Fact-checked by Grok 2 weeks ago

Electron transfer

Electron transfer (ET) is the process in which an electron moves from a donor , , or to an acceptor , , or , serving as the core mechanism underlying reactions without necessarily involving the formation or breaking of covalent bonds. This nonradiative relocation of charge is a fundamental elementary step in oxidation-reduction chemistry, enabling the conversion of into electrical or mechanical forms. ET reactions occur in diverse environments, from homogeneous solutions to biological interfaces, and are characterized by their , which depend on factors such as the driving force, reorganization energy, and electronic coupling between donor and acceptor. The theoretical foundation for understanding ET rates was established by in the 1950s and 1960s, culminating in his in 1992 for developing a framework that predicts reaction rates based on classical and quantum mechanical principles. posits that the arises primarily from the reorganization of solvent molecules and intramolecular vibrations to accommodate the changed charge distribution post-transfer, leading to parabolic relationships with distinct normal and inverted regions. This model has been validated across inner-sphere (bond-mediated) and outer-sphere (through-space) mechanisms, with experimental confirmation through techniques like ultrafast spectroscopy. ET processes are pivotal in natural and technological systems, powering biological in —where electrons flow through protein complexes to generate ATP—and , as well as enabling applications in batteries, solar cells, and corrosion prevention. In synthetic chemistry, controlled ET facilitates and , while in , it underpins molecular electronics and dye-sensitized photovoltaic devices with efficiencies tied to rapid charge separation. Ongoing research integrates quantum effects and proton-coupled variants to enhance conversion efficiencies in sustainable technologies.

Fundamentals

Definition and Importance

Electron transfer refers to the relocation of an from one , known as the donor, to another, the acceptor, which results in the oxidation of the donor and the of the acceptor without any net displacement of atoms between the species. This process lies at the heart of reactions, where the of electrons drives chemical transformations across diverse systems. The systematic conceptualization of electron transfer emerged in the 1950s through detailed studies of inorganic redox reactions, particularly in metal complexes. Pioneering experiments by elucidated mechanisms such as inner-sphere electron transfer, where a facilitates the process, earning him the in 1983 for these foundational contributions. Electron transfer holds profound importance across multiple disciplines. In , it enables energy storage and conversion in batteries, where electrons shuttle between and during charge-discharge cycles to power devices. In , it is essential for cellular energy production, notably in the mitochondrial , where approximately 80% of ATP yield from glucose oxidation in aerobic —around 30 of 36 ATP molecules—arises from electron transfers coupled to proton gradients. In , electron transfer is critical for photovoltaic devices like solar cells, facilitating charge separation upon light absorption to generate electricity. The of electron transfer are governed by potentials, which quantify the tendency of a to gain or lose electrons. For instance, the standard reduction potential for the oxygen-water (O₂ + 4H⁺ + 4e⁻ → 2H₂O) is +1.23 V versus the , highlighting oxygen's role as a potent oxidant in many natural and engineered processes.

Basic Principles

Electron transfer (ET) is fundamentally a quantum mechanical process in which an tunnels from a donor orbital to an acceptor orbital, typically the highest occupied (HOMO) of the donor and the lowest unoccupied (LUMO) of the acceptor. This tunneling occurs without classical activation over a barrier, allowing ET over distances up to several nanometers in suitable media, with the rate depending exponentially on the donor-acceptor separation. The electronic coupling that facilitates this transfer requires sufficient spatial overlap between the donor and acceptor orbitals to enable the tunneling probability. A key aspect of is the role of nuclear reorganization governed by the Franck-Condon principle, which posits that the electron transfer is much faster than nuclear motion, leading to a vertical transition on the . Following the electronic jump, the nuclei must relax to the equilibrium geometry of the product state, incurring a reorganization energy λ that includes contributions from intramolecular vibrations and . are critical, as the screening modulates the electrostatic interactions between charged , influencing both the driving force and the reorganization barrier. ET pathways can be classified as adiabatic, where strong electronic coupling allows the system to follow the lower , or non-adiabatic, where weak coupling results in probabilistic tunneling between diabatic states. The thermodynamics of ET are described by the standard free energy change, given by \Delta G^\circ = -n F E^\circ where n is the number of electrons transferred, F is the Faraday constant, and E^\circ is the standard electrode potential difference between donor and acceptor. The kinetics involve an activation barrier arising from the reorganization energy λ, which must be overcome for the nuclear coordinates to reach the transition state. In biological systems, such as photosynthetic reaction centers, these principles enable efficient energy transduction over long distances. ET reactions are distinguished as homonuclear self-exchange (e.g., between identical couples) or heteronuclear cross-reactions (between different couples). Self-exchange reactions, where ΔG° ≈ 0, serve as benchmarks for understanding reorganization without driving force contributions; for instance, the aqueous Fe³⁺/Fe²⁺ self-exchange has a rate constant of approximately 4 M⁻¹ s⁻¹ at 25°C, reflecting modest inner-sphere reorganization due to similar ligand fields.

