Fact-checked by Grok 2 weeks ago

Charge-transfer complex

A charge-transfer complex, also known as an electron donor-acceptor (EDA) complex, is a formed between an electron-rich donor or and an electron-poor acceptor, characterized by partial from the donor to the acceptor upon , often manifesting as intense bands in the visible or region. This partial charge transfer distinguishes these complexes from fully ionic compounds, as the typically involves weak non-covalent interactions such as π-π stacking, hydrogen bonding, or van der Waals forces, while the features a more pronounced charge-separated configuration. The concept was theoretically formalized by Robert S. Mulliken in 1952, who described these interactions using a two-state model involving no-bond (donor-acceptor ) and dative-bond (charge-transferred ) configurations to explain their spectroscopic properties. Charge-transfer complexes exhibit diverse structural motifs, ranging from loose molecular adducts in solution to crystalline solids with ordered donor-acceptor stacks, as seen in classic examples like the iodine- complex (C₆H₆·I₂), where acts as the donor and iodine as the acceptor, producing a broad CT band around 300 nm. More robust systems, such as the tetrathiafulvalene-tetracyanoquinodimethane (TTF-TCNQ) complex, form segregated or mixed-stack lattices that display metallic due to transfer and band overlap in the solid state. These complexes are typically identified through their characteristic electronic spectra, where the energy of the CT transition correlates with the potential of the donor and the of the acceptor, often following Mulliken's equation: h\nu_{CT} = I_D - E_A - C + \frac{2\beta_0^2}{I_D - E_A - C}, with I_D as donor potential, E_A as acceptor , C as Coulombic term, and \beta_0 as resonance integral. The significance of charge-transfer complexes extends across multiple fields in chemistry and , serving as models for understanding processes in biological systems, such as , where donor-acceptor pairs mimic natural pigment-protein interactions. In synthetic applications, they enable the design of with tunable bandgaps, as demonstrated in TTF-TCNQ's role as a prototype for one-dimensional conductors with conductivities on the order of 10³ S/cm at . Recent advances highlight their utility in , including organic photovoltaics where CT states facilitate dissociation, achieving power conversion efficiencies exceeding 20% as of 2025 in non-fullerene acceptors, and in , where EDA complexes promote metal-free aerobic oxidations via light-induced . Additionally, their sensitivity to environmental perturbations makes them ideal for gas sensing, with complexes like perylene-TCNE detecting volatile organic compounds through shifts in CT absorption.

Fundamentals

Definition

A charge-transfer complex is a formed by the association of two or more molecules or ions, typically an and an , stabilized primarily by electrostatic attraction arising from partial charge separation between them. In this interaction, the donor molecule acquires a partial positive charge (δ⁺), while the acceptor gains a partial negative charge (δ⁻), resulting in a non-covalent aggregate that exhibits properties distinct from those of the individual components. These complexes are characterized by a degree of charge transfer that ranges from weak (partial electron delocalization) to strong (approaching ionic or salt-like structures), depending on the relative strengths of the donor and acceptor and environmental factors such as solvent . The formation of a charge-transfer complex can be represented for weak cases by the equilibrium D + A ⇌ [Dδ+⋯Aδ⁻] in the ground state, where D denotes the and A the ; upon of , an promotes the system to an approximated as [D+⋯A], involving more complete charge separation. are with relatively low potentials, facilitating release, such as aromatic hydrocarbons like or . In contrast, possess high affinities, enabling acceptance, exemplified by nitroaromatic compounds like trinitrobenzene or quinones. Unlike covalent bonds, which involve significant electron sharing and strong orbital overlap leading to full bond formation, charge-transfer complexes rely on weaker, non-covalent interactions without substantial covalent character, though partial charge transfer contributes to their stability and unique spectroscopic signatures. This distinction underscores their supramolecular nature, where the integrity of the complex is maintained by intermolecular forces rather than intramolecular bonding.

