Abstract

Understanding the kinetics of electron transfer reactions involves active research in physics, chemistry, biology, and nano-tech. Here, we propose a model to apply in a broader framework by establishing a connection between thermodynamics and kinetics. From a purely thermodynamic point of view, electronic transfer Marcus’ theory is revisited; consequently, calculations of thermodynamic variables such as mobility, energy, and entropy are provided. More significantly, two different regimes are explicitly established. In the anomalous region, an exergonic process associated with negative heat capacity appears. Further, in the same region, mobility, energy, and entropy decrease when the temperature increases.
1. Introduction
Electron transfer (ET) occurs in many phenomena in nature, such as the processes of photosynthesis, light harvesting in solar cells, and oxidation–reduction reactions in biology and chemistry.1−4 Several approaches have been proposed to understand the kinetics of ET-based reactions, and one of the most important is the theory developed by Marcus.1−14
Generally, Marcus’ phenomenological theory considers ET between electron-donor and electron-acceptor species in a finite system.1−3 That is, an open system with two sub-systems exchanging charges, while the surrounding environment is reorganized (solvent sheath of ligands and/or adjacent free solvent), and their individualities are retained. Although ET is basically a quantum phenomenon, the environment influences the reaction because it modulates the electron jump. This formulation has been improved and extended gradually to describe a wide variety of reactions, such as reactions between ions or molecules and electrodes, photoelectric emission spectra of ionic solutions, chemiluminescent ETs, ET through frozen media, and ET through thin hydrocarbon-like films on electrodes, among others.15−19 In fact, it is a transport theory for mesoscopic systems.20 Marcus’ theory allows us to evaluate the rate kET of reactions, which involves both static and dynamic effects.5−8 Static effects such as stabilization of reactants, products, and dynamic effects correspond to relaxations of nuclear and solvent modes.
The kET rate depends on quantities that can be computed or determined experimentally (free energy, reorganizational energy, and electronic coupling). The electronic coupling process becomes a relevant variable because it describes the adiabatic or non-adiabatic nature of the reactions. The non-adiabatic regime corresponds to a small coupling, while large electronic coupling accounts for adiabatic transfer. In this work, from a purely thermodynamic point of view, electronic transfer Marcus’ theory is revisited, consequently, calculations of thermodynamic variables such as mobility, energy, entropy, and others are provided. More importantly, two different regimes are explicitly established (normal and anomalous). Novelties such as negative heat capacity or mobility decreasing with temperature, among others, also appear. The paper is structured as follows. Marcus’ model is revisited in Section 2. The mobility for the normal and the anomalous region is described in Section 3. In addition, the corresponding threshold separating both regions is explicitly obtained. Provided that for an open quantum system, the lifetime-width is related to the energy through the uncertainty principle, the internal energy (Section 4), the partition function, the free energy, heat capacity, and entropy are calculated in Section 5. These magnitudes likewise display a threshold and split the thermodynamic behavior into a normal and anomalous regime. Finally, Section 6 presents conclusions and scopes.
2. Marcus’ Theoretical Model for ET
This theory can be derived through semiclassical expressions.21 In this context, the relaxation time τ is given by1−3
| 1 |
where the apparition of Planck’s constant ℏ and temperature T tells us that the electronic motion is linked to quantum decoherence. The Gibbs free energy G is the driving force for the reaction, and the reorganizational energy λ corresponds to the required rearrangement of the reactants and solvents. The reorganization energy is composed of two terms: internal reorganization energy, λi, which accounts for the change in the bond angles and bond lengths, and external reorganization energy, λs, which considers the energy of the solvent shell for the required rearrangement.13 Finally, the superposition between the donor and acceptor wave functions is described by the electronic energy coupling H or degree of quantum mechanical mixing.
Regarding eq 1, the reaction rate, kET = 1/τ, is maximum when the free energy of activation EA ∝ (G + λ)2 is minimum (G = −λ). Figure 1a shows the dependence for kET as a function of free energy G. The magnitudes of G and λ determine three situations for the transfer process. The lower panels represent the Gibbs free energy curves of the donor GD and the acceptor GA corresponding to the three expected situations: (b) the normal regime, when −G < λ, in which kET increases when the energy −G goes up. (c) The “activationless” region (maximum kET) corresponding to the top of the curve of panel (a) satisfies the condition (G = −λ). Lastly, (d), an interesting anomalous regime occurs when −G > λ, in which the reaction becomes slower as −G goes up. This regime, called the Marcus inverted region,13 has been experimentally verified for long-distance intermolecular ET reactions.22−24 The reaction coordinate (x axis) takes into account the energy difference between the donor and acceptor structure. The greater the difference in structure, the greater the displacement of the energy curves. Furthermore, as the energy barrier also increases with the displacement of the energy curves, the slower the reaction.
