is given by η = (V_oc × J_sc × ) / P_in, where P_in is the incident (100 mW/cm²) and is the fill factor, defined as = P_max / (V_oc × J_sc) with P_max the maximum power output. The quantifies the deviation from rectangular J-V behavior due to series resistance, shunt paths, and recombination. High values (>70%) indicate efficient charge extraction with minimal losses. OPV efficiencies have evolved dramatically since the mid-1990s, starting from around 1% for early BHJ devices and reaching a certified 19.2% as of September 2025 for single-junction cells, driven by advances in and processing. The introduction of non-fullerene acceptors (NFAs) like ITIC in the mid-2010s enabled complementary absorption and reduced energy loss, boosting efficiencies beyond 10% and surpassing fullerene-based systems like PCBM. Seminal NFA blends, such as PBDB-T:ITIC, achieved initial records of ~11%, paving the way for further optimizations with Y-series NFAs that pushed certified efficiencies to 19.2% as tracked by NREL. Stability remains a challenge, with UV-induced degradation primarily causing photo-oxidation of the active layer and electrodes, leading to reduced J_sc and over time under prolonged exposure. Encapsulation strategies, such as barrier films with UV filters (e.g., or laminates with low transmission rates <10^{-4} g/m²/day), have extended operational lifetimes to over 1000 hours at 85°C/85% RH, mitigating oxygen and moisture ingress while blocking harmful UV wavelengths. For practical deployment, OPVs excel in scalability via printing techniques like roll-to-roll slot-die coating, enabling large-area modules exceeding 1 m² with power conversion efficiencies around 10%. These modules demonstrate viable output for applications like building-integrated photovoltaics, with geometric fill factors >70% to minimize inactive areas.00099-0)

Organic Field-Effect Transistors

Organic field-effect transistors (OFETs) are thin-film devices that utilize to control flow in a channel via an applied at the , enabling switching and functions central to electronics. Unlike inorganic counterparts, OFETs leverage the unique properties of organic materials, such as processability and mechanical flexibility, to achieve low-cost fabrication on large-area substrates. The channel conduction is modulated by accumulating s at the semiconductor-dielectric interface, typically operating in accumulation mode where the device is off at zero voltage and turns on with increasing bias. The standard structure of an OFET consists of a gate electrode, a layer, an channel, and source/drain electrodes, with the bottom-gate/top-contact (BGTC) configuration being one of the most common architectures. In this setup, a heavily doped wafer serves as the bottom gate, coated with a dielectric such as SiO₂ or a like PMMA, followed by the organic semiconductor layer (e.g., pentacene for small molecules or P3HT for polymers), and topped with source and drain electrodes typically made of or silver. The channel length (L) and width (W) are defined by the spacing between source and drain, often ranging from micrometers to millimeters. This staggered facilitates good charge injection while minimizing parasitic capacitances. Operation of OFETs relies on field-induced charge accumulation, primarily in the saturation regime for most characterizations, where the drain current I_d follows the gradual channel approximation: I_d = \frac{W}{2L} C_i \mu (V_g - V_{th})^2 Here, V_g is the gate voltage, V_{th} is the threshold voltage marking the onset of strong inversion, C_i is the gate dielectric capacitance per unit area, and \mu is the field-effect mobility. Mobility is extracted from the saturation regime using: \mu = \frac{2L}{W C_i V_d} \left( \frac{d \sqrt{I_d}}{d V_g} \right)^2 where V_d is the drain voltage. Typical devices exhibit on/off current ratios exceeding $10^6, enabling clear switching between conducting and insulating states, though V_{th} can shift due to trap filling at the interface. Contact effects at the source/drain electrodes significantly influence performance, as Schottky barriers form due to mismatches between metal work functions and organic semiconductor energy levels, leading to high injection barriers and contact resistance that can distort current-voltage characteristics. These barriers, often 0.1-0.5 eV high, limit charge injection efficiency, particularly in short-channel devices, and can overestimate mobility if unaccounted for. Mitigation strategies include inserting interlayers such as self-assembled monolayers (SAMs) or metal oxides (e.g., MoO_x), which tune the work function and reduce barrier heights, lowering contact resistance from kΩ·cm to Ω·cm levels in materials like pentacene or C8-BTBT. Alternative architectures address specific limitations; top-gate configurations (e.g., top-gate/bottom-contact) place the gate and atop the , providing encapsulation against and smoother interfaces for higher , though they require precise deposition to avoid damaging the layer. Vertical OFETs (VOFETs) orient the perpendicular to the , with length defined by layer thickness (nanometers), enabling ultra-short channels and high current densities up to 1 kA/cm²—orders of magnitude beyond lateral devices—suitable for high-speed applications while maintaining low operating voltages. Performance metrics have advanced significantly, with polycrystalline thin-film OFETs achieving mobilities of 1-10 cm²/V·s using materials like P3HT, while high-performance and single-crystal devices have reached mobilities exceeding 90 cm² V⁻¹ s⁻¹ as of 2025, approaching levels. in transfer characteristics, arising from charge traps at grain boundaries or interfaces, broadens the subthreshold swing and reduces operational stability, often mitigated by trap-passivating interlayers or high-purity . These enhancements enable practical on/off ratios >10^6 and threshold voltages near 0 V in optimized devices. Applications of OFETs include logic circuits, where arrays of p-type and n-type transistors form inverters, ring oscillators, and gates on flexible substrates, demonstrating clock frequencies up to 100 kHz for low-power . In sensing, OFETs exploit their sensitivity to environmental perturbations, serving as chemical or biosensors by modulating conductance in response to analytes like gases or biomolecules, with detection limits down to ppb levels due to the large surface area.

