Abstract

The primary step in the mechanism by which migratory birds sense the Earth’s magnetic field is thought to be the light-induced formation of long-lived magnetically sensitive radical pairs within cryptochrome flavoproteins located in the birds’ retinas. Blue-light absorption by the non-covalently bound flavin chromophore triggers sequential electron transfers along a chain of four tryptophan residues toward the photoexcited flavin. The recently demonstrated ability to express cryptochrome 4a from the night-migratory European robin (Erithacus rubecula), ErCry4a, and to replace each of the tryptophan residues by a redox-inactive phenylalanine offers the prospect of exploring the roles of the four tryptophans. Here, we use ultrafast transient absorption spectroscopy to compare wild type ErCry4a and four mutants having a phenylalanine at different positions in the chain. We find that each of the three tryptophan residues closest to the flavin adds a distinct relaxation component (time constants: 0.5, 30, and 150 ps) in the transient absorption data. The dynamics of the mutant containing a phenylalanine at the fourth position, furthest from the flavin, are very similar to those of wild type ErCry4a, except for a reduced concentration of long-lived radical pairs. The experimental results are evaluated and discussed in the framework of real-time quantum mechanical/molecular mechanical electron transfer simulations based on the density functional-based tight binding approach. This comparison between simulation results and experimental measurements provides a detailed microscopic insight into the sequential electron transfers along the tryptophan chain. Our results offer a route to the study of spin transport and dynamical spin correlations in flavoprotein radical pairs.
1. Introduction
Cryptochromes (Cry) are blue-light-sensitive flavoproteins that have a variety of functions.1−4 They are essential, for example, for maintaining circadian rhythms in animals5,6 and plants7 and for controlling aspects of plant growth8 and flowering.9 Moreover, there is increasing,10 albeit so far indirect, evidence that cryptochromes are involved in the mechanism that allows migratory songbirds to sense the direction of the Earth’s magnetic field.11−14
Some cryptochromes10,15,16 non-covalently bind a flavin adenine dinucleotide (FAD) chromophore linked to the surface of the protein by a ∼2.5 nm chain of aromatic amino acid residues. In plant Crys, the chain comprises three tryptophans (Trp),12,17 while in animal and animal-like Crys, there are often four tryptophans.10,18 Blue light absorption via a π → π* transition in the isoalloxazine (flavin) moiety of the fully oxidized FADox ground state triggers a series of electron transfers along the Trp chain and results in the formation of a pair of radicals, [FAD•–TrpH•+], in a spin-correlated electronic singlet state.10,19,20 The lifetime of this state (denoted SRP1), which is typically in the microsecond range,10,11,21 should be long enough to allow a weak magnetic field to modulate the efficiency of coherent interconversion with the corresponding triplet state (TRP1).12,22−24 This process is mainly induced by hyperfine couplings and thus depends on the alignment of nuclear spins coupled to the unpaired electrons.12,18,23 RP1 is anticipated to be sensitive to the direction of an external magnetic field and, therefore, to be a suitable candidate for the magnetic compass sensor in birds.10 RP1 deprotonates on a microsecond time scale, thereby forming a longer-lived, magnetically insensitive, secondary radical pair, [FAD•–Trp•].21 Protonation of FAD•–, to give FADH•, and reduction of Trp• further stabilize the reduced flavin and lead to the formation of a signaling state with an anisotropic quantum yield that encodes the direction of the magnetic field. The signaling state returns to the fully oxidized ground state on a much longer time scale. This light-dependent radical-pair mechanism is thought to be sufficiently sensitive to provide night-migratory birds with magnetic compass information even under dim light conditions.11,25