Figure 1.

(a) Dependence between the logarithm of the reaction rate kET as a function of Gibbs free energy Geq 1. The lower panels show the Gibbs free energy curves at (b) the normal region −G < λ, (c) activationless region (G = λ), and (d) anomalous region −G > λ with spontaneous energy emission (no barrier between both minimum).
Finally, estimations based on eq 1 commonly underestimate the rate transfer in the inverted region. The previously proposed model, as Marcus–Jortner–Levich theory, overcame some of the limitations of the original Marcus’ model by separating the internal and external reorganization energy and by explicitly including the contribution from the vibronic coupling between the reactant and product state.25
3. Mobility: Normal and Anomalous Behavior
Mobility μ is a key factor that characterizes charge transport. Several models have been proposed to describe the carrier drift in materials.26−29 In the hopping model, which dominates at room temperature, mobility can be connected with the diffusion constant through Einstein’s relation.30 In the standard band theory model, which is suitable for ordered systems at low temperatures, the transport behavior is governed by the Boltzmann equation which can be solved using relaxation-time approximation.31 In this case, the mobility derived according to first principles32−42 is μ = qτ/m*(H), where q is the electrical charge, τ is the average relaxation time, and m*(H) is the effective mass which depends proportionally on the hopping transport.29,43 Besides, the electronic mobility can be related to the conductivity through the Landauer equation44,45 by using the Einstein relation between conductivity and diffusion.46,47 Therefore, from eq 1, the electronic mobility becomes
| 2 |
Figure 2 shows the
dimensionless mobility μ/μ0 [where
] as a function of the dimensionless temperature t = 4λkBT/(G + λ)2. The more relevant topic
is the apparition of two regimes depending on temperature. These two
behaviors also exist for other quantities such as energy, specific
heat, and additional thermodynamic functions (Sections 4 and 5). Figure 2 inset shows an amplification
around the threshold point tc = 2.
Figure 2.

Dimensionless
mobility μ/μ0 as a function
of the dimensionless temperature t = 4λkBT/(G + λ)2 and the mobility parameter
). Two regions exist depending on the temperature.
In the first one, the mobility drops when the temperature increases
(normal regime). In the second, the mobility rises when the temperature
grows (anomalous). The inset shows an amplification around the threshold
red point tc = 2.
In this way, there is a singular point tc separating two kinds of thermodynamic behaviors:
For temperatures t < tc the mobility decreases with the temperature (normal behavior).
When t > tc the mobility rises with the temperature (anomalous). Also expected for electronic conductivity.
As mentioned, two regimes will likewise be deducted for other thermodynamic quantities. Chiefly, it will be also true for the specific heat with two branches (positive and negative48−61). The negative branch is related to exergonic processes as it will be discussed in detail in the next section.
Finally, the charge transport properties of conjugated small molecules/polymers are relevant parameters to evaluate their performance in organic optoelectronic applications.62,63 Indeed, the molecular parameters H and λ become of paramount importance for high charge transport ability. Several studies have shown that the combination of density functional theory (DFT), with Marcus’ theory and the Einstein–Smoluchowski equation, allows the prediction of the mobility of positive and negative charge carriers in conjugated systems.64−67 Although comparable results have been obtained between predicted mobilities and experimental measurements, the computational cost of DFT calculations increases significantly as systems become more complex. In this sense, eq 2 allows estimating the charge mobility more simply. As an example, we have estimated the hole mobility μ for two typical π-electronic core organic semiconductors, the pentacene (C22H14)68,69 and the rubrene (C42H28).70−72 These semiconductors present a negative temperature coefficient of the mobility (dμ/dT < 0) under ambient conditions, which can be considered as a fingerprint of the “band-like” transport behavior.73 The room-temperature hole mobilities obtained by eq 2 gives values of order of units cm2/(V s) (see Supporting Information, Figure S2), which are in good agreement with the reported data for pentacene69,74−76 and rubrene.70,72,74 Let us remark that the effective mass is not a quantity that can be easily experimentally determined which makes the estimation of μ0 complicated. However, the effective mass can be obtained by the analytical description m* = −ℏ2/(∂2ϵ(k)/∂k2), where ϵ(k) are the eigenvalues, and k is the wavevector.77−80
4. Internal Energy and Negative Specific Heat
It is well known that for quantum systems, the lifetime-width, out of the stationary regime, and the representative spread spectrum E (or level width) can be connected through81−83
| 3 |
which can be seen as a fluctuation–dissipation relation.32−36 Consequently, eqs 1 and 3 give the energy estimation as a function of temperature
| 4 |
with a two-regime behavior as a function of temperature, as shown in Figure 3. Namely, it is analogous to the mentioned case of mobility. Importantly, from eq 4, we can also obtain the variation of the energy at a constant temperature
| 5 |
Figure 3.