Emerging Devices

Organic electrochemical transistors (OECTs) represent a promising class of emerging devices in organic electronics, distinguished by their ion-gated in aqueous environments, which enables seamless with biological systems. Unlike traditional field-effect transistors, OECTs modulate through volumetric doping by ions, allowing high values, defined as g_m = \frac{dI_d}{dV_g}, reaching up to 100 mS in devices based on poly(3,4-ethylenedioxythiophene): (PEDOT:PSS). This exceptional sensitivity stems from the mixed ionic-electronic conduction in the polymer , making OECTs ideal for bio-sensing applications such as real-time detection of biomarkers like glucose or neurotransmitters in physiological fluids. Recent advancements have focused on fiber-based and flexible OECTs, enhancing their suitability for wearable diagnostics with stable performance over thousands of cycles in humid conditions. Organic resistive random-access memory (RRAM) devices leverage formation in organic active layers to achieve non-volatile , offering a low-power alternative to inorganic memories. In these devices, switching occurs via the formation and rupture of conductive filaments—typically metallic or oxygen vacancy-based—within polymers like or small-molecule organics, transitioning between high- and low-resistance states under applied voltage. Endurance exceeding $10^6 cycles has been demonstrated in organic-inorganic structures, attributed to controlled that minimize , with retention times over 10 years at . These devices are particularly valued for their compatibility with flexible substrates, enabling integration into and for portable . Organic sensors have advanced significantly for environmental and health monitoring, with gas and chemical detectors exploiting swelling-induced resistance changes in responsive polymers. For instance, conductive polymer composites, such as those incorporating or PEDOT, swell upon analyte absorption, altering inter-chain distances and thus electrical resistance, enabling selective detection of volatile organic compounds or toxic gases at parts-per-billion levels. Wearable strain sensors based on these materials achieve gauge factors greater than 100, far surpassing metallic foils, through piezoresistive mechanisms where strain deforms percolating networks of carbon nanotubes or within an elastomeric matrix. This high sensitivity supports applications in human motion tracking, with devices maintaining linearity over 100% strain and durability for over 10,000 cycles. In bioelectronics, neural interfaces utilizing cytocompatible organic polymers like poly(3-hexylthiophene) (P3HT) facilitate recording of neural signals with minimal damage. P3HT's and mixed conduction properties allow conformal coating on electrodes, reducing impedance and enabling high-fidelity extracellular recordings from or peripheral nerves over extended periods. These interfaces have demonstrated stable signal-to-noise ratios in rodent models, supporting applications in prosthetics and brain-machine interfaces by promoting neural adhesion without inflammatory responses. Hybrid organic-inorganic perovskites have emerged as key enablers for tandem solar cells, combining the broad absorption of layers with the of inorganic components to surpass single-junction efficiencies. In these structures, wide-bandgap inorganic perovskites like CsPbI3 serve as top cells atop narrow-bandgap bottom cells, achieving certified power conversion efficiencies of 26.7% as of 2025 while enhancing operational through defect passivation and encapsulation. Stability improvements, such as retaining 90% efficiency after 1,000 hours under standard testing, arise from suppressed ion migration and moisture resistance in all-inorganic variants. As of 2025, self-powered wearables integrating organic photovoltaics (OPV) with sensors exemplify key trends in organic electronics, enabling autonomous operation without batteries. These systems harvest ambient light via OPV modules to power on-skin sensors for continuous health monitoring, such as sweat analysis or , with integrated efficiencies supporting multi-day use. Advances in ultraflexible OPV-sensors hybrids have reduced form factors to sub-micrometer thicknesses, promoting and for next-generation platforms.

