Mike
Castellano
a,
Christoph
Kaspar
b,
Michael
Thoss
b and
Thorsten
Koslowski
*a
aInstitut für Physikalische Chemie, Universität Freiburg, Albertstraße 21, 79104 Freiburg, Germany. E-mail: koslowsk@uni-freiburg.de
bInstitut für Physik, Universität Freiburg, Hermann-Herder-Straße 3, 79104 Freiburg, Germany
First published on 6th November 2023
Potential differences for protein-assisted electron transfer across lipid bilayers or in bio-nano setups can amount to several 100 mV; they lie far outside the range of linear response theory. We describe these situations by Pauli-master equations that are based on Marcus theory of charge transfer between self-trapped electrons and that obey Kirchhoff's current law. In addition, we take on-site blockade effects and a full non-linear response of the local potentials into account. We present analytical and numerical current–potential curves and electron populations for multi-site model systems and biological electron transfer chains. Based on these, we provide empirical rules for electron populations and chemical potentials along the chain. The Pauli-master mean-field results are validated by kinetic Monte Carlo simulations. We briefly discuss the biochemical and evolutionary aspects of our findings.
Within a cycle of work, both of these enzymes transfer two electrons. The biochemically controlled transfer of electrons is not limited to pairs: the NrfHA nitrite reductase catalyses the reduction of nitrite to an ammonium cation, a reaction that involves the transfer of six electrons. In its minimum functional unit, NrfHAn (n = 1 or 2) contains nine to 14 heme groups, and the enzyme is attached to the inner membrane of bacteria by a hydrophobic anchor.10
The structural biology of membrane proteins has seen considerable progress in recent years, in particular due to the availability of high-resolution cryo transmission electron microscopy.11,12 In addition, the combination of redox titration experiments and in situ spectroscopy has added to insight into the energetics of electron transfer reactions, often down to pinning individual transfer steps and the redox states of single electron acceptors. Usually, these studies are performed at thermodynamic equilibrium in aqueous solution, while the biologically operative protein complexes are associated with membranes and experience a redox potential gradient.13 For example, electron transfer in the respiratory complex I is driven by the standard redox potential difference of at least 520 mV between NAD+/NADH and 1,4-naphthoquinone-based mobile redox carriers.14,15 For NrfH, we find a standard potential difference of 650 mV between the nitrite and the ammonium ion.16 For light-induced processes, potential differences span a considerable fraction of the spectrum of visible light.
The situations under review in this paper are depicted in Fig. 1 as cartoon models. Fig. 1a shows a redox-active protein at equilibrium, the redox states of the excess electron centers of localization can be populated by tuning an external chemical potential, μred. In Fig. 1c, the protein is embedded in a membrane and subject to two chemical potentials, to which we will refer to as μL and μR. Such membranes, e.g. lipid bilayers, separate cell or bacterial compartments in which different physical and chemical conditions can be realized. In the case studied here, this compartmentalization is essential for realizing different chemical potentials on either side of the bilayer, their difference drives the electron transport process. These chemical potentials depend on the standard redox potential and on the concentration of the redox-active species on either side of the membrane. The major aim of our work is the computation of the charge current, the charge carrier populations and the chemical potentials of the electron acceptors as a function of μL, μR and the energies characteristic of biological charge transfer. The situation of a protein bridging two compartments is related to that of a biopolymer spanning metallic contacts or leads, as shown in Fig. 1b. The major physical difference to Fig. 1c is the presence of a continuum of states above and below μL and μR rather than a single state representing the chemical potential.
The setup shown in Fig. 1b is reminiscent of that found in the field of molecular electronics, i.e. small- to medium-sized molecules spanning a small gap or bridging a scanning tunneling microscope tip and a surface.17 In nanostructures such as molecular junctions conductivity is often accompanied by strong coupling of the transport electrons to the mechanical degrees of freedom such as local vibrational modes or phonons.18,19 For higher bias voltages, i.e., far from equilibrium, this coupling gives rise to non-conservative forces and unusual dissipative phenomena, which have important consequences. For example, theoretical studies have predicted that in situations far from equilibrium current-induced forces may cause so-called runaway vibrational modes. Furthermore, negative friction may exist, both resulting in mechanical instability of the structures. Theoretical studies of these phenomena have, for example, been based on quantum mechanical rate theories,20 classical concepts such as Langevin equations21,22 or the hierarchical equations of motion approach.19,23 The latter is a density matrix approach that provides a fully quantum mechanical framework to study nonequilibrium transport in open quantum systems.
