Fact-checked by Grok 2 weeks ago

Photoinduced electron transfer

Photoinduced electron transfer (PET) is a light-triggered photochemical process in which an is transferred from a photoexcited donor to a ground-state acceptor, generating charge-separated states that can drive subsequent chemical reactions. This phenomenon occurs in donor-bridge-acceptor (D-B-A) architectures, either through covalent bonding or , and can proceed via oxidative (electron from excited donor to acceptor) or reductive (electron from ground-state donor to excited acceptor). The efficiency of PET is influenced by key factors including the potentials of the donor and acceptor, the distance between them, the surrounding medium, and the driving force of the reaction, as described by theories such as . In natural systems, PET is essential to , where it initiates the conversion of into ; for instance, in bacterial reaction centers like that of Rhodobacter sphaeroides, light absorption by a special pair of bacteriochlorophylls leads to rapid via pheophytin to a acceptor, achieving high . This process powers the , ultimately enabling ATP synthesis and carbon fixation in and . Beyond biology, PET underpins artificial photosynthetic systems designed to mimic natural processes for production, such as to generate or CO₂ reduction to form carbon-based fuels. Technological applications of PET are prominent in photovoltaic devices, particularly dye-sensitized solar cells (DSSCs), where photoexcitation of a sensitizer adsorbed on a TiO₂ leads to ultrafast injection into the conduction band, facilitating solar-to-electricity conversion with efficiencies exceeding 14% in recent designs. PET also plays a role in , for , and molecular , where controlled charge separation enables light-driven switching and information processing. Ongoing focuses on optimizing PET dynamics through nanostructured materials and hybrid systems to enhance energy conversion yields and stability.

Fundamentals

Definition and Scope

Photoinduced () is an that occurs from an electronic generated by the of . In this reaction, excites an in to , enabling the subsequent transfer of the and resulting in charge separation into radical ion pairs. This distinguishes PET from ground-state electron transfer, which proceeds without photoexcitation and typically involves less favorable energetics due to the absence of the excited-state driving force. The scope of PET spans diverse fields, including semiconductors, biological systems, and synthetic molecular architectures, where it facilitates efficient light-to-chemical or electrical energy conversion. In semiconductor applications, such as dye-sensitized solar cells, PET enables rapid electron injection from an excited dye sensitizer into the conduction band of a wide-bandgap oxide like TiO₂, initiating generation. In biological contexts, PET drives charge separation in light-harvesting complexes of photosynthetic organisms, where excited molecules transfer electrons to primary acceptors, powering the conversion of into chemical fuels. Similarly, in synthetic systems like organic , PET at donor-acceptor interfaces generates free charge carriers essential for device efficiency. Successful PET requires careful alignment of key parameters, including compatible redox potentials between donor and acceptor to ensure exergonic electron transfer and overlapping absorption spectra to match the incident light source for effective excitation. Unlike Förster resonance energy transfer (FRET), which transfers excitation energy through dipole-dipole coupling without net charge displacement, PET produces spatially separated charges that can drive subsequent redox chemistry or energy harvesting.

Historical Development

The earliest documented observations of light-induced processes, precursors to modern understanding of photoinduced (PET), date back to the . In 1805, Seekamp reported the photooxidation of under light, demonstrating how light could drive the decomposition of organic compounds in the presence of metal ions. Similarly, in 1830, Döbereiner observed the photooxidation of by ions in , providing further evidence of light-mediated in simple chemical systems. These findings, though not interpreted in terms of electron transfer at the time, laid the groundwork for recognizing photochemical reactions. In the mid-20th century, the concept of electron flow in biological systems gained prominence, particularly in research from 1949 to 1959. Eugene Rabinowitch's comprehensive works, including his 1956 treatise on photosynthetic transport, helped establish the framework for light-driven electron movement across membranes. This period also saw the discovery of non-cyclic by Daniel Arnon in 1954, which demonstrated ATP synthesis coupled to from to NADP+ under illumination, solidifying the role of sequential electron transfers in energy conversion. Rabinowitch's theoretical framework helped conceptualize the separation of two light reactions, highlighting PET as central to oxygenic . Theoretical advancements transformed PET from empirical observations to a predictive science. Rudolph A. Marcus's seminal 1956 paper introduced a quantum mechanical framework for rates, accounting for solvent reorganization and electronic coupling, which became foundational for understanding PET kinetics. This work, expanded in subsequent publications, earned Marcus the 1992 for contributions to electron transfer theory. In synthetic chemistry, Vincenzo Balzani and Franco Scandola's 1991 book Supramolecular Photochemistry formalized PET mechanisms in designed molecular assemblies, emphasizing vectorial electron transfer in multicomponent systems for mimicking natural processes. Post-2000 developments have integrated into , focusing on . Daniel G. Nocera's research advanced molecular catalysts, such as the cobalt-phosphate in 2008, enabling light-driven with high turnover numbers. His lab's hybrid systems, combining molecular catalysts with semiconductors or biological components, achieved efficient charge separation for production, with ongoing refinements in and reported through 2025.