Historical Development

The earliest observation of a charge-transfer complex dates to 1844, when discovered quinhydrone, a 1:1 complex of p-benzoquinone and , during studies on the relationship between and its . This dark green crystalline material intrigued chemists due to its intense color, which contrasted with the pale hues of its components, though the underlying electronic mechanism remained unexplained for over a century. The modern understanding of charge-transfer complexes emerged in the mid-20th century, with Robert S. Mulliken's seminal work formalizing donor-acceptor interactions through . In his 1952 publication, Mulliken described these complexes as weak associations where an and acceptor form a stabilized by charge-transfer excitations, distinguishing them from simple van der Waals adducts and introducing charge-transfer bands in electronic spectra as a key diagnostic feature. This framework spurred experimental investigations throughout the , establishing charge-transfer as a distinct class of intermolecular bonding. Key synthetic advancements in the late and enabled the creation of more stable and versatile complexes. Tetracyanoethylene (TCNE), a potent , was first synthesized in 1957 by and colleagues via dehydration of malononitrile derivatives, allowing the formation of deeply colored complexes with aromatic donors. This was followed by the preparation of 7,7,8,8-tetracyanoquinodimethane (TCNQ) in 1962 by Hertler et al., which exhibited even stronger accepting properties and facilitated complexes with enhanced stability and conductivity. These acceptors shifted focus from neutral or partial charge-transfer species to those approaching full ionic character, though complete ionization was not routinely achieved until late 20th-century refinements in and . Research accelerated in the 1970s with the discovery of tetrathiafulvalene-tetracyanoquinodimethane (TTF-TCNQ) in 1973 by Ferraris, Cowan, and coworkers, recognized as the first organic metal due to its room-temperature conductivity exceeding 10^3 S/cm along segregated stacks of donor and acceptor molecules. This milestone, independently confirmed by Heeger et al., ignited interest in conducting charge-transfer solids and inspired donor modifications like tetramethyltetrathiafulvalene (TMTTF). By 1980, Jérome, Bechgaard, and collaborators reported superconductivity in (TMTSF)_2PF_6 under 12 kbar pressure at 1.2 K, the first such phenomenon in an organic material, extending charge-transfer concepts to low-temperature electronic phases. In recent decades up to 2025, charge-transfer complexes have integrated with , enhancing optoelectronic applications such as photoresponsive sensors and transistors through electrocrystallization on nanostructured substrates. These developments leverage tunable charge separation for improved efficiency in devices like organic photovoltaics and light-emitting diodes.

Theoretical Framework

Mulliken's Theory

Mulliken proposed a quantum mechanical framework in to describe the formation and properties of charge-transfer complexes, emphasizing interactions between electron donors (D) and acceptors (A). In this model, weak charge-transfer complexes are represented as resonance hybrids of a no-bond state (D···A), where the donor and acceptor are loosely associated without significant charge separation, and an ionic state (D⁺ A⁻), involving partial from donor to acceptor. The of the complex is predominantly the no-bond configuration, with only a small contribution from the ionic form, whereas the is largely ionic. This stabilizes the complex, with the extent of mixing determined by the energy difference between the two states and their interaction matrix element. The energy of the charge-transfer (CT) transition, corresponding to promotion to the excited state, is approximated by the equation E_{CT} = IP_D - EA_A - C + \Delta where IP_D is the ionization potential of the donor, EA_A is the electron affinity of the acceptor, C accounts for Coulombic stabilization between the resulting ions, and \Delta encompasses additional interactions such as polarization and solvation effects. This transition involves excitation of an electron from the donor's highest occupied molecular orbital (HOMO) to the acceptor's lowest unoccupied molecular orbital (LUMO), producing characteristic intense absorption bands in the visible or near-ultraviolet spectrum. The stability of these complexes is quantified by the association constant K = \frac{[DA]}{[D][A]}, which increases with stronger donors (lower IP_D) and acceptors (higher EA_A). For instance, the benzene-tetracyanoethylene (TCNE) complex exhibits K \approx 0.128 L/mol in at , reflecting its weak binding nature. Mulliken's primarily applies to such weak complexes, where the ionic contribution remains minor; stronger complexes, with larger K values, show greater deviation toward predominantly and require extensions beyond the simple two-state model.

Electronic Structure and Bonding

In charge-transfer complexes, the electronic structure is primarily described by the frontier molecular orbital approach, wherein the highest occupied (HOMO) of the overlaps with the lowest unoccupied (LUMO) of the . This overlap results in the formation of new bonding and antibonding s, with partial mixing that facilitates a degree of charge transfer from the donor to the acceptor. The extent of this charge transfer is quantified by the parameter δ in the structural notation [D^{δ+} \cdots A^{δ-}], where δ denotes the fraction of an transferred per donor-acceptor pair. Weak charge-transfer complexes exhibit δ < 0.5, characterized by minimal electron delocalization and neutral-like behavior, whereas strong complexes display δ > 0.5, indicating substantial charge separation and approaching ionic character around a critical value near 0.5. In the solid state, π-π stacking arrangements play a key role in the electronic structure, manifesting as either segregated stacks—where donor and acceptor molecules form distinct columnar arrays—or mixed stacks featuring alternation of donor (D) and acceptor (A) units along the stack axis. Mixed stacks often promote dimerization, as the alternating D-A pattern enhances local interactions and leads to paired molecular units with altered orbital overlaps. Density functional theory (DFT) calculations provide insights into the binding energies and structural perturbations upon complex formation, typically revealing values of 10–20 kcal/mol for representative systems, driven by combined electrostatic, dispersion, and charge-transfer contributions. These computations also highlight geometric changes, such as reduced intermolecular distances and planar distortions in the donor and acceptor moieties, which optimize the HOMO-LUMO overlap. Distinct from coordination compounds, which involve dative covalent bonds from ligands to a central atom, charge-transfer complexes rely on non-covalent interactions—primarily electrostatic and dispersion forces—with charge transfer providing an additional stabilizing component but without forming sigma dative linkages. This molecular orbital framework extends Mulliken's valence bond resonance model by incorporating explicit orbital mixing and modern computational validation.