Dimensionless energy
as a function of normalized temperature t = 4λkBT/(G + λ)2 [i.e.,
]. There are two behaviors, or regimes,
as occurs with mobility (Figure 2). In the anomalous region t >
2,
the specific heat c = ∂E∂T becomes negative and compatible with spontaneous energy
loss (exergonic process).
hence, in the anomalous regime (G + λ) < 0 (Figure 1) and for spontaneous process ΔG < 0 there is energy emission ΔE < 0 (exergonic process).
Specific heat c = ∂E/∂T32−35 presents negative values48−61 in the anomalous regime (t > 2, Figure 3). Usually, a negative specific heat is associated with non-isolated subsystems composing a system.33,50 Always at the anomalous regime, a diminution of the dimensionless energy Δe < 0 is encompassed by an augmentation of Δt (Figure 3), but as long as t = 4λkBT/(G + λ)2 a variation Δt ∝ −ΔG ensues. Accordingly, if Δt < 0 then ΔG < 0, also corresponding to the mentioned exergonic behavior in the region of negative specific heat (i.e., the system loses spontaneously energy, Δe < 0).
5. Partition Function, Free Energy, and Entropy
At this stage, it is convenient to use so-called activation energy EA(1−3) related to energy E by
| 6 |
which measures transport difficulties through a potential barrier. As long as the activation energy EA and the partition function Z are connected through EA = ∂/∂β ln(Z) [with β = 1/(kBT)],32−35 it follows
| 7 |
which, from eq 6, gives the expression for the energy E described by (eq 4). Helmholtz free energy F = −ln(Z)/β becomes related to the entropy S through the usual expression F = EA – TS, but in terms of the energy E it must be modified as
| 8 |
In this way, the entropy S = ∂F/∂T becomes evaluated as
| 9 |
which as Figure 4 shows also presents a two-regime behavior with a singular point at tc = 2.
Figure 4.

Dimensionless entropy S = 2S0((4 + 2t + t2)/t2)e–2/t as a function of dimensionless temperature t = 4λkBT/(G + λ)2, where S0 = πH4kB/(G + λ)4 is an auxiliary parameter. There are two behaviors, or regimes, as occurs with mobility and energy (Figures 2 and 3). In the anomalous regime, encompassed by an exergonic behavior, the system tends to weak order. The inner figure shows the entropy function around tc = 2 corresponding to a maximum.
The importance of the entropy function is related to the second principle of thermodynamics defining irreversibility as it agrees with a dissipative system. The core of the ET mechanism is associated with dissipation and consequently related to conductance or mobility (Section 3). On the other hand, the entropy function can be associated with the degree of the disorder32−35 and, in the anomalous regime (t < 2), it diminishes when the dimensionless temperature rises. Consequently, a (weak) degree of order in this singular regime can be expected.
Figure 4 exhibits the entropy as a function of dimensionless temperature t. In the normal regime (t < 2), the entropy grows with temperature. Conversely, in the anomalous regime (t > 2), the entropy slightly decreases when the temperature rises. Finally, and formally, when t → 0 the entropy goes to zero (Nernst principle32) and at the limit of large temperature t → ∞ the entropy goes to a constant.
6. Conclusions
Open incoherent/dissipative electronic transport between two sub-systems (donor/acceptor) has been widely discussed through Marcus’ theory of ET. This scheme gives the relaxation rate of electronic motion as a function of temperature and appropriate energy parameters (eq 1). It becomes connected to mobility (eq 1, Figure 2) and supports two regimes, normal and anomalous. In the normal regime, mobility decreases with the temperature and likewise the electrical conductivity and, probably, the thermal conduction. Conversely, in the anomalous regime, mobility rises with the temperature. Additionally, the threshold between both regimes was explicitly evaluated. Although comparable results have been obtained between predicted mobilities and experimental measurements, the computational cost of DFT calculations increases significantly as systems become more complex. In this sense, our outcomes allow estimating the charge transport properties in a simpler and less computationally expensive way. From the usual connection between energy and relaxation time eq 3, the thermal energy of the system emerges. It retains clearly, similar to mobility, a two-regime behavior (Figure 3). Even more interesting, in the anomalous regime, the specific heat becomes negative, and spontaneous energy loss exists (exergonic process). Free energy and entropy also were evaluated showing the same two behaviors. For entropy, Figure 4, the temperature augmentation is encompassed by entropy diminution in the anomalous case. Moreover, always in this region, the system becomes tiny ordered. Entropy, energy, and mobility have (dimensionless) temperature thresholds. Finally, in future works, we expect to connect our results with experiments on electronic transport in more complex organic molecules.