Theoretical Framework

Basic Mechanism

Photoinduced electron transfer (PET) begins with the step, where a is absorbed by either the donor (D) or the acceptor (A) , promoting an from the (S₀) to an excited (S₁) or triplet (T₁) state. This process generates a highly reactive species, such as D* or A*, with sufficient to drive . For instance, the can be represented as D + hν → D*, where the excited donor possesses a higher , facilitating subsequent interactions. The initial transfer occurs in two primary modes: photooxidation, where the excited donor donates an to the ground-state acceptor (D* + A → D⁺ + A⁻), or photoreduction, where the excited acceptor accepts an from the ground-state donor (A* + D → A⁻ + D⁺). These processes create charge-separated states essential for further dynamics in systems. In metal es, specific examples illustrate these mechanisms; for photooxidation, excitation of a metal followed by electron donation to an acceptor yields [MLₙ]²⁺ + hν → [MLₙ]²⁺* → [MLₙ]³⁺ + acceptor⁻, as observed in ruthenium polypyridyl systems, such as [Ru(bpy)₃]²⁺* reducing methylviologen to form [Ru(bpy)₃]³⁺ + ⁺•. Similarly, for photoreduction, the excited accepts an electron from a donor: [MLₙ]³⁺ + hν → [MLₙ]³⁺* → [MLₙ]²⁺ + donor⁺, for example in cobalt(III) ammine es quenched by alcohols. Key factors influencing the initiation of PET include orbital overlap between the donor and acceptor, which enables efficient electron tunneling, and solvent effects that modulate the excited-state lifetimes, typically ranging from 10⁻⁹ to 10⁻⁶ s. Polar solvents can stabilize charge-separated intermediates but may also quench excited states through interactions, while nonpolar environments often extend lifetimes by reducing such quenching. These elements ensure the excited state persists long enough for transfer to compete with other deactivation pathways.

Rate Theories

The classical framework for quantifying the rates of photoinduced (PET) is , which models the process as a non-adiabatic transition between donor and acceptor states following photoexcitation. Developed by , this theory posits that the electron transfer rate depends on the overlap of vibrational wavefunctions in the initial and final states, accounting for nuclear reorganization in the surrounding medium. The rate constant k_{ET} for such transfers is expressed as: k_{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] where V represents the electronic coupling between donor and acceptor, \lambda is the reorganization energy associated with solvent and inner-sphere changes, \Delta G^\circ is the standard free energy change (driving force), k_B is Boltzmann's constant, T is temperature, and \hbar is the reduced Planck's constant. This semiclassical expression assumes a Gaussian distribution of nuclear configurations and has been foundational for predicting PET kinetics in solution and molecular systems. A of is the "inverted region," where for highly exergonic (large negative \Delta G^\circ, typically exceeding \lambda), the transfer rate decreases due to insufficient vibrational overlap caused by the need for the system to adjust to a higher-energy configuration. This counterintuitive behavior arises from the quadratic dependence in the activation free energy term, leading to slower rates despite greater thermodynamic favorability. The inverted region was theoretically anticipated in Marcus's original work but experimentally confirmed in PET systems through intramolecular transfers in rigid anions, where rates peaked at moderate driving forces and declined for more negative \Delta G^\circ. For non-adiabatic PET, particularly over distances greater than ~10 Å, quantum mechanical effects become prominent, with electron tunneling dominating via superexchange mechanisms. In superexchange, the donor and acceptor are weakly coupled through virtual intermediate states in bridging molecules or solvent, resulting in an effective electronic coupling V that decays exponentially with donor-acceptor separation r as V \propto \exp(-\beta r / 2), where \beta (typically 0.6–1.4 Å⁻¹) reflects the medium's electronic structure. Semiclassical extensions, such as those incorporating multiphonon transitions, address quantum tunneling in photoexcited states by treating low-frequency solvent modes classically and high-frequency intramolecular vibrations quantum mechanically, enhancing accuracy for PET in polar environments. Central parameters influencing PET efficiency include the driving force -\Delta G^\circ, which for optimal rates in the normal region falls in the range of 0.5–2 , balancing activation barriers with thermodynamic drive, and the distance dependence, where rates drop by an every 1–2 beyond close contact due to diminished [V](/page/V.). These factors underscore the theory's utility in designing systems for controlled charge separation, though solvent dynamics and medium can modulate \lambda and thus shift the optimal .