There are, however, major differences between molecular junctions and biopolymers, which render the application of the schemes listed above difficult from a numerical perspective. Proteins and DNA exist in aqueous environment, and the negative DNA glycophosphate backbone charges have to be compensated. As a consequence, these systems are dominated by strong electron–phonon coupling, emerging from a continuous vibrational density of states that ranges down to the Terahertz range of the electromagnetic spectrum. As detailed below, this coupling is expressed in a single characteristic quantity, the reorganization energy λ. For a single site that bridges two metallic contacts, models of this type have been studied by Sowa et al.24,25 In our paper, we discuss both electron transport chains coupled to single entry and exit levels and chains coupled to metallic leads, as they are intimately related mathematically and computationally.
As a biopolymer that exhibits long-range electron transfer, DNA has been in the spotlight for many years. Two different types of experiments have been in the focus of research. Light-induced DNA fragmentation has been used in the work of Barton, Giese, Schuster, Wagenknecht, to name a few.26–29 Here, a transition metal complex intercalated into DNA induces charge separation upon irradiation, and an excess hole can travel along the DNA strand until it is captured by a (GC)3 moiety, with consecutive oxidative fission in this place. DNA amplification and consecutive chromatography has revealed a notably large length scale of charge hopping up to ∼200 Å. In a second type of experiments, short DNA strands bridge a metal surface and a scanning electron microscope tip, permitting the observation of current–voltage curves. Upon the application of a voltage drop of 2–4 V along the strands, conductivities amount to some 10 nA.30 Both processes have been modeled theoretically and computationally.31–33 While for the latter, the current–voltage curves have been reproduced quantitatively, approximations have been made, such as postulating a linear voltage drop along the chain, or restrictions upon the total number of charge carriers that populate the chain. With application to a different kind of biopolymers, electron transfer proteins, we present a scheme that overcomes these constraints and approximations.
The remaining part of this work is organized as follows. In the next section, we present the model, which is based upon hopping conduction in a polarizing environment. We also detail the computational approach that is used to solve the resulting Pauli-master equations at constant flow, and a kinetic Monte Carlo approach that introduces a stochastic element to the description of the transport process. In the third section, analytical results are given for a specific set of parameters. In addition, computational results are presented in the third section for setups that model proteins spanning membranes and nanoscopic setups. In the final section, the results are discussed in a general context, and conclusions are derived.
![]() | (1) |
Away from equilibrium, we have to take into account the drop in electronic chemical potential (or the voltage) along the transport chain, which modifies the effective local driving force and leads to
![]() | (2) |
As the centers of localization accommodate excess charge on a molecular scale, they exhibit a strong on-site electron–electron repulsion, usually limiting the number of excess electrons or holes to a single one. From a mean-field perspective, this local blockade effect can be taken into account using the local excess charge carrier population pi. From this point of view, charge transfer from a site i to a site j occurs with a probability proportional to pi(1 − pj). In the end, we will drop the mean-field character of the populations p and simulate charge transfer between states characterized by binary integer charges using a kinetic Monte Carlo algorithm.
All the elements described above can be arranged into a system of Pauli-master equations:
![]() | (3) |
ki−1,ipi−1(1 − pi) − ki,i−1pi(1 − pi−1) = ki,i+1pi(1 − pi+1) − ki+1,ipi+1(1 − pi). | (4) |
![]() | (5) |
![]() | ||
Fig. 2 Conductivity model used in this work. A chain of n sites 1,…,i, j,…,n is connected to a left (L) and right (R) lead, which exhibit a constant potential. Hopping rates kij within the system, and kL, kR, ![]() |
For a chain coupled to a left and a right single-level redox potential μL and μR (the situation depicted in Fig. 1c), the following equations hold for the entry and the exit site. For the first site, we have
![]() | (6) |
![]() | (7) |
For the left lead, we now have to modify the f for the entry process
![]() | (8) |
![]() | (9) |
For small populations, small reorganization energies and small chemical potential differences (the latter two compared to kBT), we recover Ohmic conductivity and Kirchhoff's current law by transforming the chemical potentials μ into voltages U,
![]() | (10) |
Detailed balance is satisfied under those conditions where Boltzmann statistics can be applied, i.e. in the absence of external driving forces and for a number of particles that is much smaller than the number of states, i.e. pi ≪ 1.