Classification and Examples

Donor-Acceptor Molecular Complexes

Donor-acceptor molecular complexes represent a fundamental class of charge-transfer complexes, formed through the interaction of electron-rich organic donors, such as aromatic hydrocarbons, with electron-poor acceptors like nitroaromatic compounds or cyano-substituted alkenes. These complexes typically exhibit partial charge transfer from the highest occupied molecular orbital (HOMO) of the donor to the lowest unoccupied molecular orbital (LUMO) of the acceptor, resulting in distinctive optical properties and structural arrangements. Classic examples include pairs of simple aromatics with strong acceptors such as tetracyanoethylene (TCNE) or nitrobenzenes. For instance, forms a weak with TCNE, while more electron-rich donors like or yield stronger interactions due to increased in their π-systems. Stability notably increases with donor electron richness, as seen in the progression from benzene-TCNE to hexamethylbenzene-TCNE, where the transferred charge and binding strength correlate linearly with the number of electron-donating methyl substituents. A representative case is the hexamethylbenzene-TCNE , which demonstrates enhanced association compared to unsubstituted analogs, reflecting the role of alkyl groups in elevating the donor's energy. A prominent example is quinhydrone, a 1:1 complex between p-benzoquinone (acceptor) and (donor), characterized by alternating π-stacked layers and skew pancake bonding that facilitates charge transfer. This complex appears deep purple in both solution and solid state, owing to a broad charge-transfer absorption band centered around 530 nm. Picrate complexes provide early crystalline illustrations of these interactions, formed between aromatic hydrocarbons (donors) and (acceptor), often as colorful salts with distinct spectral signatures indicative of charge-transfer bonding. These were historically employed in qualitative for identifying aromatic compounds, leveraging their characteristic colors and behaviors to distinguish hydrocarbons like or . In , these complexes are generally weak and governed by dynamic equilibria, with constants influenced by polarity—nonpolar media favor formation by minimizing of the polar CT state. In contrast, solid-state variants are more stable, frequently adopting crystalline structures with π-π stacking between donor and acceptor units, leading to ordered columnar arrangements that enhance intermolecular charge delocalization. Stability trends follow predictable patterns: electron-donating substituents (e.g., alkyl or alkoxy groups) on the donor lower its ionization potential, increasing the for complex formation, while electron-withdrawing groups (e.g., or cyano) on the acceptor increase its , similarly promoting charge transfer and binding affinity. These effects underscore the tunability of donor-acceptor es for applications in molecular recognition and materials design.

Halogen and Interhalogen Complexes

Charge-transfer complexes involving and interhalogens, such as iodine (I₂) and (), form through interactions with n-electron donors like amines and ethers, or π-electron donors such as . These complexes arise from the electron-accepting ability of the halogen molecule, where the donor provides electrons to the antibonding orbital of the halogen, leading to a separation. A prominent example is the starch-iodine complex, in which the helical structure of traps linear polyiodide species like I₃⁻ or chains of I₂ molecules within its cavity. This interaction produces a characteristic blue-purple color due to a broad charge-transfer absorption band in the visible region, typically around 600 nm, resulting from excitation of an from the donor to the acceptor. The complex is widely employed for qualitative detection of iodine in and as a security feature in counterfeit currency detection, where the color change indicates the presence of starch or iodine-based inks. The bonding in these complexes is characterized by σ-donation from the lone pair of the n-donor to the σ* antibonding orbital of the molecule, accompanied by partial charge transfer that polarizes the halogen bond. The I₂···NH₃ complex serves as a prototypical example, where the lone pair donates to the I-I σ* orbital, resulting in a linear N-I···I arrangement with a of approximately 4-5 kcal/mol and a measurable charge transfer of about 0.1 electrons. This donation enhances the electrophilicity of the , particularly at the terminal atom. In these complexes, function as mild oxidants, with the partial positive charge on the halogen facilitating electrophilic additions to nucleophilic substrates. For instance, amine-iodine complexes enable selective iodination of in aqueous media under mild conditions, where the complex acts as a source of electrophilic I⁺ equivalent without requiring harsh oxidants or catalysts. This reactivity contrasts with free , as the donor moderates the oxidizing power, allowing controlled addition reactions. The stability of charge-transfer complexes is generally weaker in the gas phase, where interactions rely primarily on and electrostatic forces without support, often exhibiting energies below 5 kcal/mol. In polar , however, stability increases due to the ion-pair character from separation, which is stabilized by solvent screening and specific of the charged components. This solvent enhancement can raise association constants by orders of magnitude compared to non-polar media.