Key Processes

Primary Electron Transfer

Primary electron transfer constitutes the initial ultrafast injection of an electron from a photoexcited donor to an adjacent acceptor in molecular systems such as donor-acceptor dyads, typically occurring within femtoseconds to picoseconds. This process is central to photoinduced charge , where the of the donor, often a or similar , rapidly transfers the to an acceptor like a , achieving charge separation on the order of 4 picoseconds in rigid porphyrin-quinone dyad structures designed for . Such dynamics ensure competition with non-productive decay pathways, enabling efficient harvesting of photonic energy. The directionality of primary —forward from donor to acceptor versus backward recombination—depends critically on the reaction energetics, with forward transfer favored under exergonic conditions. The Rehm-Weller equation provides a key framework for estimating the standard change (ΔG°), given by ΔG° = E_{ox}(D) - E_{red}(A) - E_{00} + w_p, where E_{ox}(D) and E_{red}(A) are the oxidation potential of the donor and of the acceptor, respectively, E_{00} is the zero-zero of the donor, and w_p accounts for Coulombic work terms in solution. This relation, derived from fluorescence studies, predicts rates approaching the limit for moderately exergonic transfers (ΔG° ≈ -0.3 to -1 eV), guiding the design of dyads with appropriate tuning. Experimental characterization of primary electron transfer relies on time-resolved transient absorption spectroscopy, which detects the formation of radical ions through their distinct spectral signatures, such as the near-infrared absorption of the donor cation radical (D^{+•}) and the acceptor anion radical (A^{-•}). For instance, in perylene-based systems, this technique resolves the rise of radical ion bands within 100 femtoseconds, confirming the ultrafast nature of the transfer and distinguishing it from slower secondary processes. Medium effects, particularly solvent viscosity, influence the mechanism of primary electron transfer by modulating the balance between diffusion-controlled (dynamic) quenching, where molecules encounter during the excited-state lifetime, and static quenching, involving pre-formed donor-acceptor complexes. In viscous media, such as mixtures, dynamic contributions diminish as rates slow, allowing static quenching to dominate and revealing intrinsic transfer rates independent of collisional encounters.

Charge Separation and Recombination

In photoinduced electron transfer, charge separation refers to the spatial isolation of the oxidized donor (D⁺) and reduced acceptor (A⁻) following the initial event, which is essential to minimize rapid geminate recombination and enable productive . This isolation is often achieved through covalent bridging in dyad or molecules, where rigid spacers such as oligophenylenevinylene or norbornyl units enforce through-bond separation, reducing electronic coupling between D⁺ and A⁻. In supramolecular assemblies, solvent-separated pairs form via non-covalent interactions, such as axial coordination in porphyrin-fullerene complexes, further stabilizing the charges by effects in polar media like . Recombination of the separated charges primarily occurs through geminate pathways, where D⁺ and A⁻ return to the neutral . Radiative recombination manifests as fluorescence quenching of the donor or charge-transfer emission, though it is often suppressed in designed systems; non-radiative recombination dominates, dissipating energy as through vibrational relaxation involving phonons. In some cases, multiexciton recombination contributes under high-intensity , but of charge-separated states in optimized assemblies typically span from nanoseconds to microseconds, such as 310 μs observed in a zinc imidazoporphyrin-C60 dyad in benzonitrile at 278 . These extended arise from increased donor-acceptor distances and energetic barriers that slow back . Following charge separation, subsequent reactions can involve directed migration of holes or electrons along molecular chains, particularly in antenna-like architectures that mimic photosynthetic light-harvesting complexes. For instance, in carotenoid-polyene-porphyrin- pentads, sequential electron transfers propagate the positive charge across a polyene bridge while the electron resides on the fullerene acceptor, achieving separation over distances of about 4.4 and preventing recombination. Energy from recombination or non-productive paths is ultimately dissipated as phonons in the surrounding or . The efficiency of charge separation is quantified by the Φ_CS, which exceeds 0.9 in well-designed supramolecular systems, such as ruthenium-based assemblies for catalytic applications.