Acknowledgments
A.M.C. and J.C.F.A. thank the support from UTA-Project 4722-22. A.M.C. thanks the support from ANID Project SA77210039. I.A.J. thanks the support from UTA-Project 4769-22. F.H. thanks the support from UTA-Project 4721-21.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.2c04094.
Brief connection with experimental and theoretical data and estimation of hole transfer rates kET and hole mobility μ as a function of the temperature for pentacene (C22H14) and rubrene (C42H28) (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Marcus R. A. Electron transfer reactions in chemistry. Theory and experiment. Rev. Mod. Phys. 1993, 65, 599–610. 10.1103/revmodphys.65.599. [DOI] [Google Scholar]
- Blumberger J. Recent Advances in the Theory and Molecular Simulation of Biological Electron Transfer Reactions. Chem. Rev. 2015, 115, 11191–11238. 10.1021/acs.chemrev.5b00298. [DOI] [PubMed] [Google Scholar]
- Williams R. M.; Koeberg M.; Lawson J. M.; An Y. Z.; Rubin Y.; Paddon-Row M. N.; Verhoeven J. W. Photoinduced electron transfer to C60 across extended 3- and 11-bond hydrocarbon bridges: Creation of a long-lived charge-separated state. J. Org. Chem. 1996, 61, 5055–5062. 10.1021/jo960678q. [DOI] [Google Scholar]
- Kuznetsov A. M.; Ulstrup J.. Electron Transfer in Chemistry and Biology: An Introduction to the Theory; John Wiley & Sons Ltd., 1999. [Google Scholar]
- Jortner J. Temperature dependent activation energy for electron transfer between biological molecules. Chem. Phys. 1976, 64, 4860–4867. 10.1063/1.432142. [DOI] [Google Scholar]
- Gray H. B. Long-range electron transfer. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 3534–3539. 10.1073/pnas.0408029102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barthel E. R.; Martini I. B.; Schwartz B. J. How Does the Solvent Control Electron Transfer? Experimental and Theoretical Studies of the Simplest Charge Transfer Reaction. J. Phys. Chem. B 2001, 105, 12230–12241. 10.1021/jp011150e. [DOI] [Google Scholar]
- Skourtis S. S.; Waldeck D. H.; Beratan D. N. Fluctuations in Biological and Bioinspired Electron-Transfer Reactions. J. Electroanal. Chem. 2010, 61, 461–485. 10.1146/annurev.physchem.012809.103436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marcus R. A. On the Theory of Oxidation-Reduction Reactions Involving Electron Transfer. I. Chem. Phys. 1956, 24, 966–978. 10.1063/1.1742723. [DOI] [Google Scholar]
- Marcus R. A. Tutorial on rate constants and reorganization energies. J. Electroanal. Chem. 2000, 483, 2–6. 10.1016/s0022-0728(00)00011-5. [DOI] [Google Scholar]
- Marcus R. A. Exchange reactions and electron transfer reactions including isotopic exchange. Theory of oxidation-reduction reactions involving electron transfer. Part 4.—A statistical-mechanical basis for treating contributions from solvent, ligands, and inert salt. Annu. Rev. Phys. Chem. 1960, 29, 21–31. 10.1039/df9602900021. [DOI] [Google Scholar]
- Marcus R. A. Chemical and Electrochemical Electron-Transfer Theory. Annu. Rev. Phys. Chem. 1964, 15, 155–196. 10.1146/annurev.pc.15.100164.001103. [DOI] [Google Scholar]
- Marcus R. A.; Sutin N. Electron transfers in chemistry and biology. Biochim. Biophys. Acta, Rev. Bioenerg. 1985, 811, 265–322. 10.1016/0304-4173(85)90014-x. [DOI] [Google Scholar]
- Peters B. Common features of extraordinary rate theories. J. Phys. Chem. B 2015, 119, 6349–6356. 10.1021/acs.jpcb.5b02547. [DOI] [PubMed] [Google Scholar]