Properties and Challenges

Advantages Over Inorganic Electronics

Organic electronics offer superior mechanical properties compared to inorganic counterparts like , which is inherently brittle and prone to fracture under bending or stretching. and devices can achieve bend radii below 1 mm without significant performance degradation, enabling applications in conformable electronics such as wearable sensors and foldable displays. Furthermore, they exhibit stretchability exceeding 20% , with some organic light-emitting diodes maintaining or even enhancing under such deformation, contrasting sharply with the rigidity of silicon-based systems that fail at strains below 1%. Processing advantages stem from the ability to fabricate organic electronics via low-temperature solution-based methods, typically below 200°C, which allow deposition on flexible substrates using cost-effective techniques like inkjet or roll-to-roll . This contrasts with inorganic electronics requiring high-temperature vacuum processes, resulting in manufacturing costs for organic cells around $50–100 per m², comparable to at ~$30–50 per m² as of 2025, with potential for further reduction through scalable . These methods facilitate large-area production on rolls, producing lightweight devices with active layer weights under 1 g/m², ideal for scalable applications like flexible displays and panels. Molecular engineering in organic electronics enables precise tunability of electronic properties, such as bandgaps spanning 1.55–3.1 , corresponding to emission or absorption wavelengths from 400 to 800 , through structural modifications of conjugated polymers or small molecules. This versatility allows customization for specific applications, like blue-to-red emitters in displays, far surpassing the fixed bandgaps in inorganic semiconductors. Additionally, their carbon-based composition promotes by reducing reliance on rare-earth elements and enabling recyclability of semiconductors, supporting a with lower environmental impact than metal-intensive inorganic alternatives. The of organic electronics arises from their low and soft, tissue-like mechanical properties, making them suitable for implantable devices such as neural interfaces or bio-sensors, where metallic implants can cause inflammatory responses due to release or mismatch. In contrast to metals like or , which may exhibit over time, organic materials minimize adverse biological reactions, facilitating long-term integration in medical applications.

Limitations and Research Directions

One major limitation in organic electronics is the poor long-term stability of materials and devices, particularly under photo and thermal stress. Organic semiconductors are susceptible to from UV exposure, which generates leading to chain scission and trap , and thermal degradation that induces in active layers. For instance, in organic photovoltaics (OPVs), outdoor lifetimes often fall below 5 years due to cumulative UV doses and temperature fluctuations accelerating morphological instability, though as of 2025, lab projections indicate operational lifetimes exceeding 10 years for advanced devices. To address this, are exploring antioxidants like non-fused non-fullerene acceptors that preserve 84% efficiency after 1900 hours of illumination, and cross-linking strategies in polymerized acceptors that double operational stability to 100 hours by reducing mixing . Efficiency in organic electronics is constrained by low charge carrier mobilities, typically ranging from 10^{-3} to 10 cm²/V·s, which current densities and overall compared to inorganic counterparts exceeding 1000 cm²/V·s. This gap arises from disordered molecular packing in , hindering efficient . Efforts to improve this include fabricating ordered through techniques, such as transfer processing, which enables high-mobility semiconductors with mobilities up to 10 cm²/V·s in field-effect transistors by promoting uniaxial . Scalability remains a challenge for large-area production, where printing methods suffer from non-uniformity in film thickness and composition, leading to device-to-device variability. Solution-based fabrication like inkjet or gravure printing requires precise control over ink rheology to achieve consistent deposition over substrates larger than 1 . Recent advances incorporate algorithms to optimize ink formulations, predicting and for improved uniformity in roll-to-roll processes, as demonstrated in gravure-printed electrodes with quantified defect reduction. Environmental concerns in organic electronics stem from the toxicity of common processing solvents, such as chlorinated hydrocarbons, which pose health risks and contribute to ecological harm during manufacturing and disposal. Halogenated solvents like have high toxicity profiles, complicating sustainable production. A shift toward water-based is underway, enabling high-efficiency OPVs with power conversion efficiencies over 14% while minimizing environmental impact, as seen in non-toxic formulations that maintain device performance without compromising solubility. Looking to 2025 frontiers, hybrid systems integrating quantum dots with materials have achieved efficiencies up to 21% in tandem architectures by enhancing light absorption and charge separation. Additionally, synapses are advancing , mimicking biological neural plasticity with low-power memristive devices for efficient brain-inspired processing in wearable . Economic barriers hinder widespread adoption, with organic electronics, particularly OLED displays, holding approximately 20-30% in premium display segments like smartphones and TVs as of 2025, though overall penetration in the broader display market remains limited due to LCD dominance. Projections indicate growth to over 40% in key segments by 2030, driven by cost reductions in OLED production and expanding applications in flexible screens.

References

  1. [1]
    Molecular and Organic Electronics | Nature Research Intelligence
    Detailed reviews have highlighted high‐mobility emissive organic semiconductors that effectively combine robust charge transport with tunable optical properties ...
  2. [2]
    Highly Ordered Small Molecule Organic Semiconductor Thin-Films ...
    Organic semiconductors have opened up many new electronic applications, enabled by properties like flexibility, low-cost manufacturing, and biocompatibility ...<|control11|><|separator|>
  3. [3]
    Organic Optoelectronic Materials: Mechanisms and Applications
    The goal of this review is to provide a balanced assessment of the current understanding of the physical mechanisms that determine the optoelectronic properties ...
  4. [4]
    Introduction: Organic Electronics and Optoelectronics
    Organic electronics has been the focus of a growing body of investigation in the fields of physics and chemistry for more than 50 years.
  5. [5]
    Introduction: Organic Bioelectronics | Chemical Reviews
    Feb 23, 2022 · The key advantage of organic semiconductors in bioelectronics is that they can be designed at the molecular level to transport both ionic and ...