Applications

In Biological Systems

Photoinduced electron transfer (PET) is fundamental to energy conversion in biological systems, particularly in photosynthesis, where it facilitates the light-dependent reactions that sustain life on Earth. In oxygenic photosynthesis, occurring in chloroplasts of plants, algae, and cyanobacteria, light is absorbed by antenna complexes and funneled to the reaction centers of photosystems I (PSI) and II (PSII). In PSII, the primary donor, a chlorophyll a dimer termed P680, absorbs a photon to form the excited state P680*, which donates an electron to the nearby pheophytin (Pheo) acceptor in approximately 0.9 picoseconds, yielding the charge-separated pair P680⁺ Pheo⁻. This ultrafast step prevents recombination and initiates charge separation, with the electron subsequently transferred to the bound quinone QA, while P680⁺ is reduced by the oxygen-evolving complex via tyrosine Z. In PSI, a parallel process involves P700* transferring an electron to an iron-sulfur cluster A0, enabling further reduction downstream. The electrons from these PET events flow through the photosynthetic electron transport chain, depicted as the Z-scheme, linking PSII and PSI via the cytochrome b₆f complex. This linear pathway oxidizes water at PSII to produce O₂ and protons, while reducing NADP⁺ to NADPH at PSI via . The Q-cycle in cytochrome b₆f translocates protons across the thylakoid membrane, establishing an that drives ATP synthesis through . Cyclic electron flow around PSI supplements ATP production without NADPH formation. The light reactions achieve an energy conversion efficiency of approximately 30–40%, capturing a significant portion of absorbed visible to generate the ATP and NADPH needed for carbon fixation in the Calvin-Benson . Beyond oxygenic , PET operates in anoxygenic systems, such as bacterial reaction centers. In purple bacteria like Rhodobacter sphaeroides, light excites the special pair bacteriochlorophyll dimer (P*) in the reaction center, prompting to bacteriopheophytin (H_A) within 3–4 picoseconds to form P⁺H_A⁻, followed by transfer to the primary Q_A in hundreds of picoseconds. This process uses electron donors like or H₂S instead of , powering ATP synthesis via a proton gradient. The conservation of PET mechanisms across biological systems underscores their evolutionary significance. Oxygenic photosynthesis, with its PSII-driven water oxidation, emerged around 3 billion years ago in ancient , as evidenced by isotopic signatures in rocks and microfossils, fundamentally altering Earth's atmosphere by enabling the and paving the way for aerobic life.

In Technological Systems

Photoinduced electron transfer (PET) plays a pivotal role in engineered photovoltaic systems, particularly in dye-sensitized solar cells (DSSCs), also known as Grätzel cells, first demonstrated in 1991 using ruthenium(II) polypyridyl complexes as sensitizers adsorbed on nanocrystalline TiO₂ films. In these devices, visible light excites the sensitizer, leading to rapid electron injection into the TiO₂ conduction band, followed by regeneration of the sensitizer by a mediator, achieving power conversion efficiencies up to 13.7% in co-sensitized configurations as of August 2025. This PET-driven process enables low-cost fabrication and operation under diffuse light, distinguishing DSSCs from traditional silicon-based . In , PET facilitates visible-light by generating reactive intermediates for selective C-H activation and bond cleavage, often employing or complexes as photoredox catalysts. For instance, [Ru(bpy)₃]²⁺ derivatives promote single-electron transfer to substrates, enabling transformations like α-amino generation for bond formation under mild conditions, as reviewed in studies on photoredox mechanisms. Similarly, catalysts, such as Ir(ppy)₃, drive oxidative C-H functionalization via PET quenching, offering sustainable alternatives to thermal methods with high tolerance. PET-based fluorescent probes serve as sensitive tools in sensors and , where is quenched by intramolecular from a donor moiety until to an analyte disrupts the process. These probes, often featuring fluorophores like fluorescein conjugated to receptor units, exhibit "turn-on" emission upon coordination with metal ions such as Zn²⁺ or Cd²⁺ in aqueous media, enabling detection limits in the nanomolar range for environmental and biological monitoring. For example, Schiff base derivatives leverage PET inhibition for selective ratiometric sensing of , providing real-time capabilities in cellular contexts. Emerging technological applications harness in and solar cells, where photoexcitation initiates efficient charge separation at interfaces, contributing to certified efficiencies exceeding 27.2% for single-junction as of November 2025. In these systems, like MAPbI₃ facilitate ultrafast to electron-transport layers such as PCBM, minimizing recombination losses and enhancing stability through interface engineering. variants, incorporating colloidal nanocrystals, achieve up to 18.3% efficiency via size-tunable , with potential for tandem architectures pushing overall performance beyond 25%.