Light-induced electron transfer along the Trp chain is thus the primary step in the photochemical cycle of cryptochromes. The ultrafast dynamics of flavins and flavoproteins have been the subject of numerous experimental studies during the last two decades.19,26−36 In FAD in aqueous solution, an intramolecular electron transfer from the adenine to the excited flavin occurs in a stacked conformation with a time constant of 5 ps, followed by charge recombination on a faster time scale.31 In the open conformation, at low pH, a much longer, nanosecond lifetime has been measured.31 For protein-bound FAD, rapid electron transfer from a nearby tryptophan residue forms a [FAD•–TrpH•+] radical pair within ∼1 ps in animal type I Crys19,29,37 and with sub-picosecond time constants in plant Crys and photolyases.28,38−41 Dynamics on longer time scales are mainly interpreted in terms of the sequential electron transfer steps along the Trp chain. It remains difficult, however, to unambiguously assign rate constants to the individual charge separation steps due to competing nonequilibrium processes such as vibrational cooling.29
Very recently, it became possible to recombinantly express and purify42 wild type cryptochrome 4a from the night-migratory European robin10 (ErCry4a) with more than 97% FAD-bound.10 Transient absorption spectra of wild type ErCry4a and four mutants, in which one of the tryptophans involved in the electron transfer had been site-selectively replaced by a redox-inactive phenylalanine (Phe), were recorded with sub-nanosecond time resolution.10 The study gave evidence for the creation of light-induced radicals with lifetimes exceeding 100 ns in the wild type protein and in the mutant in which the terminal Trp had been replaced by Phe. The experiments also showed changes in the transient absorption signals of up to 15% produced by ∼10 mT magnetic fields. Due to the limited time resolution, the sequential charge transfer steps could not be resolved experimentally but were estimated computationally based on molecular dynamics (MD) simulations and empirical Moser–Dutton theory. Since those rates are critical for understanding the magnetic sensitivity of ErCry4a, their independent measurement using ultrafast spectroscopy is urgently needed. Such experiments would open new frontiers for probing intraprotein charge and spin transfer dynamics43 and provide important opportunities for experiment/theory comparison.10,18,44,45
In the present study, we provide a comparative ultrafast optical study of the Trp-tetrad electron transfer cascade in wild type ErCry4a and in the four Trp → Phe mutants of ErCry4a. This approach allows us to isolate the dynamics and yields of the first three electron transfers in ErCry4a. It provides a benchmark for modeling radical pair formation in cryptochromes, as explored here by a direct comparison with hybrid quantum-mechanics/molecular mechanics simulations. The emerging synergy between the experiment and simulations provides a first step toward quantitative modeling of electron transport in ErCry4a and forms a basis for the investigation of spin transport and correlations in magnetically sensitive flavin-tryptophan radical pairs.
2. Results
2.1. Spectroscopic Properties of ErCry4a Mutants
Spectroscopic studies were performed on wild type ErCry4a and four Trp → Phe mutants expressed and purified as described earlier.10 The structure of ErCry4a has not yet been determined due to the difficulty of obtaining crystals of sufficient quality. Gene sequencing46 and computational modeling10,46 suggest that it is similar to that of pigeon (Columba livia, Cl) Cry4, which also has four tryptophans involved in photo-activation, namely, TrpA (W395), TrpB (W372), TrpC (W318), and TrpD (W369).16 The structure of the FAD chromophore and the Trp tetrad, as obtained from such simulations, is shown in Figure 1e. Site-specific mutagenesis10 has been used to express four mutants, WXF (X = A, B, C, D), in which one of the four tryptophans (W) has been selectively replaced by phenylalanine (F) to block the electron transfer at different positions along the chain, as illustrated in Figure 1a–d.
Figure 1.