- Libby W. F. Theory of Electron Exchange Reactions in Aqueous Solution. J. Phys. Chem. 1952, 56, 863–868. 10.1021/j150499a010. [DOI] [Google Scholar]
- Small D. W.; Matyushov D. V.; Voth G. A. The Theory of Electron Transfer Reactions What May Be Missing. J. Am. Chem. Soc. 2003, 125, 7470–7478. 10.1021/ja029595j. [DOI] [PubMed] [Google Scholar]
- Matyushov D. V.; Voth G. A. Modeling the free energy surfaces of electron transfer in condensed phases. Chem. Phys. 2000, 113, 5413–5424. 10.1063/1.1289886. [DOI] [Google Scholar]
- Matyushov D. V.; Voth G. A. New Developments in the Theoretical Description of Charge Transfer Reactions in Condensed Phases. J. Am. Chem. Soc. 2003, 18, 147–210. 10.1002/0471433519.ch4. [DOI] [Google Scholar]
- LeBard D. N.; Matyushov D. V. Glassy Protein Dynamics and Gigantic Solvent Reorganization Energy of Plastocyanin. J. Phys. Chem. B 2008, 112, 5218–5227. 10.1021/jp709586e. [DOI] [PubMed] [Google Scholar]
- Bueno P. R.; Benites T. A.; Davis J. J. The Mesoscopic Electrochemistry of Molecular Junctions. Sci. Rep. 2016, 6, 18400. 10.1038/srep18400. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barbara P. F.; Meyer T. J.; Ratner M. A. Contemporary Issues in Electron Transfer Research. J. Phys. Chem. 1996, 100, 13148–13168. 10.1021/jp9605663. [DOI] [Google Scholar]
- Piechota E. J.; Meyer G. J. Introduction to Electron Transfer: Theoretical Foundations and Pedagogical Examples. J. Chem. Educ. 2019, 96, 2450–2466. 10.1021/acs.jchemed.9b00489. [DOI] [Google Scholar]
- Miller J. R.; Calcaterra L. T.; Closs G. L. Intramolecular long distance electron transfer in radical anions. The effects of free energy and solvent on the reaction rates. J. Am. Chem. Soc. 1984, 106, 3047–3049. 10.1021/ja00322a058. [DOI] [Google Scholar]
- Imahori H.; Tamaki K.; Guldi D.; Luo C. P.; Fujitsuka M.; Ito O.; Sakata Y.; Fukuzumi S. Modulating Charge Separation and Charge Recombination Dynamics in Porphyrin-Fullerene Linked Dyads and Triads: Marcus-Normal versus Inverted Region. J. Am. Chem. Soc. 2001, 123, 2607–2617. 10.1021/ja003346i. [DOI] [PubMed] [Google Scholar]
- Chaudhuri S.; Hedström S.; Méndez-Hernández D. D.; Hendrickson H. P.; Jung K. A.; Ho J.; Batista V. S. Electron Transfer Assisted by Vibronic Coupling from Multiple Modes. J. Chem. Theory Comput. 2017, 13, 6000–6009. 10.1021/acs.jctc.7b00513. [DOI] [PubMed] [Google Scholar]
- Takeuchi E.; Muramatsu M.; Yoneda Y.; Katayama T.; Iwamoto A.; Nagasawa Y.; Miyasaka H. Vibrational decoherence induced by ultrafast intramolecular charge separation of an asymmetric bianthryl derivative. J. Chem. Phys. 2020, 153, 84307. 10.1063/5.0018482. [DOI] [PubMed] [Google Scholar]
- Deng W.-Q.; Goddard W. A. Predictions of Hole Mobilities in Oligoacene Organic Semiconductors from Quantum Mechanical Calculations. J. Phys. Chem. B 2004, 108, 8614–8621. 10.1021/jp0495848. [DOI] [Google Scholar]
- Coropceanu V.; Cornil J.; da Silva Filho D. A.; Olivier Y.; Silbey R.; Brédas J.-L. Charge Transport in Organic Semiconductors. Chem. Rev. 2007, 107, 926–952. 10.1021/cr050140x. [DOI] [PubMed] [Google Scholar]
- Zhao S.; Yu F.; Yang G.; Zhang H.; Su Z.; Wang Y. Theoretical study on the charge transport property of Pt(CNtBu)2(CN)2 nanowires induced by Pt-Pt interactions. Dalton Trans. 2012, 41, 7272–7277. 10.1039/c2dt00009a. [DOI] [PubMed] [Google Scholar]
- Schein L. B.; McGhie A. R. Band-hopping mobility transition in naphthalene and deuterated naphthalene. Phys. Rev. B: Condens. Matter Mater. Phys. 1979, 20, 1631–1639. 10.1103/physrevb.20.1631. [DOI] [Google Scholar]