A
i
, Bi, Ci and Di depend on the n populations, which can be assembled into a vector, , and on the n local chemical potentials,
. The resulting system of n equations is nonlinear, and it is subject to the constraints 0 ≤ pi ≤ 1 ∀i. For vanishing equilibrium driving forces, ΔGij = 0, and a linear topology, μL ≥ μi ≥ μR∀i holds. In this situation, the chemical potentials also have to obey μi+1 ≤ μi. For a given
, we have
Ai − Bi − Ci + Di = 0 = gi(![]() | (11) |
![]() ![]() ![]() | (12) |
J(![]() ![]() ![]() | (13) |
To solve for and
simultaneously, we follow the strategy of Koslowski and Wilkening applied in the Ohmic regime with an added local blockade and backward transfer.44 Starting from an initial guess for both quantities, the μi are subject to a Monte Carlo (MC) optimization. In each MC step, a random site i is selected, and the corresponding chemical potential μi is altered by ±Δμi, which is drawn from a uniform distribution. The change is accepted if μi fulfills the constraints listed above – if applicable – and if the conductivity increases. For the old and the new μi, we solve for
old and
new by Newton's method. Eqn (13) is solved by the dgesl and dgefa subroutines of the linpack library.45 For small system sizes, the potentials may also be scanned by n coarse-grained outer loops in a brute force manner.
We make use of a rejection Monte Carlo algorithm,46,47 starting with a randomly populated system at t = 0, and we calculate its index m. We choose one of the states out of L accessible by m, and compute the rate km
by identifying the corresponding electron hopping process and its rate. The move is accepted with a probability km
/kmax, where kmax is the maximum of all hopping rates. We compute a second random number 0 < r ≤ 1 and increment the time by Δt = −(Lkmax)−1
ln
r. We loop through the process until a time limit is reached or a predefined number of electrons has passed through the system.
![]() | (14) |
![]() | ||
Fig. 3 Analytical solution (eqn (14)) for the current as a function of the external drop in the chemical potential for a reorganization energy of λ = 0.75 eV and n = 1 (red, solid line) and n = 3 (black, solid line) centers of excess electron localization. Open circles, ○, depict the results of kinetic Monte Carlo simulations for n = 3. The underlying model is shown in Fig. 1c. We also present numerical results for metallic leads and n = 3, as depicted in Fig. 1b. The results are shown as a black chain line. |
It is interesting to compare these analytical results obtained for systems coupled to two redox potentials (Fig. 1c) to numerical values computed for those attached to metallic leads (Fig. 1b). For identical values of k0 and λ as used above, the current-chemical potential curve is shown in Fig. 3 for a system consisting of three chromophores. As compared to the Marcus result, the system attached to conducting leads is slightly shifted to higher values of Δμ, and it saturates slightly above Δμ = λ(n + 1) to a maximum current. This behaviour is due to the presence of metallic states in a uniform density of states below and above the Fermi level, which are always available as a source or drain of electrons.
Eqn (14) can be expanded into a Taylor series around Δμ = 0. With λ = 1 eV, at room temperature (kBT = 0.025 eV) and with k0 = 0.25 (in arbitrary units), we have two leading terms of 20exp(−10)Δ
and 3400
exp(−10)Δ
3/3. For n = 1, a linearization gives rise to a relative error that is less than 10% for Δμ < 80 meV. For uniform chains, this range becomes more extended with increasing n. This energy scale has to be compared to ±2 V that can be achieved in a bio-nano setup or to the biologically relevant differences in redox potentials of up to 500 meV referenced above.