Reverse Processes

Reverse electron transfer (RET) refers to the back transfer of an electron from the reduced acceptor (A⁻) to the oxidized donor (D⁺) in a photoinduced (PET) system, reforming the ground-state donor-acceptor pair (D A). This process often competes with charge separation, leading to of the and reduced efficiency in PET-driven systems. In molecular dyads and triads, RET is a primary decay pathway for the charge-separated state, particularly when the donor and acceptor are in close proximity, limiting the lifetime of useful charge separation. The dynamics of RET are governed by , analogous to forward , but RET frequently operates in the inverted regime when the process is highly exergonic (where the driving force |–ΔG| exceeds the reorganization energy λ). In this regime, increasing the exergonicity paradoxically slows the RET rate, which is advantageous for suppressing unproductive recombination. For endergonic RET, rates follow the normal Marcus regime, but experimental observations in donor-acceptor systems confirm the inverted behavior for exergonic back transfer, with rates tuned by reorganization and electronic coupling. In (ECL), RET is coupled with electrochemical oxidation and reduction steps to generate s without initial photon absorption. A classic example is the Ru(bpy)₃²⁺/, where electrochemical generation of Ru(bpy)₃³⁺ and a reduced coreactant (e.g., TPrA•) leads to RET: Ru(bpy)₃³⁺ + TPrA• → *Ru(bpy)₃²⁺ + TPrA⁺, producing the luminescent *Ru(bpy)₃²⁺. This pathway, or coreactant-mediated variant, relies on efficient RET to achieve high ECL intensity, with by reverse processes influencing signal output. In biological systems like photosynthetic reaction centers, RET from the primary acceptor (Q_A⁻) to the oxidized donor (P⁺) can limit by recombining charges before secondary transfer occurs. In , RET in the P700⁺A₁⁻ state operates in the Marcus inverted region, slowing recombination rates to near 10⁻² s⁻¹ at 298 K and enhancing forward transfer yields to over 98%. Suppression of RET by increasing donor-acceptor distance, as in bacterial reaction centers, further boosts by reducing electronic coupling and favoring charge stabilization.

Induced Photon Production

In photoinduced electron transfer (PET) processes, induced photon production arises when subsequent recombination or energy transfer events populate emissive excited states, leading to emission distinct from initial photoexcitation. This phenomenon extends PET dynamics to scenarios where external potentials, chemical reactions, or intermediate states trigger radiative decay, often mimicking or enhancing the outcomes of direct charge separation. Such mechanisms are pivotal in converting electronic excitations into detectable photons, with applications spanning and bioimaging. Potential-induced emission occurs in semiconductors through radiative electron-hole recombination, where an applied voltage generates charge carriers analogous to photoexcitation in , enabling release upon their . In light-emitting diodes (LEDs), electrons injected from the n-type region and holes from the p-type region recombine in the active layer, releasing energy as with wavelengths determined by the bandgap energy. This process, termed radiative recombination, achieves efficiencies up to 90% in III-V semiconductors like GaAs, where the recombination rate is proportional to carrier concentration and radiative lifetime. Unlike direct , the voltage-driven carrier injection bypasses light absorption but parallels the charge generation step, with non-radiative pathways minimized through material engineering. Chemiluminescence via electron transfer manifests in exergonic reactions like peroxyoxalate systems, where populates high-energy intermediates that decay radiatively, often releasing CO₂ and . In these systems, an activator (e.g., 9,10-diphenylanthracene) undergoes electron donation to a peroxoester intermediate formed from and , followed by back that excites the activator to its for . This chemically initiated electron-exchange luminescence (CIEEL) mechanism yields quantum efficiencies of 20-30%, far surpassing many thermal processes, and is catalyzed by bases like to accelerate intermediate formation. The process exemplifies how intermediates can drive production without external light, with seminal studies establishing the role of solvent-independent in chemiexcitation, analogous to aspects of . Upconversion through (TTA) leverages PET intermediates to enable anti-Stokes emission, converting low-energy photons to higher-energy ones via sequential energy transfers. In donor-acceptor sensitizers like BODIPY-pyrene dyads, photoexcitation induces PET to form a charge-transfer state, which undergoes spin-orbit charge-transfer (SOCT-ISC) to populate triplets (e.g., at 1.69 ), subsequently transferring energy to acceptors like for TTA. Two triplets annihilate to yield one singlet emitter, achieving upconversion quantum yields up to 6.9% in deoxygenated solvents, with PET tuning via substituents controlling efficiency. This PET-mediated pathway avoids heavy-atom effects, enhancing applicability in and sensing. Applications of induced photon production include organic LEDs (OLEDs) incorporating quenching layers to optimize by suppressing non-radiative decay, and bioluminescent probes exploiting in luciferin oxidation for targeted imaging. In OLEDs, from electron-rich layers quenches triplet states in adjacent emitters, reducing efficiency roll-off and enabling color-pure output, as seen in phosphorescent devices with donor-acceptor interfaces achieving external quantum efficiencies over 25%. For bioluminescent probes, oxidation involves single from the phenolate to O₂, forming a dioxetanone intermediate that decomposes to an excited oxyluciferin emitter (peak at 560 nm), with quantum yields of ~0.41; engineered BioLeT probes modulate this via enzyme-induced for detection with limits down to 1 μM. Recent developments as of 2024 include bioluminescent probes for selective detection of and , enhancing sensitivity in living systems. These systems highlight the role of in scalable, biocompatible sources.