Types of Electron Transfer

Inner-Sphere Electron Transfer

Inner-sphere electron transfer is a in which an moves between two metal centers connected by a , forming a covalent linkage in the . This process typically proceeds through three stages: the formation of a precursor where the bridge is established, the actual electron transfer across the bridge, and the dissociation of the successor . Unlike direct electron tunneling, the facilitates the transfer by providing an orbital pathway, often involving partial bond breaking or making during the reaction. A key characteristic of inner-sphere electron transfer is the involvement of bond-making and bond-breaking steps, which distinguish it from non-bridged pathways. These reactions are frequently rate-limited by the initial to form the bridged precursor, particularly when one reactant is labile. The can be , pseudohalide, or organic groups capable of coordinating both metals. Representative activation parameters, such as low enthalpies of (ΔH‡ ≈ 10 kcal mol⁻¹) and negative entropies of (ΔS‡ ≈ -25 cal mol⁻¹ K⁻¹), support an associative with restricted degrees of freedom in the . The seminal example is the reduction of chloropentaamminecobalt(III) by hexaquachromium(II): [\ce{(NH3)5CoCl}]^{2+} + \ce{Cr(H2O)6}^{2+} \rightarrow [\ce{(NH3)5Co(H2O)}]^{3+} + \ce{Cr(H2O)5Cl}^{2+} Here, the chloride ligand bridges the Co(III) and Cr(II) centers, transferring from cobalt to chromium during the reaction. This system proceeds rapidly with a second-order rate constant of 6 × 10⁶ M⁻¹ s⁻¹ at 25 °C, following the rate law rate = k [[Co(III)]][[Cr(II)]]. Experimental evidence for the bridged pathway came from studies in the 1950s by , who used radioactive (Cl*) and observed that the in the chromium product originated from the cobalt complex rather than the solvent, confirming atom transfer via the bridge. These studies, along with comparisons to non-bridging ligands like (which yield much slower rates), established the inner-sphere mechanism. Discrimination between inner-sphere and outer-sphere mechanisms often relies on stereochemical probes; for instance, retention of at the metal center in bridged transfers versus no stereochemical change in direct outer-sphere processes provides diagnostic evidence for involvement.

Outer-Sphere Electron Transfer

Outer-sphere electron transfer involves the tunneling of an between two metal centers while their coordination spheres remain intact and unchanged, resulting in no net chemical transformation beyond the movement of the itself./V%3A__Reactivity_in_Organic_Biological_and_Inorganic_Chemistry_3/01%3A_Reduction_and_Oxidation_Reactions/1.09%3A_Outer_Sphere_Electron_Transfer) This process occurs without the formation or breaking of any bonds, distinguishing it as a purely event mediated by overlap of molecular orbitals across the intervening space, often molecules. The reactants approach each other diffusively, and the electron transfer happens at close but non-contact distances, typically on the order of 7–12 . The mechanism proceeds through a five-step sequence. First, the oxidant and reductant diffuse together to form an encounter complex, held by electrostatic or weak . Second, reorganization occurs, involving adjustments in bond lengths and angles within each , accompanied by reorganization to accommodate the changing charge distribution. Third, the tunnels quantum mechanically from the donor to the acceptor orbital, facilitated by through the solvent barrier. Fourth, the successor complex relaxes as the nuclei and adapt to the new charge states. Finally, the products dissociate and diffuse apart. These steps ensure that the activation barrier arises primarily from the reorganization energy required for the . Representative examples include the self-exchange reaction between and ions, [Fe(CN)6]3– + [Fe(CN)6]4– ⇌ [Fe(CN)6]4– + [Fe(CN)6]3–, which exhibits a second-order rate constant of approximately 103 M–1 s–1 at 25 °C in . Another classic case is the self-exchange between and ions, MnO4 + MnO42– ⇌ MnO42– + MnO4, proceeding via outer-sphere transfer with a rate constant on the order of 102–103 M–1 s–1, as determined by NMR line-broadening methods. These reactions highlight the prevalence of outer-sphere mechanisms in symmetric couples where substitution is unfavorable. The rate of outer-sphere electron transfer is strongly influenced by several factors. Distance plays a critical role, with the rate decaying exponentially as exp(–βr), where β ≈ 1.4 Å–1 reflects the tunneling probability through solvent or vacuum-like barriers, limiting efficient transfer to separations below ~14 Å. Orientation effects arise from the need for favorable alignment of donor and acceptor orbitals to maximize electronic coupling, which can vary with molecular geometry and solvent dielectric properties. Experimental determination of these rates often employs stopped-flow spectroscopy to monitor rapid absorbance changes in the encounter complex formation and electron transfer steps, as demonstrated in studies of ruthenium ammine complexes. Rates can be predicted using Marcus theory, which relates the activation free energy to reorganization and driving force parameters.