For a two-site system, we have three rates, of which we set two to k0 and use the third (k0L, k0R or k012) as a bottleneck by decreasing the rate by a factor of 100. We also study an additional symmetric system with a small but equal entry and exit hopping rate. We apply external chemical potential differences of 0.5 and 1 eV, the former defining a maximum of biological ground state protein electron transfer, the latter lying in the range of nanoscopic setups involving narrow-band metals or doped semiconductors. In Table 1, we list the populations and the chemical potentials calculated for these values of Δμ. All of these calculations refer to the model with single level contacts depicted in Fig. 1c.
System | μ L (eV) | p 1 | p 2 | μ 1 (eV) | μ 2 (eV) |
---|---|---|---|---|---|
L-1·2-R | 0.5 | 0.66 | 0.34 | 0.43 | 0.07 |
L-1·2-R | 1.0 | 0.77 | 0.23 | 0.80 | 0.20 |
L·1-2-R | 0.5 | 0.38 | 0.48 | 0.16 | 0.06 |
L·1-2-R | 1.0 | 0.26 | 0.46 | 0.42 | 0.17 |
L-1-2·R | 0.5 | 0.52 | 0.62 | 0.44 | 0.34 |
L-1-2·R | 1.0 | 0.54 | 0.74 | 0.83 | 0.58 |
L·1-2·R | 0.5 | 0.40 | 0.60 | 0.27 | 0.23 |
L·1-2·R | 1.0 | 0.36 | 0.64 | 0.58 | 0.42 |
Regardless of the Δμ value applied, we always find the largest drops in the chemical potential between the sites – including the leads – that exhibit the smallest hopping rates. The hopping rates of the first and the last model listed in Table 1 are symmetric w.r.t. to an exchange of the hopping rates, this symmetry is reflected both in the populations and chemical potentials. Populations add up to unity for the symmetric models, and the individual contributions can deviate considerably from the analytical solutions for uniform models with p = 1/2, as computed above. This deviations increase with an increasing Δμ. It is important to note that nonuniform hopping rates lead to changes both in the populations pi and in the chemical potentials μi, as compared to the uniform case and its analytical solution. In retrospect, this behaviour justifies the strategy of treating both quantities as variables, and of designing a scheme that permits their simultaneous computation.
We have also tested our numerical approach for two charge proteins of biological relevance. Our choice of the proteins has been motivated by our previous experience with the systems and the availability of reliable computed energy landscapes for electron transfer. Furthermore, we believe that NrfH is a very suitable candidate for a protein in bio-nano setup. It has a membrane anchor that can be attached to a lipid covering a conducting metal surface on one end, and a free valence on the heme iron on the other end that can be contacted e.g. by a His–Ni moiety.
First, we turn to an electron transfer protein that has been mentioned in the introduction, the cytochrome subunit of the photoreaction center of the purple bacterium Rps. viridis. This protein contains four heme groups, and its geometry is displayed in Fig. 4 using the X-ray structure of Deisenhofer et al.7 Following the approach of Gnandt et al.,38 the driving forces and the reorganization energies have been computed by a numerical solution of the Poisson–Boltzmann equation applying the Delphi program package.48,49 For the atomic charges and radii that enter these computations, we have used the values of the Amber force field.50,51 Within the protein and the lipid bilayer, a dielectric constant of ε = 3.5 has been used; for water, we take the room temperature value ε = 78.4. Rather than computing the electronic coupling that enters eqn (1), we apply the empirical Dutton–Moser rule36,52 to estimate the value of k0. The rule relates k0 to the edge-to-edge distance between two chromophores, r, via
k0(r) = k0(r0)exp(−γ(r − r0)) | (15) |
![]() | ||
Fig. 4 Protein structures used in the work. Amino acid atoms drawn as grey spheres, non-iron heme atoms as red sticks, iron atoms as orange spheres. (a) Cytochrome subunit of the photoreaction center of Rps. viridis after the X-ray structure of Deisenhofer et al.,7 protein databank identifier 1PRC, (b) NrfH subunit of the nitrite reductase of D. vulgaris after the X-ray structure of Rodrigues et al.,10 protein databank identifier 2J7A. The hemes H1 and 335 are closest to the membrane, respectively. |
Reaction | k 0 (s−1) | ΔG (eV) | λ (eV) |
---|---|---|---|
L → 333 | 8.00 × 1010 | 0.00 | 1.0 |
333 → 334 | 7.45 × 1010 | 0.05 | 0.81 |
334 → 336 | 1.84 × 1010 | −0.27 | 0.72 |
336 → 335 | 8.57 × 1010 | −0.28 | 0.82 |
335 → R | 8.00 × 1010 | 0.00 | 1.0 |