Conducting Charge-Transfer Solids

Conducting charge-transfer solids are characterized by segregated stack structures, where pure donor and acceptor molecules form separate one-dimensional chains, facilitating band-like conduction along the stacking direction. This arrangement contrasts with mixed-stack configurations and allows for partial charge transfer between donor and acceptor stacks, enabling metallic behavior in these materials. The prototype example is tetrathiafulvalene-tetracyanoquinodimethane (TTF-TCNQ), a 1:1 charge-transfer complex first synthesized in 1973. In TTF-TCNQ single crystals, segregated stacks of TTF donors and TCNQ acceptors result in highly anisotropic , with values reaching approximately 10^3 S/cm along the b-axis at , marking it as one of the earliest metals. At low temperatures, TTF-TCNQ undergoes Peierls distortions, leading to a metal-insulator ; the TCNQ chains exhibit a charge-density-wave around 54 K, while the TTF chains follow at approximately 38 K, reducing below these thresholds. In contrast, mixed-stack charge-transfer solids, where donors and acceptors alternate in a single column, typically display insulating properties due to charge localization and stronger electron-hole binding, limiting delocalization. Segregated stacks in conducting variants promote transfer, with a degree δ ≈ 0.5–0.6 electrons per donor-acceptor pair in TTF-TCNQ, fostering mixed-valence states that support coherent transport. Superconducting behavior emerges in related segregated-stack systems, such as bis(tetramethyltetraselenafulvalene) hexafluorophosphate, (TMTSF)_2PF_6, discovered in 1980 as the first organic superconductor. This material exhibits superconductivity at a critical temperature of 0.9 K under applied pressure of about 12 kbar, attributed to its quasi-one-dimensional segregated chains with partial charge transfer enabling Cooper pair formation. Doping through partial oxidation further enhances conductivity in these solids by introducing carriers that partially fill electronic bands, mimicking semiconductor behavior while suppressing insulating transitions. For instance, controlled oxidation levels in donor stacks can tune the carrier density, improving metallic conduction without fully localizing charges.

Properties

Optical and Spectroscopic Properties

Charge-transfer complexes exhibit characteristic optical bands, known as charge-transfer () bands, which arise from transitions involving the promotion of an from the highest occupied (HOMO) of the donor to the lowest unoccupied molecular orbital (LUMO) of the acceptor. These bands are typically intense and broad, appearing in the ultraviolet-visible (UV-Vis) region, with molar extinction coefficients (ε) often exceeding 10^4 L mol^{-1} cm^{-1}, reflecting the allowed nature of the transition. In solution, such can indicate contact charge-transfer interactions between donor and acceptor molecules that are not fully bound in stable complexes. The visible color of charge-transfer complexes originates from these CT bands; weak complexes, involving minimal charge separation, often appear pale yellow or red, while stronger ones with greater charge transfer display deeper colors, such as the purple hue of quinhydrone, a 1:1 complex of hydroquinone and p-benzoquinone. This coloration provides a straightforward visual indicator of complex formation and strength, as the energy of the CT transition falls within the visible spectrum for many systems. Ultraviolet-visible spectroscopy serves as the primary technique for detecting and characterizing charge-transfer complexes, revealing new absorption bands absent in the spectra of the individual donor or acceptor components. The position of the CT band maximum (λ_max) correlates approximately with the difference in ionization potential of the donor (IP_D) and electron affinity of the acceptor (EA_A), following the relation λ_max ≈ 1240 / (IP_D - EA_A) nm, where energies are in eV and the coulombic stabilization term is neglected for simplicity. This empirical approximation, rooted in Mulliken's excited-state model, aids in predicting spectral features based on donor-acceptor energetics. Infrared (IR) spectroscopy reveals shifts in the vibrational modes of donor and acceptor molecules upon complexation, typically due to changes in bond strengths from partial charge redistribution. For instance, acceptor carbonyl stretches often shift to lower frequencies, indicating increased electron density. Raman spectroscopy complements IR by highlighting stacking interactions in solid-state complexes, where charge-transfer effects can enhance scattering intensities for modes sensitive to intermolecular geometry. Electron spin resonance (ESR) spectroscopy is particularly useful for studying radical ions formed in conducting charge-transfer solids, detecting unpaired electrons in donor radical cations or acceptor radical anions with hyperfine splitting patterns that reveal the extent of charge delocalization. Solvatochromism in CT bands, manifested as shifts in λ_max with changing solvent polarity, underscores the ion-pair character of the ; polar solvents stabilize the charge-separated more than the , often leading to bathochromic shifts with increasing solvent polarity. This sensitivity provides insights into the polarity and dynamics of the complex in solution.

Electrical and Magnetic Properties

Charge-transfer complexes exhibit a range of electrical behaviors depending on the degree of charge (δ) and molecular stacking arrangement. In mixed-stack configurations, where donor and acceptor molecules alternate along the stack axis, complexes with weak charge (δ < 0.1) are typically insulating due to localized charge carriers and large gaps. Semiconducting arises in these systems through activated , where charge carriers are thermally excited across the band , as observed in many donor-acceptor complexes. In contrast, segregated-stack structures, such as those in tetrathiafulvalene-tetracyanoquinodimethane (TTF-TCNQ), enable metallic conduction via band-like when δ exceeds approximately 0.5, leading to partial filling of the conduction band by transferred electrons. The band model for these complexes describes how partial charge transfer populates the conduction band, facilitating electron delocalization along molecular chains. In organic systems, charge carrier mobility typically ranges from 1 to 10 cm²/V·s, influenced by intermolecular overlaps and interactions that broaden the . Recent advances in CT complexes have achieved mobilities exceeding 10 cm²/V·s through optimized stacking and reduced disorder. The transition between metallic and insulating states is governed by stack dimensionality and δ; one-dimensional chains are particularly susceptible to instabilities, while higher dimensionality stabilizes metallic behavior by reducing electron-phonon coupling effects. For instance, in TTF-TCNQ, metallic conduction persists above the Peierls transition temperature but gives way to an insulating state below 38 K due to lattice distortion and gap opening in the one-dimensional stacks. Magnetic properties of charge-transfer complexes often stem from unpaired spins in radical-ion components. Radical-ion salts, such as those involving the tetracyanoquinodimethane anion (TCNQ⁻), display arising from the S = 1/2 of the , leading to Curie-like at high temperatures. In certain solids with stronger interactions, antiferromagnetic ordering emerges, as seen in TTF-chloranil (TTF-CA) below its Néel temperature, where coupling aligns antiparallel. These magnetic behaviors are modulated by the same factors influencing , with partial δ promoting spin delocalization in metallic phases. Emerging research explores these properties for spintronic applications, leveraging tunable in low-dimensional CT systems as of 2024.