  6. [6]
    High-performing organic electronics using terpene green solvents ...
    Dec 15, 2022 · Organic electronics (OEs) using carbon-based conjugated polymers and small molecules produce fewer carbon emissions in comparison with their ...
  7. [7]
    Charge Transport in Organic Semiconductors | Chemical Reviews
    π-Conjugated polymers are generally exploited as p-type materials. However, polymers with high electron affinity, such as the poly( ...
  8. [8]
    Review on Charge Carrier Transport in Inorganic and Organic ...
    In organic semiconductors, charge transport is typically described using a hopping mechanism, where charge carriers move between localised states. This is due ...
  9. [9]
    Intrinsic Charge Transport on the Surface of Organic Semiconductors
    Aug 20, 2004 · In this Letter, we report on the observation of the intrinsic transport of field-induced charges on the surface of single crystals of rubrene ( ...
  10. [10]
    Rubrene-Based Single-Crystal Organic Semiconductors: Synthesis ...
    This work lays a foundation to improve our understanding of charge-carrier transport phenomena in organic single-crystal semiconductors.
  11. [11]
    A review for polaron dependent charge transport in organic ...
    The purpose of this review is to provide a theoretical description of the polaron dependence of charge transport and to highlight the most essential results ...
  12. [12]
    Charge Transport in Organic Semiconductors: The Perspective from ...
    Feb 23, 2022 · Charge propagation occurs via thermally activated small polaron hopping events. For OSs with larger mobilities (naphtalene, anthracene, perylene ...
  13. [13]
    Elimination of charge-carrier trapping by molecular design - Nature
    Jun 29, 2023 · Organic semiconductors often show relatively poor charge transport properties compared with their inorganic counterparts. There are two ...
  14. [14]
    Review Characterisation of charge carrier transport in thin organic ...
    The paper reviews recent advances in characterisation of charge carrier transport in organic semiconductor layers by time-of-flight photocurrent measurements.
  15. [15]
    Colloquium: Electronic transport in single-crystal organic transistors
    Sep 29, 2006 · The benchmark for the study of charge transport in organic semiconductors was established by the time-of-flight (TOF) experiments with ultrapure ...
  16. [16]
    [PDF] Conductive polymers - Nobel Prize
    In 1977, however, Shirakawa, MacDiarmid and Heeger discovered that oxidation with chlorine, bromine or iodine vapour made polyacetylene films 109 times more ...
  17. [17]
    Synthesis of electrically conducting organic polymers: halogen ...
    Shirakawa, E. J. Louis, A. G. MacDiarmid, C. K. Chiang and A. J. Heeger, J. Chem. Soc., Chem. Commun., 1977, 578 DOI: 10.1039/C39770000578. To request ...
  18. [18]
    Electrical Conductivity in Doped Polyacetylene | Phys. Rev. Lett.
    Oct 24, 1977 · Doped polyacetylene forms a new class of conducting polymers in which the electrical conductivity can be systematically and continuously varied over a range of ...
  19. [19]
    [PDF] Alan J. Heeger - Nobel Lecture
    In 1976, Alan MacDiarmid, Hideki Shirakawa and I, together with a talented group of graduate students and post-doctoral researchers discovered con- ducting ...<|separator|>
  20. [20]
    The Nobel Prize in Chemistry 2000 - NobelPrize.org
    Heeger, Alan G. MacDiarmid and Hideki Shirakawa "for the discovery and development of conductive polymers". To cite this section. MLA style: The Nobel Prize ...
  21. [21]
    Studies on Molecular Conductors: From Organic Semiconductors to ...
    1.2. Electrical and Magnetic Properties of Charge-Transfer Complexes and Ion Radical Salts. In 1954, Akamatu, Inokuchi, and Matsunaga reported that the freshly ...
  22. [22]
    Novel superconducting phenomena in quasi-one-dimensional ...
    First, the superconductivity of Q1D Bechgaard salts shares a common border ... Jérome, H. Schulz. Adv. Phys., 31 (1982), p. 299. Crossref View in Scopus.Missing: paper | Show results with:paper
  23. [23]
    Modern History of Organic Conductors: An Overview - MDPI
    Akamatu, H.; Inokuchi, H.; Matsunaga, Y. Electrical conductivity of the perylene-bromine complex. Nature 1954, 173, 168–169. [Google Scholar] [CrossRef] ...
  24. [24]
    Charge transfer complexes: a review survey - ScienceDirect
    This review examines the synthesis, characteristics, and documented examples of CTCs.
  25. [25]
    OLED history: Invention to application tour
    Jan 11, 2025 · Sony unveils a 56" 4K2K Oxide-TFT OLED TV prototype; January ... 2003. December - Univision reaches monthly capacity of 6,000 OLED ...
  26. [26]
    Age of Flexible Electronics: Emerging Trends in Soft Multifunctional ...
    Jan 23, 2024 · With the commercialization of first-generation flexible displays and mobiles in the late 2010s, humanity has stepped into the age of flexible ...Selection Criteria for Soft... · Emerging Materials and... · Emerging Trends in Soft...