Reaction | k 0 (s−1) | ΔG (eV) | λ (eV) |
---|---|---|---|
L → H1 | 1.00 × 1013 | 0.00 | 1.0 |
H1 → H2 | 1.32 × 1013 | −0.61 | 1.37 |
H2 → H3 | 4.60 × 1011 | 0.00 | 1.3 |
H3 → H4 | 4.32 × 1012 | 0.07 | 1.64 |
H4 → R | 1.00 × 1013 | 0.00 | 1.0 |
We assume that the parameters relevant for the contacts of proteins to metallic leads are essentially governed by the same relations as those relevant to charge transfer within the protein, i.e. a dependence of the tunnel splitting upon the contact-chromophore distance that follows an exponential Dutton–Moser law, a reorganization energy that is determined by the dielectric response of the aqueous environment of the system, and with driving forces that can be tuned by the material of the contacts and the external voltages.
For the contacts, we use ΔGL = ΔGR = 0, λ = 1 eV and a k0L entry and a k0R exit hopping rate close to the maximum of k0ij within the chain.
For the cytochrome subunit, we have a monotonous drop of ΔG along the transport chain, a nearly uniform λ and a hopping rate bottleneck between the hemes 334 and 336 as the parameters entering the computation of the currents, the populations and the local chemical potentials as a function of Δμ. Within the computations, μR is fixed and defines the zero of the chemical potential, while μL is varied. We note that the driving forces depend on the protonation pattern of the cytochrome,53 and that the charge flow in the biological system is still unknown due to the lack of the identification of the electron donor binding site.
The current-chemical potential curves are displayed in Fig. 5 for single-level (cf.Fig. 1c) and metallic (Fig. 1b) contacts. As above, the curve in the positive Δμ branch is shifted to slightly higher chemical potentials for a computation involving metallic leads. The curves are asymmetric w.r.t. Δμ = 0 and exhibit a current that is close to zero in range from −1.5 eV to 0.2 eV. This behaviour partly reflects the extent of the original free energy surface (cf.Table 2): barriers of 0.05 eV (transfer to the left) or 0.65 eV (transfer to the right) have to be overcome to facilitate electron transfer.
![]() | ||
Fig. 5 Currents I as a function of the applied difference in the external chemical potential, Δμ for two four-heme protein subunits. Red, solid line: NrfH, single-level contacts; black, solid line: photoreaction center cytochrome, single-level contacts; (both as in Fig. 1c); red, dashed line: NrfH, metallic leads; black, dashed line: photoreaction center, metallic leads (the latter two as in Fig. 1b). Note that for the latter curves, the current has been multiplied by a factor of 100. |
Similar curves have been obtained for electron transfer through the NrfH protein, which also constitutes a charge transfer process that is essentially downhill. They are also shown in Fig. 5. Compared to the cytochrome, the currents are about an order of magnitude larger throughout the entire range of Δμ. Here, larger λ values are overcompensated by a tighter packing of the chromophores, which in turn leads to larger values of k0. We note that the Amber charges usually lead to an overestimate of the reorganization energies54 and that we have not applied the so-called Pekar factor that reduces the reorganization energies. At Δμ = 1 eV, the currents lie in the range of 100 to 500 pA, which puts them into the range of scanning tunnel microscope or patch clamp experiments. We note that currents of 100 pA can be detected with an accuracy of 1 ppm using commercially available instrumentation.55 For biological soft-matter systems, ionic currents around 1 pA can be measured using patch clamp techniques,56,57 electronic currents recorded by scanning tunnel microscope experiments permit a resolution much better than 1 nA.30
While the currents as a function Δμ are as continuous as the local chemical potentials, the local populations show a different behaviour. For NrfH, they are displayed in Fig. 6. Approaching the range of a particularly small conductivity from the left, p1 and p4 show values in the range of 0.5 to 0.85, while p2 and p3 gradually converge to unity and hence block the conductivity channel by effectively taking out these sites as electron acceptors, cf.eqn (3)–(6). At zero voltage, p2 starts to decline, and the channel is unblocked with p1, p3 and p4 not far away from one half. We find a very similar behaviour both for single-site and metallic contacts, and overlapping curves reflecting these situations are not shown in Fig. 6.