Heterogeneous Electron Transfer

Heterogeneous electron transfer refers to the process of exchange occurring at the between two different s, typically involving a in and a such as a metal or . This type of transfer is fundamental in electrochemical systems where the must across the boundary, often influenced by the structure of the and the environment of the . The mechanism of heterogeneous electron transfer is commonly described by the Butler-Volmer equation, which relates the to the and accounts for both anodic and cathodic processes at the . In this framework, the (η), defined as the difference between the applied potential and the equilibrium potential, drives the reaction rate, while charge transfer resistance (R_ct) quantifies the kinetic barrier to electron exchange. For solution-side aspects, this process can connect to outer-sphere principles where minimal reorganization occurs in the redox species. Key concepts influencing heterogeneous electron transfer include the structure of the electric double layer at the interface, which modulates the local potential and affects the energy barrier for electron tunneling. Adsorption of reactants onto the surface often follows the Langmuir isotherm, assuming coverage without lateral interactions, which is critical for reactions involving surface-bound intermediates. Rate analysis frequently employs Tafel plots, derived from the high- limit of the Butler-Volmer equation, where the overpotential η is linearly related to the logarithm of the i: \eta = a + b \log i Here, a is a constant related to the , and b (the Tafel slope) reflects the symmetry of the energy barrier, typically around 120 mV/decade for a single electron with a coefficient of 0.5. Representative examples illustrate these principles. In the (HER) on electrodes, heterogeneous electron facilitates the reduction of protons to gas, with Pt's low (near 0 V vs. ) arising from optimal adsorption and fast kinetics. Similarly, in -sensitized cells, electron injection from an excited to the conduction of a TiO₂ exemplifies photoinduced heterogeneous , enabling efficient charge separation with rates exceeding 10^9 s^{-1} due to favorable alignment. Experimental characterization of heterogeneous electron transfer relies on techniques like (), which sweeps the and measures the resulting current to determine the standard heterogeneous rate constant k⁰. In , the peak separation between anodic and cathodic waves increases with slower kinetics, allowing k⁰ to be extracted via methods such as Nicholson's working curve, providing quantitative insight into the transfer barrier.

Vectorial Electron Transfer

Vectorial electron transfer describes the unidirectional propagation of electrons across a series of redox-active centers in multi-center molecular systems, such as proteins or synthetic assemblies, propelled by gradients or asymmetric spatial arrangements that minimize back-transfer. This process is essential in constrained environments like biological membranes, where it enables efficient energy conversion without diffusive loss. The mechanism predominantly involves tunneling, in which the virtually couples donor and acceptor sites through intervening bridge units—such as σ-bonds in peptides or π-conjugated systems in aromatic residues—without significant occupation of bridge states. The transfer rate decays exponentially with donor-acceptor distance r, following k \propto \exp(-\beta r), where the attenuation factor \beta ranges from approximately 0.6 to 1.4 Å⁻¹ depending on the bridge composition and medium polarity. For instance, β values near 1.1 Å⁻¹ are common in β-sheet protein motifs, reflecting moderate tunneling barriers through hydrogen-bonded networks. Prominent examples include the sequential electron hopping between [4Fe-4S] clusters in , where low reorganization energies (<0.5 eV) and weak electronic coupling facilitate vectorial flow over ~10 Å distances, as observed in Chromatium vinosum ferredoxin variants. In DNA, vectorial charge transfer often manifests as hole migration along the base stack, culminating in guanine radical formation at low-potential sites (e.g., GG steps), which supports long-range oxidative damage propagation or repair without significant distance attenuation up to 100 Å. Protein folding significantly influences pathway efficiency by precisely orienting redox centers and selecting optimal tunneling routes, such as through β-strands or aromatic amino acid chains, thereby tuning electronic coupling and reducing off-pathway leakage. Experimental characterization relies on time-resolved spectroscopy techniques, including femtosecond transient absorption, which capture initial transfer events on fs-ps timescales in systems like azurin or photosynthetic reaction centers.