  27. [27]
    Single-Junction Organic Solar Cell with over 15% Efficiency Using ...
    Apr 17, 2019 · Here, we have designed and synthesized a new class of non-fullerene acceptor, Y6, that employs an electron-deficient-core-based central fused ...
  28. [28]
    Conducting polymers: a comprehensive review on recent advances ...
    Conducting polymers are extensively studied due to their outstanding properties, including tunable electrical property, optical and high mechanical properties.Missing: extrinsic | Show results with:extrinsic
  29. [29]
    Synthesis and Morphology of Conducting Polymers - ACS Publications
    May 15, 2023 · Various strategies have been reported for the synthesis of conducting polymers including chemical oxidation, electrochemical polymerization, vapor phase ...Missing: seminal | Show results with:seminal
  30. [30]
    Controlled/living polymerization towards functional poly( p ...
    Jan 12, 2016 · Two polymerization approaches are compared, living polymerization towards well-defined PPVs via ring-opening metathesis polymerization (ROMP) and chain- ...
  31. [31]
    Enhanced Electrical Conductivity of Molecularly p-Doped Poly(3 ...
    Oct 11, 2017 · As a result, we achieve a value of up to 12.7 S cm–1 for P3HT:F4TCNQ. We determine the F4TCNQ anion concentration and find that the number of ( ...
  32. [32]
    The damaging effects of the acidity in PEDOT:PSS on ...
    Mar 24, 2020 · PEDOT:PSS is well-known to cause degradation in devices and limit their stability due to the acidity of the PSS chain.Missing: counterion | Show results with:counterion
  33. [33]
    [PDF] Introduction to the Physics of Organic Semiconductors
    The first studies of the dark and photoconductivity of anthracene crystals (a prototype organic semiconductor, see e.g. [1]) date back to the early 20th centu-.
  34. [34]
    An Improved Synthesis of Pentacene: Rapid Access to a Benchmark ...
    Apr 20, 2012 · Pentacene is a benchmark organic semiconductor in thin-film organic electronic devices due to its π-conjugated electronic structure, its ...
  35. [35]
    Charge transport properties of a metal-free phthalocyanine discotic ...
    Discotic liquid crystals can self-align to form one-dimensional semiconducting wires, many tens of microns long. In this letter, we describe the preparation ...
  36. [36]
    Control of π–π Stacking via Crystal Engineering in Organic ...
    In the solid state, π–π stacking of small molecules is significant in the charge transport process. Since π–π stacking of conjugated small molecules is ...
  37. [37]
    Dependence of Exciton Binding Energy on Bandgap of Organic ...
    The lowest optical band gaps are found for linkers that either raise the HOMO or lower the LUMO most, while the highest optical band gap is found for Ph linkers ...
  38. [38]
    The significance of fullerene electron acceptors in organic solar cell ...
    The discrete acceptor energy levels represent increasingly electronegative fullerene derivatives, from ICTA (LUMO = 3.77 eV) to PCBM (LUMO = 4.07 eV).
  39. [39]
    Synthesis of thiophene/phenylene co‐oligomers. I. Phenyl‐capped ...
    May 6, 2009 · An alternative synthetic route to the phenyl‐capped oligothiophenes that utilizes the Suzuki coupling reaction is presented as well. Citing ...<|separator|>
  40. [40]
    Conjugated Polymer Controlled Morphology and Charge Transport ...
    Mar 9, 2020 · In this study, we report an effective approach to tune the crystallization, microstructure and charge transport of solution-processed organic semiconductors.
  41. [41]
    Vacuum Thermal Evaporation for OLEDs - The Advanced Portfolio
    Oct 16, 2025 · Vacuum thermal evaporation (VTE) is the dominant method for manufacturing high-performance small-molecule organic light-emitting diodes (OLEDs).
  42. [42]
    Thermal Evaporation - an overview | ScienceDirect Topics
    Thermal evaporation is a well-known method for coating a thin layer in which the source material evaporates in a vacuum due to high temperature heating.
  43. [43]
    Organic vapor phase deposition for the growth of large area organic ...
    Dec 11, 2009 · Our results suggest that OVPD is scalable to very large substrates, making it a key growth technology for low-cost organic electronics. The ...
  44. [44]
    Thin-Film Growth and Patterning Techniques for Small Molecular ...
    Jun 7, 2013 · This article reviews several deposition and patterning methods for small molecular thin films and devices, including organic molecular beam deposition.<|separator|>
  45. [45]
    None
    Summary of each segment:
  46. [46]
    Inkjet printing of organic electronics – comparison of deposition ...
    The nascent inkjet printing technique is compared to commonly used solution deposition methods, like spin-coating and doctor blading.
  47. [47]
    Printing and Coating Techniques for Scalable Organic Photovoltaic ...
    Overall, spin coating is one of the most common coating methods used in OPV research due to its reproducibility and ability to achieve high device performances.
  48. [48]
    [PDF] Organic Vapor Phase Deposition for Optoelectronic Devices
    OVPD transports organic molecules to a cold substrate using a hot inert carrier gas, improving control over doping and enabling rapid, particle-free deposition.