References

  1. [1]
    Photoinduced Electron Transfer in Organized Assemblies—Case ...
    In this review, photoinduced electron transfer processes in specifically designed assembled architectures have been discussed in the light of recent results.
  2. [2]
    PHOTOINDUCED ELECTRON TRANSFER - Fox - 1990
    In both systems, the critical event of photosynthesis is the transfer of an electron from the photoexcited special pair of bacteriochlorophylls located at ...
  3. [3]
    Review Energy Conversion in Natural and Artificial Photosynthesis
    May 28, 2010 · The transfer of electrons across the photosynthetic membranes drives chemical reactions and can also produce electricity. Similarly, electricity ...
  4. [4]
    Dye-sensitized solar cells strike back - RSC Publishing
    Sep 30, 2021 · Photoinduced electron transfer from a dye molecule to the conduction band of TiO2 is the first step in the working mechanism of a dye ...
  5. [5]
    Photoinduced electron transfer in supramolecular systems for ...
    Photoinduced electron transfer in ... Molecular-Modified Photocathodes for Applications in Artificial Photosynthesis and Solar-to-Fuel Technologies.
  6. [6]
    photoinduced electron transfer (P04617) - IUPAC
    An electron transfer resulting from an electronic state produced by the resonant interaction of electromagnetic radiation with matter. Source:
  7. [7]
    Photoinduced Electron Transfer in Inclusion Complexes of Carbon ...
    Dec 16, 2023 · Photoinduced electron transfer (PET) in complexes is an excited state process that proceeds from the electron donor (D) to the electron acceptor ...
  8. [8]
    Fundamentals of Photoinduced Electron Trmsfer - ResearchGate
    Aug 7, 2025 · Photoinduced electron transfer (PET) is a term reserved to describe the transfer of an electron between photoexcited and ground-state molecules.
  9. [9]
    Charge-transfer states in photosynthesis and organic solar cells
    This contribution aims to provide an introductory review and comparison of the light-induced charge-transfer mechanisms occurring in natural photosynthesis and ...Missing: paper | Show results with:paper
  10. [10]
    Photoinduced Electron Transfer in Organic Solar Cells
    Feb 8, 2016 · Electron transfer (ET) is the key process in light-driven charge separation reactions in organic solar cells. The current review summarizes ...Abstract · Introduction · Models for Electron Transfer · Three Key Parameters in...
  11. [11]
    Recent Advances in Photoinduced Electron Transfer Processes of ...
    From the redox potentials, the free energy change for photoinduced intermolecular electron transfer for the mixture system (−ΔGPET) can be calculated to be 0.3 ...
  12. [12]
    Competition between Photoinduced Electron Transfer and ...
    Mar 24, 2021 · Our findings show that there is a net preference for photoinduced ET over other ways of energy transfer, at least partially, due to a lack of donors.Introduction · Experimental Section · Results and Discussion · Conclusions
  13. [13]
    Separation of Two Light Reactions in Noncyclic Photo ... - Nature
    Separation of Two Light Reactions in Noncyclic Photo-Phosphorylation of Green Plants ... Govindjee, R., Thomas, J. B., and Rabinowitch, E., Science, 132, 421 ( ...
  14. [14]
    On the Theory of Oxidation‐Reduction Reactions Involving Electron ...
    R. A. Marcus; On the Theory of Oxidation‐Reduction Reactions Involving Electron Transfer. I. J. Chem. Phys. 1 May 1956; 24 (5): 966–978. https://doi.org ...
  15. [15]
    Supramolecular Photochemistry - SpringerLink
    Towards a Supramolecular Photochemistry: Assembly of Molecular Components to Obtain Photochemical Molecular Devices. V. Balzani, L. Moggi, F. Scandola. Pages ...
  16. [16]
    All publications - Nocera Lab