Applications and Mechanistic Roles

In Materials and Electronics

Charge-transfer complexes have found significant applications in materials science and electronics due to their tunable electrical conductivity and optoelectronic properties. Organic conductors based on these complexes, such as tetrathiafulvalene-tetracyanoquinodimethane (TTF-TCNQ), exhibit metallic-like conductivity, making them suitable for low-cost flexible electronics. For instance, TTF-TCNQ thin films demonstrate high conductivity (approximately 30 S/cm) and transparency, enabling their use in freestanding organic wires and interconnects for wearable devices. Analogs of TTF-TCNQ have been integrated into flexible sensors, where their charge transport facilitates responsive detection layers for environmental monitoring. In , charge-transfer complexes play a key role in organic solar cells by promoting efficient dissociation at donor-acceptor interfaces. Donor-acceptor blends, such as those involving diimides or acceptors, form charge-transfer states that enable charge separation, contributing to power conversion efficiencies of 5-10% in bulk architectures. These complexes enhance light harvesting and reduce recombination losses, positioning them as vital components in low-cost photovoltaic devices. Charge-transfer complexes are also employed in sensing applications, leveraging their color changes upon interaction with analytes. The classic iodine-starch complex, formed between iodine and helices, produces a distinctive color due to charge-transfer interactions, enabling simple detection of in biological and food samples. Similarly, tetracyanoethylene (TCNE)-based complexes with electron donors exhibit colorimetric responses to nitroaromatic , where from the donor to the explosive acceptor induces visible spectral shifts for rapid, on-site detection. In catalytic materials, charge-transfer complexes enhance reactivity in by facilitating electron donation and separation. For example, (III) complexes forming charge-transfer pairs with substrates promote selective C-H activation under visible light, improving efficiency in transformations. In gas sensing, these complexes enable sensitive NO2 detection; exposure to NO2 leads to , increasing sensor resistance and allowing ppb-level monitoring at . Recent advancements in the have integrated charge-transfer complexes with and materials to boost charge separation in optoelectronic devices. In perovskite heterostructures, such as those mixing phenylethylammonium lead and tin , charge-transfer excitons form at interfaces, enhancing and for light-emitting diodes (LEDs). For transistors, perovskites with enhanced interlayer charge transport, such as fluorinated lead- structures, achieve field-effect mobilities up to approximately 0.002 cm²/V·s, supporting with improved charge transport. These hybrid systems benefit from inherent electrical conductivity, aiding device performance without additional doping.

In Chemical Reactions

Charge-transfer complexes play a crucial role as intermediates in () reactions, particularly as π-complex precursors that facilitate the approach of the to the aromatic ring. In the bromination of , for instance, the arene acts as a π-donor forming a charge-transfer complex with Br₂, which stabilizes the and lowers the for the subsequent formation of the σ-complex (Wheland intermediate). This preorganization step enhances reaction efficiency, as evidenced by spectroscopic detection of the CT complex and kinetic studies showing its influence on the rate-determining . Similarly, in the tricyanovinylation of N,N-dimethylaniline with tetracyanoethylene (TCNE), the initial π-complex exhibits partial charge transfer, with a barrier of approximately 79 kJ/mol in solution, leading to regioselective substitution at the para position. In nucleophile-electrophile interactions, transient charge-transfer states mediate key steps in reactions such as Grignard additions to carbonyls and halogenation of amines. For Grignard reagents reacting with ketones, from the organomagnesium species to the carbonyl oxygen forms a transient state, initiating (SET) and leading to radical-ion intermediates that collapse to the addition product; this pathway correlates with the ionization potentials of the reactants and explains solvent-dependent rates. In the halogenation of amines, charge-transfer complexes between amines (e.g., ) and halogens like or Br₂ exhibit n→σ* donation, facilitating halogen atom transfer and influencing the of N-halogenation or oxidation products, as supported by and theoretical analysis of the intermolecular forces. Photochemical processes involving charge-transfer complexes often proceed via to or energy-transfer states, particularly in dye-sensitized systems. Upon of , intramolecular charge in ruthenium-based dyes leads to rapid injection into a (e.g., TiO₂), effectively dissociating the charge pair and enabling energy for subsequent reactions like ; this occurs on picosecond timescales, with efficiency tied to the dye's donor-acceptor architecture. In oxidation-reduction reactions, CT complexes serve as precursors to SET mechanisms, where partial in the evolves into full upon activation, as seen in inner-sphere processes with organic donors and acceptors, promoting formation in synthetic transformations. These roles have broader implications for understanding and reaction rates in . For example, iodine-olefin charge-transfer complexes in reactions to alkenes (e.g., isomers) dictate stereochemical outcomes and , with the CT interaction stabilizing the bridged iodonium intermediate and accelerating trans addition over cis pathways, as determined by rate studies in nonpolar solvents. Overall, CT complexes provide a mechanistic framework for predicting substituent effects and optimizing synthetic routes.