Theoretical Frameworks

Marcus Theory

Marcus theory, developed by Rudolph A. Marcus during the 1950s and 1960s, offers a foundational semiclassical framework for predicting the rates of thermal electron transfer reactions in solution. Central to the theory is the representation of the potential energy surfaces for the reactant and product states as intersecting parabolas in a multidimensional nuclear coordinate space, capturing the harmonic approximation for vibrational modes and solvent polarization. This geometric model highlights how electron transfer occurs most efficiently at the crossing point of these surfaces, where the nuclear configurations minimize the energy barrier, and incorporates the to account for the overlap of vibrational wavefunctions. Marcus received the 1992 Nobel Prize in Chemistry for this work, which has profoundly influenced fields ranging from inorganic chemistry to biochemistry. The rate constant for nonadiabatic electron transfer, k_\text{ET}, is expressed by the following semiclassical formula derived from Fermi's golden rule: k_\text{ET} = \frac{2\pi}{\hbar} |V|^2 \frac{1}{\sqrt{4\pi \lambda k_B T}} \exp\left[ -\frac{(\lambda + \Delta G^\circ)^2}{4 \lambda k_B T} \right] Here, V represents the electronic coupling matrix element between donor and acceptor orbitals, \lambda is the total reorganization energy encompassing inner-sphere (molecular vibrational) and outer-sphere (solvent) contributions, \Delta G^\circ is the standard free energy change of the reaction, k_B is , T is the temperature, and \hbar is the . This equation quantifies how the rate depends on the balance between the driving force -\Delta G^\circ and the energy required to reorganize the system to the transition state. A hallmark prediction of the theory is the parabolic relationship between the activation free energy and the reaction driving force, resulting in a normal region where rates increase with increasing exergonicity (when |\Delta G^\circ| < \lambda) and an inverted region where rates diminish for highly exergonic processes (when |\Delta G^\circ| > \lambda), due to the growing mismatch in nuclear coordinates between initial and final states. The theory also derives linear free energy relationships, known as the Marcus cross-relations, linking self-exchange rate constants (where \Delta G^\circ = 0) to those of cross-reactions, allowing estimation of unknown rates from measurable self-exchange data. Experimental validation of the inverted region came in 1984 through studies of intramolecular electron transfer between and its in rigid organic glasses, where rates decreased markedly for driving forces exceeding 1 eV, confirming the theoretical prediction. In photosynthetic reaction centers, such as those in , the inverted region manifests in charge recombination steps, where highly exergonic transfers exhibit slower kinetics, contributing to efficient forward charge separation by suppressing back reactions. play a crucial role in modulating \lambda, with polar solvents enhancing the outer-sphere component through greater relaxation costs, as quantified by the theory's continuum electrostatic model.

Quantum Mechanical Models

Quantum mechanical models provide a more refined description of electron transfer () processes by accounting for quantum effects in both electronic and nuclear , extending beyond semiclassical approximations. These models treat ET as a nonadiabatic transition between diabatic states, where the weak electronic coupling between donor and acceptor sites allows to be applied. A foundational approach employs to compute the transition rate from an initial state \psi_i to a final state \psi_f, given by k = \frac{2\pi}{\hbar} |\langle \psi_f | \hat{H}' | \psi_i \rangle|^2 \rho(E), where \hat{H}' is the perturbation representing the , and \rho(E) is the of final states at energy E. This expression, originally derived for general quantum transitions, was adapted to ET in polar media by Levich and Dogonadze, who incorporated solvent fluctuations as a bath influencing the Franck-Condon factors. Vibronic coupling models address the interaction between electronic states and nuclear vibrations, treating the latter as a collection of oscillators coupled to the subsystem. In the nonadiabatic limit, where electronic coupling is small compared to energy gaps, the spin-boson model maps ET onto a two-level (donor-acceptor states) interacting with a bosonic bath of vibrational modes, enabling exact solutions via path integrals or numerical methods like the . This framework captures and effects, particularly in condensed-phase ET, where nuclear tunneling and coherence play key roles. Extensions of these models incorporate quantum vibrations more explicitly, as in the Marcus-Levich-Jortner (MLJ) theory, which refines the classical reorganization energy by treating high-frequency intramolecular modes quantum mechanically while retaining classical treatment for low-frequency modes. The resulting rate expression includes Franck-Condon overlaps for quantized vibrations, explaining temperature-dependent activation energies observed in biological , such as in protein matrices. In photoinduced , conical intersections—points of degeneracy between excited and ground electronic states—facilitate ultrafast nonradiative decay and , often competing with charge separation in molecular systems. These intersections enhance efficiency by allowing adiabatic passage without energy loss to vibrations, as evidenced in spectroscopic studies of molecules. Modern computational methods leverage (DFT) and time-dependent DFT (TD-DFT) to evaluate electronic couplings V in the , particularly for complex systems like . For instance, periodic DFT implementations compute transfer integrals in extended solids, revealing how lattice distortions modulate ET in . Recent studies on hybrid nanotube assemblies using TD-DFT have quantified photoinduced ET rates, showing couplings up to 0.1 eV that enable efficient charge migration over nanoscale distances. Similarly, DFT analyses of self-assembled nanostructures in the highlight how surface ligands tune vibronic overlaps, impacting ET in photocatalytic .