    2025 PUBLICATIONS. Lanthanum-Promoted Electrocatalyst for the Oxygen Evolution Reaction: Unique Catalyst or Oxide Deconstruction?
  17. [17]
    Photoinduced electron transfer reactions of metal complexes in ...
    Photochemical CO2 Reduction Catalyzed by Phenanthroline Extended Tetramesityl Porphyrin Complexes Linked with a Rhenium(I) Tricarbonyl Unit.Missing: equations | Show results with:equations
  18. [18]
    Solvent effects on intermolecular photoinduced electron transfer ...
    The electron transfer nature of the quenching reaction in all the solvents was confirmed by the transient absorption spectra of the radical ions formed.
  19. [19]
    [PDF] Rudolph A. Marcus - Nobel Lecture
    Treatment. In the initial paper (1956) I formulated the above picture of the mechanism of electron transfer and, to make the calculation of the reaction rate.
  20. [20]
    Ultrafast Photoinduced Electron Transfer in Rigid Porphyrin ...
    Jul 1, 1995 · Probing the Donor−Acceptor Proximity on the Physicochemical Properties of Porphyrin−Fullerene Dyads: “Tail-On” and “Tail-Off” Binding ...Missing: primary | Show results with:primary
  21. [21]
    The Rehm–Weller Experiment Revisited with Femtosecond Time ...
    Jan 8, 2014 · A kinetic anal. yields the expression kq = 2.0 × 1010/1 + 0.25 [exp(ΔG23/RT) + exp(ΔG⧧23/-RT)]M-1 sec.-1 where ΔG⧧23, the free enthalpy of ...
  22. [22]
    Spurious Observation of the Marcus Inverted Region in Bimolecular ...
    Jul 2, 2012 · The effect of viscosity on the bimolecular electron transfer quenching of a series of coumarins by N,N-dimethylaniline was investigated
  23. [23]
    Photoinduced electron transfer across molecular bridges
    Mar 6, 2014 · Superexchange is an effective model to rationalize photoinduced electron transfer, particularly when molecular bridges between donor and acceptor subunits are ...Missing: seminal paper
  24. [24]
    Long-Lived Charge-Separated State Produced by Photoinduced ...
    The charge-separated state has a lifetime of 310 μs in benzonitrile at 278 K, which is the longest lifetime in solution ever reported for electron donor− ...
  25. [25]
    Photoinduced electron transfer and geminate recombination in ...
    Feb 23, 2006 · Once the charge is transferred to the acceptor, there are two competing processes: geminate recombination and radical escape (the radical ...
  26. [26]
    Photoinduced double charge accumulation in a molecular compound
    Aug 25, 2025 · Fuel-forming reactions require multiple electrons, whereas photoinduced elementary reaction steps are usually single electron transfer events.
  27. [27]
    Photoinduced electron and energy transfer in molecular pentads
    ... Antenna Effects and Charge Separation. 1996, 1-95. https://doi.org ... GUST, T. A. MOORE, A. L. MOORE, A. N. MACPHERSON, A. LOPEZ, J. M. DEGRAZIANO ...
  28. [28]
    Ultrafast Photoinduced Electron Transfer in a Photosensitizer Protein
    Jun 12, 2021 · The PET driving force from the Asp and Tyr to BpC* was estimated to be +0.08 and −0.25 eV, respectively. To compare the PET driving force ...<|control11|><|separator|>
  29. [29]
    None
    ### Summary of Z-Scheme and Related Processes in Photosynthesis
  30. [30]
    Primary electron transfer in Rhodobacter sphaeroides R-26 reaction ...
    The bacterial photosynthetic reaction center (RC) is a specialized protein-cofactor complex that uses light energy to trigger electron-transfer reactions ...
  31. [31]
    Photoinduced dynamics during electronic transfer from narrow to ...
    May 30, 2024 · Another famous case is the electron transfer between the retinal chromophore in rhodopsin, which is fundamental for eye vision. In the ...
  32. [32]
    Applications of bioluminescence in biotechnology and beyond
    Mar 18, 2021 · Bioluminescent Enzyme-Induced Electron Transfer (BioLeT) is analogous to photoinduced electron transfer (PeT) which has often been ...