Visualization of the principal electron transfer route in ErCry4a and the effect of single amino acid mutations in blocking the electron transfer process at different stages. (a–d) Electron transfers in ErCry4a mutants involving Trp → Phe substitutions with W395 (TrpA), W372 (TrpB), W318 (TrpC), and W369 (TrpD). The arrows illustrate the electron transfer paths, while the crossed arrows represent the blocked electron transfers, as suggested by the experimental data in this study. (e) Edge-to-edge distances between the FAD chromophore and the components of the Trp-tetrad in wild type (WT) ErCry4a. (f) Normalized UV–visible absorption spectra of fully oxidized wild type ErCry4a and its four mutants. The pronounced vibronic structure on the 450 nm band indicates that the FAD is correctly bound to the protein. (g) Transient photoluminescence decay curves of wild type ErCry4a and its mutants compared to that of free FADox in buffer solution (solid lines) together with a mono-exponential, 4 ns decay (dashed line). Only the WAF mutant shows an initial fast component originating from the weak emission of protein-bound FAD (inset).
The UV–visible absorption spectra of wild type ErCry4a and its four mutants are very similar (Figure 1f), featuring a main peak centered at 450 nm with a well-resolved vibronic structure. Recent hybrid quantum/classical modeling,47 explicitly treating couplings to nuclear vibrations, have described the structure of these spectra and assigned the 450 nm band to a π → π* transition involving the π2 and π3 orbitals localized on the isoalloxazine moiety of the FAD cofactor.48 The vibronic structure is a distinct sign of chromophore binding to the protein and vanishes for free FAD in solution. The structured peak around 370 nm arises from the π1 → π3 transition.47
All five proteins showed weak photoluminescence (PL) with an unstructured emission around 550 nm (see Figure S1). The corresponding excitation spectra are similar and independent of the emission energy (Figure S1). In contrast to the absorption spectra (Figure 1f), the excitation spectra of all five proteins showed an absorption peak with no vibronic structure in the 450 nm band similar to that seen for free FAD in buffer solution (Figure S1). Comparison of the PL emission intensities of the proteins and of free FAD indicates that, in all of the protein samples, >97% of all FAD molecules are bound and that the weak residual PL seen in Figure S1 stems from a small amount of unbound FAD. This observation was confirmed by time-resolved PL studies (Figure 1g) showing, for all proteins except WAF, a 4-ns mono-exponential decay matching that of free FAD in buffer solution. The WAF mutant revealed an additional fast, resolution-limited emission component (Figure 1g, inset). In this mutant, TrpA, the closest tryptophan to the FAD, had been replaced by Phe, preventing the normally rapid electron transfer from TrpA to FAD. Therefore, the optically excited FADox* in WAF is longer-lived and its emission is not completely quenched by electron transfer. In all other mutants, the TrpA electron donor is present, resulting in efficient PL quenching and hence no emission from FADox. This observation already points to an efficient electron transfer from TrpA to FADox.
2.2. Transient Absorption Spectra of ErCry4a Proteins
Differential transmission (ΔT/T) spectra of the four Trp → Phe ErCry4a mutants were recorded under comparable experimental conditions at a temperature of 1 °C in buffer solution at pH 8. Linearly polarized pump pulses with 30-fs duration centered at 450 nm were used to excite the π2 → π3 transition of the FAD chromophore within the proteins. Pump-induced changes in the transmission of the sample were probed using a broadband white-light continuum. The magic angle ΔT/T spectra are shown in Figure 2a–d for the four mutants for pump–probe delay times, tW, of up to 1.5 ns. All datasets were subjected to a global data analysis and can readily be explained using a multi-exponential decay model, including a minimal set of decay-associated difference spectra (DADS).49 Each DADS spectrum (Figure S7) describes a component of ΔT/T that decays or rises with an associated time constant. As such spectra are sometimes somewhat difficult to interpret, we have also calculated evolution-associated difference spectra (EADS;49 see Methods). Here, EADS1 represents the ΔT/T spectrum before the onset of the first of the sequential incoherent relaxation steps, while in simplified terms, EADSn (n > 1) is the ΔT/T spectrum after the (n – 1)th relaxation process has been completed. The EADS spectra are presented in Figure 2e–h.
Figure 2.