  49. [49]
    Organic Vapor Phase Deposition - Baldo - 1998 - Wiley Online Library
    Organic vapor phase deposition (OVPD) opens new opportunities in the control of composition and structure of organic thin films.
  50. [50]
    A novel method for fabrication of integrated shadow mask for ...
    The pillar strips serve as an effective shadow mask for patterning the organic layers and the metal cathodes in OLED display devices.Missing: techniques | Show results with:techniques
  51. [51]
    [PDF] lithographic patterning processes for organic electronics
    Shadow Masks. Parallel process, commercially established. Resolution is on the order of tens of microns, requires high-vacuum environment. Photolithography.
  52. [52]
    Integrated shadow mask method for patterning small molecule ...
    Mar 7, 2006 · We have developed a simple and efficient method for patterning small molecule semiconductors for applications in the field of organic
  53. [53]
    Rapid fabrication of Al2O3 encapsulations for organic electronic ...
    Oct 30, 2015 · This paper report on the rapid fabrication of Al2O3 encapsulations through a unique roll-to-roll atmospheric atomic layer deposition technology ...
  54. [54]
    Reliable thin film encapsulation for organic light emitting diodes ...
    Our study shows that for low-temperature ALD the Al2O3/ZrO2 NL provides a substantial advantage over neat Al2O3 films and allows for device lifetimes well in ...
  55. [55]
    Al 2 O 3 /TiO 2 Nanolaminate Thin Film Encapsulation for Organic ...
    Flexible Ultrathin Chip-Film Patch for Electronic Component Integration and Encapsulation using Atomic Layer-Deposited Al2O3–TiO2 Nanolaminates. ACS Applied ...
  56. [56]
    Aerosol Jet Printed MXene Microsupercapacitors for Flexible and ...
    Jun 23, 2025 · AJP has emerged as a versatile low-waste and high-precision printing technology, offering sub-micron resolution (down to 10 µm), a mask-free ...
  57. [57]
    Aerosol jet printing process study of micro flexible strain sensors
    Oct 7, 2025 · The test results show that the line quality of the silver wire is good and the resistance is low in the range of 40–45 sccm of sheath gas flow, ...
  58. [58]
    Aerosol Jet Printing for Neuroprosthetic Device Development - PMC
    Jun 28, 2025 · This non-contact approach enables high-resolution patterning on a wide range of substrates with resolutions as fine as 10 µm and makes AJP ...
  59. [59]
    Organic electroluminescent diodes | Applied Physics Letters
    Sep 21, 1987 · A novel electroluminescent device is constructed using organic materials as the emitting elements. The diode has a double‐layer structure of organic thin films.
  60. [60]
    Design principles for highly efficient organic light-emitting diodes
    Aug 6, 2014 · In the vast majority of practical devices, the ratio of singlet and triplet excitons is 1:3 due to the degeneracy of the two states and ...
  61. [61]
    Highly Phosphorescent Bis-Cyclometalated Iridium Complexes
    Both singlet and triplet excited states can be trapped at the phosphor. OLEDs prepared with these heavy metal complexes are the most efficient OLEDs ...
  62. [62]
    Ideal host and guest system in phosphorescent OLEDs - ScienceDirect
    The dopant concentration in such a host-guest system is usually as high as about 6–10% by weight (wt%) because injected charges should move through guest ...Missing: review paper
  63. [63]
    Degradation Mechanisms and Reactions in Organic Light-Emitting ...
    in 1990, based on polymers, (1, 2) these devices have achieved tremendous progress. OLEDs are known for their high potential in display, signage, and lighting ...
  64. [64]
    A Review of Flexible OLEDs Toward Highly Durable Unusual Displays
    Jan 19, 2017 · Flexible OLEDs are promising due to their organic constituents, ultrathin structure, and low-temperature fabrication, but face issues with ...
  65. [65]
    Organic Photovoltaic Solar Cells - NREL
    Apr 3, 2025 · OPV is a rapidly emerging PV technology with improving cell efficiency (currently 18.2% certified), encouraging performance lifetime (>10 years ...Missing: best | Show results with:best
  66. [66]
    Polymer Photovoltaic Cells: Enhanced Efficiencies via a Network of ...
    The carrier collection efficiency (ηc) and energy conversion efficiency (ηe) of polymer photovoltaic cells were improved by blending of the semiconducting ...
  67. [67]
    Enhanced exciton diffusion in an organic photovoltaic cell by energy ...
    Typically the L D ( ∼ 10 nm ) of these materials is much less than the optical absorption length ( L A ∼ 100 nm ) ⁠,3 forcing the use of thin layers which limit ...
  68. [68]
    Bulk Heterojunction Organic Solar Cells: Working Principles and ...
    In 1995, Yu et al.29 reported an OSC based on a conjugated polymer : fullerene derivative absorber mixture with a PCE of ∼3%. Although this efficiency was ...
  69. [69]
    Can organic solar cells surpass the Shockley-Queisser limit?