References

  1. [1]
    charge-transfer complex (C01003) - IUPAC Gold Book
    An electron-donor–electron-acceptor complex, characterized by electronic transition(s) to an excited state in which there is a partial transfer of electronic ...
  2. [2]
    Molecular Compounds and their Spectra. II - ACS Publications
    Article February 1, 1952. Molecular Compounds and their Spectra. II. Click to copy article linkArticle link copied! Robert S. Mulliken. ACS Legacy Archive. Open ...
  3. [3]
    Charge‐Transfer Complexes: Fundamentals and Advances in ...
    Jul 24, 2024 · CTCs are defined as association of two or more electron donor (D) and electron acceptor (A) molecules or ions, forming a packed and well-ordered ...
  4. [4]
    Charge-transfer complexes: new perspectives on an old class of ...
    This manuscript reviews the growth, structure and properties of charge-transfer complexes and underlines recent progress in their application in organic devices ...
  5. [5]
    Molecular Compounds and their Spectra. III. The Interaction of ...
    Charge Transfer Complex between O-Phenylenediamine and 2, 3-Dichloro-5, 6-Dicyano-1, 4-Benzoquinone: Synthesis, Spectrophotometric, Characterization ...
  6. [6]
    Charge transfer complexes: a review survey - ScienceDirect
    Charge transfer complex (CTC) is a term used in chemistry to describe interactions between molecules or ions. A CTC occurs when an electron moves from one ...<|control11|><|separator|>
  7. [7]
    [PDF] Understanding the charge transfer at the interface of electron donors ...
    interaction between the two frontier orbitals and the molecular overlap. Such interaction results in partial charge transfer at a small d. As d increases ...
  8. [8]
    [PDF] Organic Charge Transfer Systems: the Next Step in Molecular ...
    Dec 13, 2013 · The actual degree of charge transfer between donor and acceptor is given as a fraction of an elemental charge δ per donor-acceptor pair. As ...
  9. [9]
    Mixed valency in organic charge transfer complexes - Journals
    Sep 10, 2007 · and that of 0.5>δ, δ′≥δc by M′, where δc is the critical degree of CT ... Charge-Transfer Complex, The Journal of Physical Chemistry A ...Mixed Valency In Organic... · 3. Ionicity Phase Diagram · (b) Mixed-Valence Complexes...
  10. [10]
    Mixed-vs-Segregated Stack Polymorphism in the N,N,N
    Apr 28, 2025 · Most CT crystals have either a segregated or mixed stack motif: in segregated stacks, D and A molecules form separate columns, while in mixed ...
  11. [11]
    TDDFT study of the model charge-transfer complexes related to the ...
    The total energy of complexation (including OH…O=C< bonding) is 15.6 kcal/mol. Spectroscopic studies of the acyl-CoA dehydrogenase model with the redox-inactive ...
  12. [12]
    [PDF] Complexation and Protein Binding Complexes or coordination ...
    In both charge transfer and donor–acceptor complexes, Charge transfer complexes are of importance in pharmacy. Iodine forms 1:1 charge transfer complexes with ...
  13. [13]
    Use of the molecular orbital theory of charge transfer spectra in ...
    . A review on electronic spectral studies of charge transfer complexes. Journal of Molecular Structure 2013, 1034 , 393-403. https://doi.org/10.1016/j ...
  14. [14]
    MP2 and DFT study of IR spectra of TCNE‐methylsubstituted ...
    Sep 9, 2009 · The calculated interplanar distance, dipole moment, and transferred charge are approximately linearly dependent on number of methyl groups N.
  15. [15]
    Effect of Pressure on Charge Transfer Complexes ... - RSC Publishing
    should cause large changes in the stability constant K and the charge-transfer (CT) energy (hv,,). The complexes should thus be sensitive to hydrostatic ...
  16. [16]
    Charge-Transfer Benzoquinone:Catechol Complexes, Their Aging ...
    Feb 17, 2023 · Quinhydrone (Qh), the first charge-transfer complex discovered in 1844 by Wöhler, (1) was also the ...<|control11|><|separator|>
  17. [17]
    [PDF] Series of charge transfer complexes obtained as crystals in a ... - HAL
    Oct 5, 2021 · The charge transfer band lays between 430 and 530 nm. The energy of this CT band is correlated with the reduction potential of the ...
  18. [18]
    Regularities in the Infrared Spectra of Picric Acid Molecular ...
    Charge–transfer complexes of pyrimidine Schiff bases with aromatic nitro compounds. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 2011 ...<|separator|>
  19. [19]
    Molecular complexes of halogens and interhalogen compounds with ...
    Gur'yanova, “The problem of the nature of the intermolecular bonds in charge-transfer complexes,” Zh. Org. Khim.,36, No. 7, 1157–1165 (1966). Google Scholar.<|control11|><|separator|>