- Bassani G. F.; Liedl G. L.; Wyder P.. Encyclopedia of Condensed Matter Physics; Elsevier: Amsterdam, 2005. [Google Scholar]
- Huang K.Statistical Mechanics, 2nd ed.; Wiley and Sons, 1987. [Google Scholar]
- Landau L. D.; Lifschitz E. M.. Statistical Mechanics; Elsevier, 2005. [Google Scholar]
- Morikazu T.; Kubo R.; Saitô N.. Statistical Physics I Equilibrium Statistical Mechanics; Springer Berlin Heidelberg, 1983. [Google Scholar]
- Feynman R.Statistical Mechanics; CRC Press, 1998. [Google Scholar]
- Pathria R.; Beale P. D.. Statistical Mechanics, 4th ed.; Academic Press, 2009. [Google Scholar]
- Economou E. N.The Physics of Solids: Essentials and Beyond; Springer: Berlin, Heidelberg, 2010. [Google Scholar]
- Galbiati M.; Motta N.; De Crescenzi M.; Camilli L. Group-IV 2D materials beyond graphene on nonmetal substrates: Challenges, recent progress, and future perspectives. Appl. Phys. Rev. 2019, 6, 41310. 10.1063/1.5121276. [DOI] [Google Scholar]
- Kittel C.Introduction to Solid State Physics; John Wiley & Sons: New York, 1986. [Google Scholar]
- Madelung O.Introduction to Solid State Physics; Springer Berlin Heidelberg, 1978. [Google Scholar]
- Ashcroft N. W.; Mermin N.. Solid State Physics; Holt, Rinehart & Winston, 2002. [Google Scholar]
- Anderson P. W.Concepts in Solids: Lectures on the Theory of Solids; World Scientific: Singapore, 1997. [Google Scholar]
- Hutchison G. R.; Zhao Y.-J.; Delley B.; Freeman A. J.; Ratner M. A.; Marks T. J. Electronic structure of conducting polymers: Limitations of oligomer extrapolation approximations and effects of heteroatoms. Phys. Rev. B: Condens. Matter Mater. Phys. 2003, 68, 35204. 10.1103/physrevb.68.035204. [DOI] [Google Scholar]
- Landauer R. Electrical resistance of disordered one-dimensional lattices. The Philosophical Magazine: J. Theor. Appl. Phys. 1970, 21, 863–867. 10.1080/14786437008238472. [DOI] [Google Scholar]
- Economou E. N.; Soukoulis C. M. Static Conductance and Scaling Theory of Localization in One Dimension. Phys. Rev. Lett. 1981, 46, 618–621. 10.1103/physrevlett.46.618. [DOI] [Google Scholar]
- Enz C. P.; Flores J. C. Localization and cross-over in linear models of random arrays of barriers. Helv. Phys. Acta 1988, 61, 1079–1086. 10.5169/seals-115983. [DOI] [Google Scholar]
- Enz C. P. A.Course on Many-Body Theory Applied to Solid-State Physics; World Scientific, 1992. [Google Scholar]
- Thirring W. Systems with Negative Specific Heat. Phys. Z. 1970, 235, 339–352. 10.1007/bf01403177. [DOI] [Google Scholar]
- Berry R.; Smirnov B. Heat capacity of isolated clusters. J. Exp. Theor. Phys. 2004, 98, 366–373. 10.1134/1.1675906. [DOI] [Google Scholar]
- Lynden-Bell D.; Wood R.; Royal A. The gravo-thermal catastrophe in isothermal spheres and the onset of red-giant structure for stellar systems. Mon. Not. Roy. Astron. Soc. 1968, 138, 495–525. 10.1093/mnras/138.4.495. [DOI] [Google Scholar]
- Schmidt M.; Kusche R.; Hippler T.; Donges J.; Kronmüller W.; von Issendorff B.; Haberland H. Negative Heat Capacity for a Cluster of 147 Sodium Atoms. Phys. Rev. Lett. 2001, 86, 1191–1194. 10.1103/physrevlett.86.1191. [DOI] [PubMed] [Google Scholar]
- Eryurek M.; Guven M. H. Thermodynamic Properties of Ar39 Cluster. AIP Conf. Proc. 2007, 899, 171–172. 10.1063/1.2733091. [DOI] [Google Scholar]