Applications

Biological Systems

In biological systems, electron transfer plays a central role in energy production, particularly through the (ETC) in mitochondria during . The ETC facilitates the transfer of electrons from NADH (and FADH₂) to molecular oxygen (O₂), generating a proton gradient across the that drives ATP synthesis. This process involves four main protein complexes: Complex I () accepts electrons from NADH and transfers them to ubiquinone while pumping protons into the ; Complex II () feeds electrons from FADH₂ to ubiquinone without proton pumping; Complex III (cytochrome bc₁ complex) transfers electrons from ubiquinol to and pumps additional protons; and Complex IV () reduces O₂ to water, completing the chain and pumping more protons. The resulting proton motive force powers (Complex V) to produce ATP via , yielding approximately 30-32 ATP molecules per glucose molecule oxidized. A parallel process occurs in photosynthesis within chloroplasts, where light-driven electron transfer follows the Z-scheme, enabling the conversion of into . In this pathway, (PSII) absorbs light to oxidize water, releasing O₂ and providing electrons that are transferred through the cytochrome b₆f complex to , then to (PSI). PSI, upon further light absorption, reduces NADP⁺ to NADPH via . This linear electron flow from H₂O to NADP⁺ establishes a proton gradient across the membrane, which uses to generate ATP. The of this electron transfer process in PSII exceeds 30%, reflecting the highly optimized light-harvesting and charge separation mechanisms evolved in photosynthetic organisms. Protein-mediated electron transfer in these systems often relies on quantum mechanical tunneling, allowing electrons to move between centers over distances up to 14 Å without direct orbital overlap, as exemplified in , which shuttles electrons between Complexes III and IV in the . Evolutionary adaptations have fine-tuned such transfers; for instance, blue copper proteins like plastocyanin in exhibit a distorted trigonal coordination that enables rapid electron transfer rates (up to 10⁶ s⁻¹) over similar distances, optimizing the Z-scheme efficiency while minimizing side reactions. These vectorial transfers are spatially directed along protein chains to ensure unidirectional flow and regulatory control. Dysfunctions in biological electron transfer pathways contribute to various diseases, notably through mitochondrial impairments. Defects in Complex I activity, often due to oxidative damage or genetic mutations, reduce ATP production and increase (ROS), linking to neurodegenerative disorders such as , where neurons show significantly reduced Complex I activity compared to controls. Recent studies on extremophiles, such as thermophilic , reveal resilient electron transfer mechanisms in rubredoxins—small iron-sulfur proteins—that maintain efficiency at temperatures above 80°C via temperature-dependent iron fluctuations, offering insights into stabilizing under stress conditions relevant to disease models.

Technological and Photoinduced Processes

Electron transfer processes are fundamental to electrochemical technologies, including fuel cells and rechargeable batteries, where they enable efficient and conversion at interfaces. In fuel cells, at the and facilitates the oxidation of and of oxygen, respectively, achieving overall efficiencies up to 60% under optimal conditions. Similarly, in lithium-ion batteries, heterogeneous electron transfer between lithium ions and materials, such as graphite and layered , supports rapid charge-discharge cycles with round-trip efficiencies exceeding 90%, minimizing energy losses during intercalation and deintercalation. These processes highlight the role of optimized surfaces in enhancing charge transport and device performance. Heterogeneous electron transfer also plays a key role in prevention strategies, where protective coatings on metals modulate electron flow to inhibit anodic dissolution. For instance, thermal-sprayed anti- coatings, such as those based on or aluminum, alter the heterogeneous electron transfer rates at the metal- , reducing currents by orders of magnitude and extending material lifespan in harsh environments. is central to light-harvesting devices like dye-sensitized solar cells, first demonstrated by O'Regan and Grätzel in 1991 using ruthenium-based dyes adsorbed on nanocrystalline TiO₂ films. In these Grätzel cells, photoexcitation of the dye leads to ultrafast electron injection into the TiO₂ conduction band, followed by regeneration via a , enabling power conversion efficiencies approaching 15% in laboratory settings as of 2025. The mechanism relies on the energetic alignment between the dye's and the semiconductor's band edge, achieving injection yields near 100% within picoseconds to minimize recombination losses. Emerging photoinduced systems leverage electron transfer in quantum dot solar cells, where size-tunable s serve as sensitizers, promoting efficient charge separation at quantum dot-metal oxide interfaces through hot-electron injection or multiple exciton generation. Organic photovoltaics, meanwhile, exploit charge-transfer states at donor-acceptor heterojunctions, such as polymer-fullerene blends, to dissociate s and generate free carriers, with device efficiencies surpassing 18% in recent non-fullerene acceptor designs. In molecular electronics, superexchange-mediated electron transfer through conjugated molecular wires, like oligophenylenevinylene chains bridged between electrodes, enables coherent tunneling over distances up to several nanometers, supporting applications in single-molecule switches and sensors. Recent advances in 2025 have pushed solar cells to certified efficiencies over 27% by engineering electron transport layers and passivation strategies that accelerate interfacial charge extraction while suppressing recombination. Computational design approaches have similarly accelerated the development of electron transfer catalysts for CO₂ reduction, identifying metal-organic frameworks and single-atom catalysts that enhance selectivity toward C₂ products like , with faradaic efficiencies exceeding 70% at low overpotentials.