  20. [20]
    The Iodine/Iodide/Starch Supramolecular Complex - PMC - NIH
    Three types of possible sources of charge transfer have been proposed: (1) chains of neutral I2 molecules, (2) chains of poly-iodine anions (complicated by the ...
  21. [21]
    Why Does Iodine Turn Starch Blue? - ChemistryViews
    Dec 6, 2016 · Electrons in such charge-transfer complexes are easy to excite to a higher energy level by light. The light is absorbed in the process and its ...
  22. [22]
    (PDF) Iodination of Phenols in Water using Easy to Handle Amine ...
    Aug 6, 2025 · In this work we describe, for the first time, the use of complexes morpholine-iodo, N-methyl-piperazine-iodo and thiomorpholine-iodo as ...
  23. [23]
    A solvent-resistant halogen bond - RSC Publishing
    Jul 30, 2014 · The stabilities of the H-bonded complexes decrease by more than three orders of magnitude with increasing solvent polarity. In contrast, the X- ...
  24. [24]
    [PDF] Metallic Conduction at Organic Charge-Transfer Interfaces - arXiv
    2, TTF and TCNQ crystals are semiconductors, whose electronic properties are similar to those of many other organic materials (e.g., pentacene) commonly used in ...
  25. [25]
    Temperature Dependence of Conductivity of Tetrathiafulvalene ...
    Aug 12, 1974 · The conductivity of single crystals of tetrathiafulvalene-tetracyanoquinodimethane (TTF-TCNQ) was measured over the temperature range 40-300 K.
  26. [26]
    IN TTF-TCNQ - Journal de Physique
    34 which is big and important in making the Peierls transition temperatures of the TCNQ and TTF chains close together (48 K and 54 K; see section 4). Thus ...
  27. [27]
    Direct Evaluation of the Charge Transfer in the Tetrathiafulvalene ...
    Jul 14, 1975 · The results indicate an average charge transfer of (0.48-0.60) ± 0.15 electrons from a TTF to a TCNQ molecule. References (14). C. J. Fritchie ...Missing: degree | Show results with:degree
  28. [28]
    [PDF] Superconductivity in a synthetic organic conductor (TMTSF)2PF6 (~)
    (TMTSF)2PF6 (and related salts) exhibits low tem- perature conductivities in excess of 105 (0. cm) - I , which were the highest reported so far in organic.
  29. [29]
    The Physics of Organic Superconductors - Science
    The upper temperature for superconductivity in organic conductors has increased from 1 kelvin in 1980, when the phenomenon was discovered.
  30. [30]
    Doping: A Key Enabler for Organic Transistors - Xu - 2018
    Aug 13, 2018 · Thus, doping can enhance mobility of OFETs with highly disordered OSCs thanks to the lowered hopping barrier and passivated charge traps with ...
  31. [31]
    [PDF] MULLIKEN'S THEORY IN CHARGE-TRANSFER COMPLEXATION
    The nature of the attraction in a charge-transfer complex is not a stable chemical bond and is much weaker than covalent forces, rather it is better ...
  32. [32]
    [PDF] UV-Visible Spectroscopic Studies of Charge- Transfer Complexes
    Mulliken was responsible for the development of what has been the most successful theoretical treatment of charge-transfer complexes. The absorption of ...
  33. [33]
    Interaction of Oxygen with Conjugated Polymers: Charge Transfer ...
    The energy associated with excitation of a charge transfer complex is given by. where IPD and EAA are the ionization potential of the donor and the electron ...
  34. [34]
    [PDF] Reference Energies for Intramolecular Charge-Transfer Excitations
    Apr 12, 2021 · ΔECT = IPD − EAA − 1/R,. (1) where IP. D is the first ionization ... The Electronic. States of Nitrobenzene: Electron-Energy-Loss Spec- troscopy ...
  35. [35]
    The Effect of Oxygen on the Polymerization of Acrylonitrile1
    **Summary of Key Points on Spectroscopic Aspects:**
  36. [36]
    Infrared Spectra of Charge-transfer Complexes. III. Complexes with ...
    The Raman spectra of the benzene—chlorine and benzene—bromine charge-transfer complexes. Spectrochimica Acta Part A: Molecular Spectroscopy 1977, 33 (11) ...
  37. [37]
    Raman spectroscopy of intermolecular charge transfer complex ...
    Sep 14, 2007 · Raman spectroscopy of intermolecular charge transfer complex between a conjugated polymer and an organic acceptor molecule Available. V. V. ...Results · Raman Scattering · Samples And Experimental...
  38. [38]
    Large solvatochromism of metal-to-ligand charge-transfer transitions ...
    Medium Effects on Charge Transfer in Metal Complexes. Chemical Reviews 1998, 98 (4) , 1439-1478. https://doi.org/10.1021/cr941180w. Jon R. Schoonover ...
  39. [39]
    [PDF] molecules - Semantic Scholar