- Gross D. H. E.; Kenney J. F. The microcanonical thermodynamics of finite systems: The microscopic origin of condensation and phase separations, and the conditions for heat flow from lower to higher temperatures. Chem. Phys. 2005, 122, 224111. 10.1063/1.1901658. [DOI] [PubMed] [Google Scholar]
- Kinoshita M.; Yoshidome T. Molecular origin of the negative heat capacity of hydrophilic hydration. Chem. Phys. 2009, 130, 144705. 10.1063/1.3112610. [DOI] [PubMed] [Google Scholar]
- Serra P.; Carignano M. A.; Alharbi F. H.; Kais S. Quantum confinement and negative heat capacity. Europhys. Lett. 2013, 104, 16004. 10.1209/0295-5075/104/16004. [DOI] [Google Scholar]
- Chomaz P.; Gulminelli F. Phase Transition in Small System. Nucl. Phys. A 2005, 749, 3–13. 10.1016/j.nuclphysa.2004.12.003. [DOI] [Google Scholar]
- Flores J. C.; Palma-Chilla L. Theoretical thermodynamics connections between Dual (Left-Handed) and Direct (Right Handed) systems: Entropy, temperature, pressure and heat capacity. Phys. Rev. B: Condens. Matter Mater. Phys. 2015, 476, 88–90. 10.1016/j.physb.2015.07.007. [DOI] [Google Scholar]
- Palma-Chilla L.; Flores J. Negative heat capacity in a left-handed system. Phys. A 2017, 471, 396–401. 10.1016/j.physa.2016.12.020. [DOI] [Google Scholar]
- Velázquez L.; Curilef S. On the thermodynamic stability of macrostates with negative heat capacities. J. Stat. Mech.: Theory Exp. 2009, 2009, P03027. 10.1088/1742-5468/2009/03/p03027. [DOI] [Google Scholar]
- Staniscia F.; Turchi A.; Fanelli D.; Chavanis P. H.; De Ninno G. Negative Specific Heat in the Canonical Statistical Ensemble. Phys. Rev. Lett. 2010, 105, 10601. 10.1103/physrevlett.105.010601. [DOI] [PubMed] [Google Scholar]
- Carignano M. A.; Gladich I. Negative heat capacity of small systems in the microcanonical ensemble. Europhys. Lett. 2010, 90, 63001. 10.1209/0295-5075/90/63001. [DOI] [Google Scholar]
- Shoaee S.; Stolterfoht M.; Neher D. The Role of Mobility on Charge Generation, Recombination, and Extraction in Polymer-Based Solar Cells. Org. Electron. 2018, 8, 1703355. 10.1002/aenm.201703355. [DOI] [Google Scholar]
- Serdiuk I. E.; Mońka M.; Kozakiewicz K.; Liberek B.; Bojarski P.; Park S. Y. Vibrationally assisted direct intersystem crossing between the same charge-transfer states for thermally activated delayed fluorescence: Analysis by Marcus Hush theory including reorganization energy. J. Phys. Chem. B 2021, 125, 2696–2706. 10.1021/acs.jpcb.0c10605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y.; Li Y.; Chen C.; Wang L.; Zhang J. Design new hole transport materials for efficient perovskite solar cells by suitable combination of donor and core groups. Org. Electron. 2017, 49, 255–261. 10.1016/j.orgel.2017.06.064. [DOI] [Google Scholar]
- Qiu M.; Pei W.; Lu Q.; Li Z.; Li Y.; Liang J. DFT Characteristics of Charge Transport in DBTP-Based Hole Transport Materials. Appl. Sci. 2019, 9, 2244. 10.3390/app9112244. [DOI] [Google Scholar]
- Wang Q.; Li Y.; Song P.; Su R.; Ma F.; Yang Y. Non-Fullerene Acceptor-Based Solar Cells: From Structural Design to Interface Charge Separation and Charge Transport. Polymers 2017, 9, 692. 10.3390/polym9120692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li S.-B.; Duan Y.-A.; Geng Y.; Li H.-B.; Zhang J.-Z.; Xu H.-L.; Zhang M.; Su Z.-M. A designed bithiopheneimide-based conjugated polymer for organic photovoltaic with ultrafast charge transfer at donor/PC71BM interface: theoretical study and characterization. Phys. Chem. Chem. Phys. 2014, 16, 25799–25808. 10.1039/c4cp03022b. [DOI] [PubMed] [Google Scholar]