  33. [33]
    Early Evolution of Photosynthesis - PMC - NIH
    Oct 6, 2010 · There is suggestive evidence that photosynthetic organisms were present approximately 3.2 to 3.5 billion years ago, in the form of ...
  34. [34]
    A low-cost, high-efficiency solar cell based on dye ... - Nature
    Oct 24, 1991 · A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Brian O'Regan &; Michael Grätzel. Nature volume 353, ...Missing: original | Show results with:original
  35. [35]
    Co-Sensitized Solar Cell Achieves 13.7% Efficiency with Bis ...
    Co-Sensitized Solar Cell Achieves 13.7% Efficiency with Bis-Hexylthiophene Dyes ... Adv Sci (Weinh). 2025 Aug 18:e09116. doi: 10.1002/advs.202509116.
  36. [36]
    Recent advances in visible light-activated radical coupling reactions ...
    Jul 26, 2021 · It has been demonstrated that ruthenium and iridium complexes make effective photoredox and energy transfer catalysts. The inclusion of heavy ...
  37. [37]
    Photoredox catalysis by [Ru(bpy)3]2+ to trigger transformations of ...
    Aug 6, 2025 · Photoinduced redox processes using visible light in conjunction with sensitizing dyes offer a great variety of catalytic transformations useful in the realm of ...<|separator|>
  38. [38]
    Fluorescent Schiff base sensors as a versatile tool for metal ion ...
    Feb 17, 2022 · The coordination of metal ions with the Schiff base sensor may favor or disrupt the PET process, resulting in emission change. Chakraborty et al ...
  39. [39]
    An efficient PET-based probe for detection and discrimination of Zn ...
    From a pool of different metal ions in aqueous medium, Zn2+ and Cd2+ ions are simultaneously detected by the probe, giving an excellent enhancement in the ...
  40. [40]
    Two Novel Fluorescent Probes as Systematic Sensors for Multiple ...
    Sep 14, 2020 · Many precedents prove that fluorescent probes are promising candidates for detection of metal ions in the environment and biological systems ...
  41. [41]
    Advancements in perovskites for solar cell commercialization: A review
    The efficiency of perovskite solar cells (PSCs) has progressed rapidly, exceeding 26% for single-junction devices and surpassing 34% in perovskite-silicon ...<|separator|>
  42. [42]
    Measuring the Efficiency of Photoinduced Electron Transfer at the ...
    Apr 10, 2025 · The incorporation of perovskite quantum dots within metal–organic frameworks can be a viable approach to design efficient semiconductor ...
  43. [43]
    Perovskite quantum dot solar cell achieves record ... - PV Magazine
    Oct 8, 2025 · Perovskite quantum dot solar cell achieves record-breaking efficiency of 18.3%. Conceived by scientists in China, the cell was built with an ...Missing: 2020s | Show results with:2020s
  44. [44]
    Electron Transfer Quenching and Electrochemiluminescence ...
    The scheme includes also the reverse electron transfer corresponding to the back electron transfer to the excited state (with the rate k32) and the singlet− ...<|separator|>
  45. [45]
    Photoinduced electron transfer in a molecular dyad by nanosecond ...
    This study is a step toward the improvement of synthetic strategies of molecular photocatalysts for light-induced charge accumulation and more generally, for ...
  46. [46]
    Inverted-region electron transfer as a mechanism for enhancing ...
    Aug 16, 2017 · Inverted-region electron transfer is therefore demonstrated to be an important mechanism contributing to efficient solar energy conversion in photosystem I.
  47. [47]
    Temperature Dependence of Forward and Reverse Electron ...
    Control of electron transfer by protein dynamics in photosynthetic reaction centers. Critical Reviews in Biochemistry and Molecular Biology 2020, 55 (5) ...