    Dec 18, 2019 · Significant blue-shift of the CT bands was observed when solvent polarity increased. This phenomenon corresponds to negative solvatochromism [41] ...Missing: epsilon | Show results with:epsilon
  40. [40]
    [PDF] Studies of an Electrically Conducting Organic Material
    TTF and TCNQ form the charge transfer complex TTF-TCNQ by simple contact ... A researcher in 1977 calculated the mobility of TTF-TCNQ at room temperature as 3 cm2 ...Missing: cm²/ | Show results with:cm²/
  41. [41]
    The 38 K transition in TTF-TCNQ viewed as a percolation ...
    The model is the first to reproduce the observed metal-insulator transition (e.g. the d.c. conductivity anomaly) near 38 K. The dielectric constant is ...
  42. [42]
    Electrochemical synthesis, characterization and magnetic studies of ...
    Two new materials, Ni(TCNQ) and Ni3(TCNQ)2, have been obtained by an electroplating technique and characterized by SEM, XPS, vibrational spectroscopy, ...
  43. [43]
    Paramagnetic Molecular Semiconductors Combining Anisotropic ...
    We report the synthesis, magnetic properties, and transport properties of paramagnetic metal complexes, [Co(DMF)4(TCNQ)2](TCNQ)2 (1), ...
  44. [44]
    [PDF] Highly Conductive, Transparent Molecular Charge-Transfer Salt with ...
    Specially, the high electrical conductivity (~ 30 S/cm) and transparency (~80%) of. TTF-TCNQ thin films allow us develop its application in all-organic.
  45. [45]
    Assessing the nature of the charge-transfer electronic states in ...
    Dec 13, 2018 · The charge-transfer electronic states appearing at the donor-acceptor interfaces in organic solar cells mediate exciton dissociation, charge generation, and ...
  46. [46]
    Charge Transfer and Interface Effects in Co‐Assembled Circular ...
    Feb 12, 2019 · Power conversion efficiency of ≈10% under 1 sun illumination has been offered from time-domain drift diffusion model. 1 Introduction. In organic ...
  47. [47]
    On the Origin of the Blue Color in The Iodine/Iodide/Starch ... - MDPI
    Dec 16, 2022 · On the other hand, I2-In- mixtures afford significantly stronger charge-transfer bands, from the Ix- σ* orbitals (HOMO) to I2 σ* (LUMO) orbitals ...
  48. [48]
    Theoretical study of the Meisenheimer and charge-transfer ...
    In this paper, we present a theoretical study of structure and UV-vis spectra of 11 colored complexes of nucleophiles with nitroaromatic energetic materials.
  49. [49]
    Photocatalytic Effect of a Charge-Transfer Complex Formed After ...
    Aug 14, 2025 · In this study, we synthesized a new Ir(III) complex that is resistant to photoinduced modifications by the use of sterically bulky NBP ...Introduction · Results and Discussion · Supporting Information · References
  50. [50]
    Establishing charge-transfer excitons in 2D perovskite heterostructures
    May 26, 2020 · We report that CTEs can be effectively formed in heterostructured 2D perovskites prepared by mixing PEA 2 PbI 4 :PEA 2 SnI 4 , functioning as host and guest ...Missing: transistors | Show results with:transistors
  51. [51]
    Enhancing Interlayer Charge Transport of Two-Dimensional ...
    Two-dimensional lead-halide perovskites provide a more robust alternative to three-dimensional perovskites in solar energy and optoelectronic applications ...
  52. [52]
    [PDF] Charge Transfer in 2D Halide Perovskites and 2D/3D Heterostructures
    ABSTRACT: Halide perovskites have emerged as a versatile class of materials, with applications spanning photovoltaics, light-emitting diodes, transistors, and ...
  53. [53]
    and π-Complexes in Electrophilic Aromatic Substitutions
    Charge-Transfer Complexes as Precursors to Electrophilic Aromatic Substitutions. ... Charge transfer facilitated direct electrophilic substitution in ...<|separator|>
  54. [54]
  55. [55]
  56. [56]
    A theoretical characterization of the trends in halogen bonding
    Aug 7, 2003 · The intermolecular interaction in the charge-transfer complexes between amines and halogens: A theoretical characterization of the trends in ...
  57. [57]
    Photoexcited Intramolecular Charge Transfer in Dye Sensitizers
    Efficient photoinduced intramolecular charge transfer (ICT) from donor to acceptor in dye molecules is the functional basis and key property in the working ...Introduction · Results and Discussion · Supporting Information · Author Information
  58. [58]
    Reactions of iodine with olefins. I. Kinetics and mechanism of iodine ...
    Mechanistic studies of atom transfer addition reactions of iodomalononitriles to alkenes. 1,2-asymmetric induction in an iodine atom transfer reaction..