- Zhang X.; Yang X.; Geng H.; Nan G.; Sun X.; Xi J.; Xu X. Theoretical comparative studies on transport properties of pentacene, pentathienoacene, and 6,13-dichloropentacene. J. Comput. Chem. 2015, 36, 891–900. 10.1002/jcc.23870. [DOI] [PubMed] [Google Scholar]
- Fan J.-X.; Chen X.-K.; Zhang S.-F.; Ren A.-M. Theoretical Study on Charge Transport Properties of Intra- and Extra-Ring Substituted Pentacene Derivatives. J. Phys. Chem. A 2016, 120, 2390–2400. 10.1021/acs.jpca.5b12641. [DOI] [PubMed] [Google Scholar]
- Delgado M.; Kim E. G.; Filho D. A.; Bredas J. L. Tuning the charge-transport parameters of perylene diimide single crystals via end and/or core functionalization: A density functional theory investigation. J. Am. Chem. Soc. 2010, 132, 3375–3387. 10.1021/ja908173x. [DOI] [PubMed] [Google Scholar]
- da Silva Filho D. A.; Kim E. G.; Brédas J. L. Transport properties in the rubrene crystal: Electronic coupling and vibrational reorganization energy. Adv. Mater. 2005, 17, 1072–1076. 10.1002/adma.200401866. [DOI] [Google Scholar]
- Sundar V. C.; Zaumseil J.; Podzorov V.; Menard E.; Willett R. L.; Someya T.; Gershenson M. E.; Rogers J. A. Elastomeric Transistor Stamps: Reversible Probing of Charge Transport in Organic Crystals. Science 2004, 303, 1644–1646. 10.1126/science.1094196. [DOI] [PubMed] [Google Scholar]
- Sakanoue T.; Sirringhaus H. Band-like temperature dependence of mobility in a solution-processed organic semiconductor. Synth. Met. 2010, 9, 736–740. 10.1038/nmat2825. [DOI] [PubMed] [Google Scholar]
- Shuai Z.; Li W.; Ren J.; Jiang Y.; Geng H. Applying Marcus theory to describe the carrier transports in organic semiconductors: Limitations and beyond. Chem. Phys. 2020, 153, 80902. 10.1063/5.0018312. [DOI] [PubMed] [Google Scholar]
- Geng H.; Peng Q.; Wang L.; Li H.; Liao Y.; Ma Z.; Shuai Z. Toward quantitative prediction of charge mobility in organic semiconductors: Tunneling enabled hopping model. Adv. Mater. 2012, 24, 3568–3572. 10.1002/adma.201104454. [DOI] [PubMed] [Google Scholar]
- Harima Y.; Kubota K.; Ishiguro Y.; Ooyama Y.; Imae I. Electrical Characteristics of Pentacene Films on Cross-Linked Polymeric Insulators of Varying Thicknesses. ACS Omega 2016, 1, 784–788. 10.1021/acsomega.6b00292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McKelvey J. P.Solid State and Semiconductor Physics; Harper’s Physics Series; Krieger Publishing Company, 1982. [Google Scholar]
- de Wijs G. A.; Mattheus C. C.; de Groot R. A.; Palstra T. T. Anisotropy of the mobility of pentacene from frustration. Synth. Met. 2003, 139, 109–114. 10.1016/s0379-6779(03)00020-1. [DOI] [Google Scholar]
- Li Z. Q.; Podzorov V.; Sai N.; Martin M. C.; Gershenson M. E.; Di Ventra M.; Basov D. N. Light quasiparticles dominate electronic transport in molecular crystal field-effect transistors. Phys. Rev. Lett. 2007, 99, 16403. 10.1103/physrevlett.99.016403. [DOI] [PubMed] [Google Scholar]
- Yamamoto A.; Murata Y.; Mitsui C.; Ishii H.; Yamagishi M.; Yano M.; Sato H.; Yamano A.; Takeya J.; Okamoto T. Zigzag-Elongated Fused π-Electronic Core: A Molecular Design Strategy to Maximize Charge-Carrier Mobility. Adv. Sci. 2018, 5, 1700317. 10.1002/advs.201700317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Messiah A.Quantum Mechanics; Dover Books on Physics Dover Publications, 2014. [Google Scholar]
- Gasiorowicz S.Quantum Physics; John Wiley: New York, 2003. [Google Scholar]
- Landau L. D.Quantum Quantum Mechanics, 3rd ed.; Pergamon, 1977. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