References

  1. [1]
    Recent Developments in the Electron Transfer Reactions and Their ...
    May 22, 2024 · Electron transfer (ET) occurs when an electron moves from one atom or molecule to another. It's the basis of chemical kinetics, ...
  2. [2]
    Electron Transfer - an overview | ScienceDirect Topics
    Electron transfer is defined as a nonradiative process in which an electron is transferred from a donor to an acceptor without the formation or breaking of ...
  3. [3]
    65 years of electron transfer | The Journal of Chemical Physics
    Jul 8, 2022 · Electron transfer (ET), as a simplest chemical reaction, is the fundamental step in the oxidation–reduction reaction, involved in a large ...
  4. [4]
    Nobel Prize in Chemistry 1992
    ### Summary of Electron Transfer Reactions from Rudolph A. Marcus' Nobel Lecture
  5. [5]
  6. [6]
    Electron Transfer - Stanford Synchrotron Radiation Lightsource
    Jun 16, 2003 · Electron transfer, or the act of moving an electron from one place to another, is amongst the simplest of chemical processes, yet certainly one of the most ...
  7. [7]
    8.3: Electrochemistry- Cells and Batteries - Chemistry LibreTexts
    Apr 15, 2023 · Electrochemistry is important in the transmission ... Redox chemistry, the transfer of electrons, is behind all electrochemical processes.Electrochemical Reactions · Dry Cells (Primary Batteries) · Lead Storage Batteries...
  8. [8]
    [PDF] Henry Taube - Nobel Lecture
    In 1951, an important symposium on Electron Transfer Processes was held at the University of Notre Dame, and the proceedings are reported in J. Phys. Chem., Vol ...
  9. [9]
    DOE Explains...Batteries - Department of Energy
    When the electrons move from the cathode to the anode, they increase the chemical potential energy, thus charging the battery; when they move the other ...
  10. [10]
    Biochemistry, Electron Transport Chain - StatPearls - NCBI Bookshelf
    Sep 4, 2023 · ATP-synthase synthesizes 1 ATP for 4 H+ ions. Therefore, 1 NADH = 10 H+, and 10/4 H+ per ATP = 2.5 ATP per NADH (**some sources round up**).
  11. [11]
    Solar Photovoltaic Cell Basics | Department of Energy
    When the semiconductor is exposed to light, it absorbs the light's energy and transfers it to negatively charged particles in the material called electrons.
  12. [12]
    P1: Standard Reduction Potentials by Element - Chemistry LibreTexts
    Dec 17, 2021 · P1: Standard Reduction Potentials by Element ; O 2 ⁡ ( g ) + 2 ⁢ H + + 2 ⁢ e − ⇌ H 2 ⁡ O 2, 0.695 ; O2(g) + 4H+ + 4e− ⇌ 2H2O(l), 1.229 ; H2O2 + 2H+ ...
  13. [13]
    Intermolecular Electron Transfer by Quantum Mechanical Tunneling
    The transfer of electrons from one molecule to another by quantum mechanical tunneling has recently been implicated in biological electron transport.Missing: orbitals | Show results with:orbitals
  14. [14]
    Long-Range Electron Tunneling | Journal of the American Chemical ...
    Feb 5, 2014 · The decay constant for tunneling through an oligoxylene bridged donor–acceptor pair was found to be 7.6(5) nm–1; (39) tunneling across 2.5 nm of ...
  15. [15]
    Resolving orbital pathways for intermolecular electron transfer - Nature
    Nov 21, 2018 · Recently, HDA has been shown to be roughly proportional to the overlap integral, SDA, between the donor and acceptor frontier molecular orbitals ...
  16. [16]
    Polar Solvent Dynamics and Electron-Transfer Reactions - Science
    In this article the dynamics of solvation in polar liquids and the influence of this dynamics on electron-transfer reactions are discussed.
  17. [17]
    Crossover between the adiabatic and nonadiabatic electron transfer ...
    Jan 19, 2021 · This study shows that in the intermediate regime, the ET kinetic results derived from the adiabatic and nonadiabatic formalisms are nearly identical.
  18. [18]
    Long-Distance Electron Tunneling in Proteins - NIH
    At the moment of level crossing, the Landau–Zener (LZ) transition occurs, in which an electron makes a tunneling jump from the donor to acceptor site. The ...2.2. The Propagator Method · 3.2. Many-Electron Aspect · 3.2. 1. Hartree-Fock...
  19. [19]
    The Cr(II) reduction of pentane-2,4-dionatobis(ethylenediamine ...
    The reaction to give [Cr(bipy)(OH2)4]3+ was characterised by k25 °C = 5.06 × 10−1 s−1 with activation parameters ΔH≠ = 9.9 ± 0.3 kcal/mol, and ΔS≠ = −26.8 ± 0. ...
  20. [20]
    Henry Taube (1915–2005): Electron Transfer - Wiley Online Library
    Jan 17, 2006 · ... inner-sphere electron-transfer reaction between two transition-metal complexes [Eq. (1)]. The genius of the experiment was understanding ...
  21. [21]