    Oct 3, 2022 · In a combined experimental/numerical study, we show that the slowness of charge carrier thermalization leads to a 0.1-0.2 eV higher open ...
  70. [70]
    Adjusting the energy of interfacial states in organic photovoltaics for ...
    Mar 19, 2021 · Introduction. According to the Shockley and Queisser description, an ideal solar cell exhibits only unavoidable radiative losses1. Operated as ...
  71. [71]
    Fill Factor - PVEducation
    The FF is defined as the ratio of the maximum power from the solar cell to the product of V oc and I sc so that:Missing: organic | Show results with:organic
  72. [72]
    Achieving a high fill factor for organic solar cells - RSC Publishing
    In this review, we focus on the third parameter, fill factor (FF), that is equally important in determining the power conversion efficiency.
  73. [73]
    [PDF] NREL Best Research-Cell Efficiencies Chart
    Jul 1, 2025 · Emerging PV. Dye-sensitized cells. Perovskite cells. Organic cells. Organic tandem cells. CZTSSe cells. Quantum dot cells.
  74. [74]
    Non-Fullerene Electron Acceptors for Use in Organic Solar Cells
    Oct 27, 2015 · Some new non-fullerene acceptors have been demonstrated to outperform fullerenes in comparative devices. This Account describes this progress.
  75. [75]
    Donor polymer design enables efficient non-fullerene organic solar ...
    Oct 26, 2016 · Here we report non-fullerene organic solar cells with efficiencies up to 10.9%, enabled by a novel donor polymer that exhibits strong temperature-dependent ...
  76. [76]
    Progress in Stability of Organic Solar Cells - PMC - PubMed Central
    The photochemical degradation in OSCs is caused by the photo‐oxidation of the active layer and the carriers transport layer in most of the cases. In the active ...
  77. [77]
    Materials and methods for encapsulation of OPV: A review
    At present, however, commercial use of organic photovoltaics is hindered by low conversion efficiency and poor overall stability of the devices. Encapsulation ...
  78. [78]
    Scalable fabrication of organic solar cells based on non-fullerene ...
    Organic solar cells have recently experienced a substantial leap in power conversion efficiency, in part driven by formulations with new non-fullerene acceptors ...
  79. [79]
  80. [80]
  81. [81]
  82. [82]
    A guide for the characterization of organic electrochemical ...
    Jan 13, 2023 · We intend this review to be a guide for the characterization methods of the OECT and the channel materials used.
  83. [83]
    Fibre-based organic electrochemical transistors: principle ... - Nature
    May 11, 2025 · Here we present the research progress of F-OECTs along three lines: working principles, evaluation methods, and applications.
  84. [84]
    Flexible and Stretchable Organic Electrochemical Transistors for ...
    Feb 20, 2023 · Organic electrochemical transistors (OECTs) are an emerging class of organic electronic building blocks exhibiting significant advantages ...
  85. [85]
    Resistive Switching Random-Access Memory (RRAM): Applications ...
    The nonvolatile switching behavior ensures good retention even at elevated temperatures, while program/erase cycling shows strong endurance, making RRAM an ...
  86. [86]
    Bioresorbable Resistive Switching Device Based on Organic ...
    Jan 23, 2024 · The endurance behavior of RRAM devices was usually altered based on device configurations and applied parameters such as an organic/inorganic ...
  87. [87]
    A review on device requirements of resistive random access ...
    Sep 6, 2023 · In this Research Update, we summarize five requirements for RRAM devices and discuss the enhancement methods for each aspect.
  88. [88]
    State-of-the-Art advances and challenges in wearable gas sensors ...
    Dec 15, 2024 · This review categorizes wearable gas sensors based on their application areas, such as health monitoring, environmental monitoring, public security and safety,
  89. [89]
    Recent development of conductive polymer composite‐based strain ...
    Jul 29, 2023 · This review introduces in detail the preparation, morphologies and strain sensing performance of different CPC sensors including one-dimensional fibers, two- ...3.1 1d Fiber · 3.3 3d Sponge/foam And... · 4 Functionalities
  90. [90]
    Wearable strain sensors: state-of-the-art and future applications
    Feb 8, 2023 · This brief review describes aspects of flexible wearable strain sensing systems along with their challenges and prospects.Missing: swelling | Show results with:swelling
  91. [91]
    In Vivo Organic Bioelectronics for Neuromodulation - ACS Publications
    Jan 20, 2022 · In this review, we introduce a variety of organic material platforms and signaling modalities specifically designed for this role as “translator”.Missing: cytocompatible | Show results with:cytocompatible
  92. [92]
    P3HT‐Based Electroactive Films for In Vitro Neuronal Cell Interfacing
    In this work, a polymer blend of P3HT and multi‐walled carbon nanotubes (MWCNT) is prepared and deposited on the surface as an electroactive thin film resulting ...<|separator|>
  93. [93]
    Latest Strategies Promoting Stable and Efficient Perovskite, Organic ...