![]() | ||
Fig. 6 Excess electron populations for the four heme molecules of the NrfH nitrite reductase protein chain as a function of the external chemical potential, Δμ. Red open symbols and line: H1, single-level contacts, black open symbols: H2, single-level contacts, blue open symbols: H3, single-level contacts, green open symbols: H4, single-level contacts, the underlying model is depicted in Fig. 1c. Full symbols (•) refer to coupling to metallic leads, as shown in Fig. 1b. |
In this way, we are not only able to show that the mean-field solutions for the populations are identical to their numerical mean-field counterparts, but recover the underlying microscopic picture. The objects of hopping transport are electrons, and sites of excess electron localization exhibit a charge of zero or the elementary charge. As a simple argument from dielectric theory suggests that the reorganization energy is proportional to the square of the charge,34,35 a hopping electron experiences the full reorganization energy λ rather than λ/4, as a mean-field value of p = 1/2 suggests.
We are not aware of a rigorous deviation of the nonlinear master equations from an extended linear scheme. As we obtain the same numerical results for a linear Fock space representation as for the mean-field nonlinear site representation (cf.Fig. 3), we are confident that the introduction of the nonlinear terms is valid.
Analytical results have been derived for a special set of parameters, from which we conclude a general trend towards an average site occupation of one half and a linear voltage – or chemical potential – drop along an electron transport chain with a linear topology. Introducing bottlenecks, we have to resort to numerical solutions that optimize the occupations and the local chemical populations simultaneously. Around these bottlenecks, both the chemical potentials and the electron populations show a drop. In retrospect, this justifies the setup of a considerable numerical apparatus that treats both quantities on an equal footing. We have verified and rationalized our methodology by relaxing the mean-field approach to one of these variables, the populations, via kinetic Monte Carlo simulations. We have demonstrated the application of the scheme to four-center proteins of biological relevance. Here, the comparatively simple current-chemical potential characteristics reflect a remarkably complex underlying pattern of chromophore populations. We are confident that the extension of our work to a catalogue of charge transfer proteins will lead to new perspectives and microscopic interpretations of biological electron transport.
The computational approach applied is robust and efficient, for the systems studied here the calculations can be performed within a few minutes on a desktop computer. Nevertheless, we find it desirable to have a more controllable numerical procedure available for the optimization of the local chemical potentials, e.g. by a nonlinear simplex method. An important physical aspect still missing are long-range intersite electron–electron interactions, which can also be obtained using dielectric theory computations.58 We have not yet touched more complex topologies involving parallel conduction channel or dead-end storage sites, although they can be addressed by the methods described here. Furthermore, the high potentials applied in nanoscopic setup may have an impact on the protein structure and on the protonation pattern, which may in turn effect the characteristic parameters of Marcus' theory.
We note that coherence effects in biological charge transfer have been discussed recently,59–61 which lie outside the range of Marcus' theory used here. The equations outlined above can be adapted to corrections of the Marcus scheme, provided they can be formulated in the context of simple rate equations.
Finally, we briefly comment on the biochemical and evolutionary aspects of our findings. Multicenter charge transfer chains can efficiently bridge membranes, or they can store electrons for many-electron redox reactions, as in the NrfHA, cytochrome and respiratory complex I protein complexes referenced above. In addition, the current–voltage characteristics are subject to the parameters of the chain and the number of cofactors. They are the subject of an evolutionary optimization process, which can adapt to requirements such as handling an increased oxygen intake on the mitochondrial level, which can increase by a factor of ten to 65 between states of rest and activity for mammals and birds.62
This journal is © the Owner Societies 2023 |