    Electron Transfer Between Metal Complexes: Retrospective - Science
    Electron Transfer Between Metal Complexes: Retrospective. Henry TaubeAuthors Info & Affiliations. Science. 30 Nov 1984.Missing: seminal | Show results with:seminal
  22. [22]
    Progress in Understanding Electron-Transfer Reactions at ...
    The second-order rate constant is therefore ket,max = dA(2π/3)ρ3νn. Substitution of dA = 3 × 10-8 cm and ρ = 3 × 10-8 cm into this expression yields ...
  23. [23]
    2. Reversibility – Chemical vs. Electrochemical - Chemistry LibreTexts
    Aug 29, 2023 · Electron transfer to solution species at solid electrodes is by nature heterogeneous. As such, reactions of this type have rate constants that ...
  24. [24]
    [PDF] Lecture 23: Heterogeneous Charge Transfer - MIT OpenCourseWare
    May 9, 2014 · Note that this equation is applicable to any general chemical reaction for which a suitable model of excess chemical potential can be found.
  25. [25]
    Recent Advances in Voltammetry - PMC - NIH
    First (Section 1), the use of the Butler–Volmer equation for modelling electrochemical reactions is considered, specifically in terms of its application to ...
  26. [26]
    The Butler-Volmer equation in electrochemical theory: Origins, value ...
    Sep 1, 2020 · The Butler-Volmer equation is widely used in electrochemical theory to describe the relation between electrode potential (vs. a suitable reference electrode) ...
  27. [27]
    Double-layer effects in the kinetics of heterogeneous electron ...
    Theory for Influence of the Metal Electrolyte Interface on Heterogeneous Electron Transfer Rate Constant: Fractional Electron Transferred Transition State ...
  28. [28]
    Tafel Slope Plot as a Tool to Analyze Electrocatalytic Reactions
    In electrocatalysis, Tafel slopes or the corresponding charge-transfer coeffs. provide valuable information about reaction mechanisms and are routinely extd.
  29. [29]
    Hydrogen Evolution Reaction Activity of Heterogeneous Materials
    Aug 17, 2020 · In this study, we present a new comprehensive methodology to quantify the catalytic activity of heterogeneous materials for the hydrogen evolution reaction ( ...Introduction · Methods · Results · Conclusions
  30. [30]
    Analysis of Electron Transfer Properties of ZnO and TiO 2 ...
    Feb 19, 2014 · In this work, we made a one-to-one comparison on the performance of ZnO and TiO2 photoanodes for dye-sensitized solar cells. Different ...
  31. [31]
    A Practical Beginner's Guide to Cyclic Voltammetry - ACS Publications
    Nov 3, 2017 · CrEr mechanism: the faster the forward rate constant of the Cr step, the more reversible the voltammogram (Keq = 0.1, kf = 1 (blue), 10 (dark ...
  32. [32]
    Determination of heterogeneous electron transfer rate constants at ...
    Rate constants were calculated using cyclic voltammetry with the Nicholson method and steady-state currents at ultramicroelectrode arrays.<|separator|>
  33. [33]
    Photoinduced electron transfer across molecular bridges
    Mar 6, 2014 · This review discusses photoinduced electron transfer using superexchange, the role of molecular bridges, and differences between thermal and ...
  34. [34]
    Long-range electron transfer - PNAS
    Solid line indicates theoretical distance dependence for a single-step, ergoneutral (ΔGRP = 0) tunneling process (β = 1.1 Å-1).
  35. [35]
    Superexchange Coupling and Electron Transfer in Homogeneous ...
    The superexchange coupling (VS) between two distant redox centers in large three-dimensional disordered media is studied by considering an electron-transfer ...
  36. [36]
    Distance dependence of electron-transfer reactions in organized ...
    Distance dependence of electron-transfer reactions in organized systems: the role of superexchange and non-Condon effects in photosynthetic reaction centers.
  37. [37]
    Intramolecular electron transfer in [4Fe-4S] proteins - PubMed
    This set of parameters, determined for the first time in an iron-sulfur protein, may help to explain how efficient vectorial electron transfer occurs with a ...
  38. [38]
    Biochemistry and Theory of Proton-Coupled Electron Transfer
    Apr 1, 2014 · Studies of Trp oxidation in proteins may have particular relevance for guanine oxidation in DNA, where long-distance radical hopping along ...
  39. [39]
    Electron-Transfer Acceleration Investigated by Time Resolved ...
    ET can be accelerated by enhancing the electronic interaction and by vibrational excitation of the reacting system and its medium.
  40. [40]
    On the Theory of Oxidation‐Reduction Reactions Involving Electron ...
    A quantitative theory of the rates of oxidation‐reduction reactions involving electron transfer in solution is presented.