    Jun 23, 2025 · This review presents a comprehensive overview of recent advancements in perovskite, organic, and tandem solar cells, highlighting key ...Photovoltaic Efficiency and... · Perovskite Absorbers · Perovskite-Organic Tandems
  94. [94]
    Improved stability and efficiency of perovskite/organic tandem solar ...
    This work demonstrates that all-inorganic perovskites are an appropriate candidate for the fabrication of efficient and stable TSCs.
  95. [95]
    Self-powered wearable sensing devices for digital health
    May 21, 2025 · This review focuses on self-powered wearable sensing devices that utilize human energy, introducing their power supply mechanism and efficiency ...
  96. [96]
    Review Organic semiconductors based wearable bioelectronics
    The fabrication of organic semiconductor devices often involves multiple steps, beginning with careful substrate preparation. This typically includes thorough ...
  97. [97]
    Non-volatile organic memory with sub-millimetre bending radius
    Apr 8, 2014 · Memories with further severe bending have rarely been demonstrated and bending radii lower than 1 mm remain exceptional, preventing rational and ...
  98. [98]
    Intrinsically stretchable organic light-emitting diodes - Science
    Feb 24, 2021 · 3C, the L·L0−1 value of is-OLEDs is maintained at 1, even when OLEDs are stretched up to 40% strain, and even increases by 30% at 20% strain.
  99. [99]
    [PDF] SOLUTION PROCESSED ORGANIC LIGHT SOURCE FOR ...
    Organic and printed functional materials are typically processed at low temperatures (< 200 °C), which provides the possibility of using cost- efficient ...
  100. [100]
    Estimating the manufacturing cost of purely organic solar cells
    We estimate that the manufacturing cost for purely organic solar cells will range between $50 and $140/m2. Under the assumption of 5% efficiency, this leads to ...
  101. [101]
    [PDF] Organic electronics - Environmental Health Perspectives
    McGehee says most crystalline silicon photovoltaic cells have an energy conversion efficiency of 12% and a cost per square meter of $300.00 (equivalent to $3.50 ...
  102. [102]
    [PDF] Sustainability Considerations for Organic Electronic Products
    ... organic electronic applications. Indeed, less than 1 g of active material is required for the fabrication of a 1 m2 solar. Page 2. panel.1 However, this does ...
  103. [103]
  104. [104]
    Recycling Organic Semiconductors: Toward Sustainable Emerging ...
    Oct 3, 2025 · Considering the economic burden of synthesizing OSCs, recycling emerges as a cost-effective and sustainable alternative. Unlike biodegradation, ...
  105. [105]
    Design Parameters and Human Biocompatibility Assessment ...
    Apr 2, 2023 · Organic semiconductors have emerged as promising neural interfacing materials due to their innate biocompatibility, soft mechanical ...
  106. [106]
    Implantable bioelectronics toward long-term stability and sustainability
    Apr 7, 2021 · None or low toxicity is usually the foremost requirement for using a material for implanted devices, which is mainly determined by the ...
  107. [107]
    Toxicity of Metals Released from Implanted Medical Devices
    Implanted metallic medical devices release many of the same metals (e.g. aluminum, nickel, cobalt, chromium) that individuals are exposed to from ...
  108. [108]
    Stability of Organic Photovoltaics: From Root Causes to Advanced ...
    May 10, 2025 · This study is centered on the underlying mechanisms of degradation, providing a systematic and in-depth analysis of their instability factors.Early- and Mid- to Long-Term... · Analytic Tools for Degradation... · Conclusion
  109. [109]
    Recent Advances in Thermo- and Photostabilities of Organic Solar ...
    Feb 27, 2025 · In this review, we first provide a brief summary of the potential thermal degradation and photodegradation routes in the OSCs. Subsequently, we ...
  110. [110]
    Comprehensive Approach to Intrinsic Charge Carrier Mobility in ...
    We review a wide variety of organic electronics in terms of their charge carrier mobilities, and we describe recent studies of macromolecules, molecular ...
  111. [111]
  112. [112]
    Quantifying Printing Quality for Printed Electrodes via Deep ...
    Jun 22, 2025 · This study proposes a method to quantify the printing quality of roll-to-roll gravure-printed electronics by combining deep learning and ...
  113. [113]
    Convergence of biocompatible printed electronics and sensing in ...
    May 23, 2025 · The developed platform combines printing-inspired automation with machine learning models to improve device up scalability and performance.
  114. [114]
    Overcoming efficiency and stability limits in water-processing ...
    Dec 17, 2018 · These toxic solvents unnecessarily increase the environmental impact of a future OPV technology. Thus, today's fabrication of state-of-the ...
  115. [115]
    Advances in Organic In‐Sensor Neuromorphic Computing
    Jul 24, 2025 · This review discusses organic in-sensor neuromorphic computing for wearable and bioelectronic systems, with a focus on memory-based and ...
  116. [116]
    Organic Electronics Market Size, Share & Growth - MarketsandMarkets
    The organic electronics market is projected to grow from USD 59.9 billion in 2023 to USD 142.1 billion by 2028, driven by OLED displays.
  117. [117]