  41. [41]
    [PDF] Rudolph A. Marcus - Nobel Lecture
    Electron transfer theory. Treatment. In the initial paper (1956) I formulated the above picture of the mechanism of electron transfer and, to make the ...
  42. [42]
    On the Theory of Electron‐Transfer Reactions. VI. Unified Treatment ...
    A unified theory of homogeneous and electrochemical electron‐transfer rates is developed using statistical mechanics.
  43. [43]
    Inverted-region electron transfer as a mechanism for enhancing ...
    Aug 16, 2017 · Inverted-region electron transfer is widely suggested to be an important mechanism contributing to photosynthetic efficiency.Missing: original | Show results with:original
  44. [44]
    Golden rule kinetics of transfer reactions in condensed phase
    Dec 16, 2013 · A. Golden rule. The formulation of the Fermi golden rule that we invoke below for the rate computation is expressed in terms of distributions ...
  45. [45]
    Path‐integral treatment of multi‐mode vibronic coupling. II ...
    A path‐integral approach to real‐time quantum dynamics is presented which is suitable to treat the dynamics of vibronic coupling or spin boson models.
  46. [46]
    [PDF] Electron Transfer in Solution - DSpace@MIT
    We examine the dynamics of spin-boson models to fourth-order in time-dependent perturba- tion theory in the diabatic coupling. In order to prevent divergences, ...
  47. [47]
    Temperature dependent activation energy for electron transfer ...
    Jun 15, 1976 · This paper considers electron transfer between biological molecules in terms of a nonadiabatic multiphonon nonradiative decay process in a dense medium.Missing: vibrations | Show results with:vibrations
  48. [48]
    Role of Conical Intersections in Molecular Spectroscopy and ...
    May 5, 2012 · This review describes how conical intersections affect measured molecular spectra and simple photofragmentation processes.
  49. [49]
    Signatures of a Conical Intersection in Two-Dimensional Spectra of ...
    Aug 26, 2025 · A deeper understanding of nonradiative decay channels is essential, as rapid deactivation pathways can compete with electron injection into the ...
  50. [50]
    Electron transfer in extended systems: characterization by periodic ...
    We describe a new computer implementation of electron transfer (ET) theory in extended systems treated by periodic density functional theory (DFT), ...Missing: TD- nanomaterials
  51. [51]
    Photoinduced dynamics during electronic transfer from narrow to ...
    May 30, 2024 · Here, we show the occurrence of an electron transfer process in one-dimensional layer-stacking of carbon nanotubes (CNTs) and boron nitride nanotubes (BNNTs).
  52. [52]
    A combined experimental and computational study of electron ...
    Jun 1, 2025 · DFT/TD-DFT and experiment defined structure-property relationships governing self-assembly and electron transfer. Abstract. A novel ...<|control11|><|separator|>
  53. [53]
    Mitochondrial electron transport chain: Oxidative phosphorylation ...
    In this review, we will provide an overview of the function of the ETC, focusing on oxidative phosphorylation and its relationship to ROS production. Further, ...
  54. [54]
    Solar energy conversion by photosystem II: principles and structures
    Feb 24, 2023 · 1), the linear electron transfer chain of photosynthesis, known as Z-scheme (reviewed in Govindjee et al. 2017), bridges a large energy gap and ...
  55. [55]
    Molecular dynamics of photosynthetic electron flow in a ...
    Dec 15, 2024 · This system absorbs sunlight with a quantum efficiency of over 30% and splits water into oxygen, protons, and electrons. These protons and ...
  56. [56]
    selection and variance in electron tunnelling proteins including ...
    The effect of selection on distance is clear and dominant. The distance ranging from van der Waals contact to 14 Å, which accommodates 95% of distances ...
  57. [57]
    Plastocyanin: Structure and function | Photosynthesis Research
    Plastocyanin is a 10 kD blue copper protein which is located in the lumen of the thylakoid where it functions as a mobileelectron carrier shuttling electrons ...
  58. [58]
    Mitochondrial dysfunction in Parkinson's disease
    Nov 11, 2023 · Mitochondrial dysfunction is strongly implicated in the etiology of idiopathic and genetic Parkinson's disease (PD).
  59. [59]
    Temperature-dependent iron motion in extremophile rubredoxins
    May 28, 2024 · Rubredoxins are electron transfer proteins, and they play important roles in photosystem II assembly, as redox partners of alkane hydroxylases ...