Transient absorption spectroscopy of ErCry4a mutants. (a–d) Differential transmission ΔT/T spectra of the mutants WAF to WDF for delays between 0.2 ps and 1.5 ns (logarithmic time axis). (e–h) Corresponding evolution-associated difference spectra (EADS) for each mutant resulting from a global data analysis. (e) WAF mutant does not show a sub-picosecond decay component because the electron transfer from TrpA to FADox is suppressed by the removal of TrpA. (f–h) 0.5 ps decay components in the EADS spectra of WBF – WDF reflect this primary electron transfer from TrpA to FAD. Each increase in the distance of the Phe mutation from the chromophore (WBF → WCF → WDF) results in one additional decay component in the EADS spectra, giving evidence for the second and third electron transfer steps in the electron transfer chain in ErCry4a. A 2–4 ps decay component in all four mutants is assigned to vibrational cooling of the photochemically formed radical pairs. The slowest component for all four mutants corresponds to radical pair recombination on a nanosecond time scale.
2.2.1. Transient Absorption of ErCry4a WBF
We first consider the WBF mutant (Figure 2b,f), where the conceptually simplest, sub-picosecond electron transfer dynamics from TrpA to FAD are expected. The distinct signature of this electron transfer is the disappearance of the stimulated emission (SE) band around 550 nm, as shown in Figure 2b on a 0.5 ps time scale. The ΔT/T spectra in Figure 2b at early time delays, represented by the first EADS spectrum (EADS1) in Figure 2f, were recorded before this electron transfer sets in. These ΔT/T spectra thus show pump-induced changes in the spectrum of only the oxidized species, FADox. The spectra are therefore not yet affected by the formation of radicals in the sample. Indeed, EADS1 (blue line in Figure 2f) can be understood on the basis of the known optical transitions of FADox (Figure 3a). The pump excitation promotes electrons to vibrationally excited states of the π3 electronic state, giving rise to a positive (ΔT > 0) ground state bleaching (GSB) contribution to the differential transmission with a spectrum matching the absorption of FADox (blue line in Figure 3a). The pronounced vibronic structure around 450 nm is evident in EADS1 and is a clear marker for optically excited FADox*.29 Rapid vibrational relaxation within π3, complete within ≤ 200 fs, gives rise to a strongly red-shifted SE (ΔT > 0) band (π3 → π2) around 550 nm (see the red line in Figures 3a and S8). The GSB below 400 nm is obscured by a pronounced excited state absorption (ESA, ΔT < 0) centered at 360 nm (with a weak fine structure around 390 nm) from the π3 → π5 transition47 (red lines in Figure 3a,c). In addition, the ΔT/T spectra suggest broadband ESA of FADox, spanning the range from 500 to 700 nm and a spectrally almost flat ESA between 400 and 470 nm that is more difficult to assign because the absorption of FADox* has yet to be modeled with high accuracy (red lines in Figure 3a,c). Similar experimental ESA spectra have also been recorded by Kutta et al. but were not well-reproduced by their quantum chemical calculations.29
Figure 3.

Analysis of the differential transmission, ΔT/T, spectra of ErCry4a mutants. (a) EADS1 spectrum of WBF with 0.51 ps decay time (open circles) and its decomposition into the ΔT/T spectra of FADox and FADox* (solid lines). (b) EADS2 spectrum of WBF with a 1.8 ps decay time (open circles) and its decomposition into the ΔT/T spectra of FADox, FAD•–, and TrpAH•+ (solid lines). (c) Differential transmission ΔT/T spectra of the various neutral and charged radicals in ErCry4a as deduced from the analysis of WDF. (d) Population dynamics of the neutral and charged radical states in WDF.
Importantly, the overall shape of EADS1 in Figure 2e–h is very similar in all four ErCry4a mutants and the relative amplitudes of the different contributions remain basically unchanged. Since mutation alters the Trp chain with little effect on the FAD chromophore, this provides further support for the assignment of EADS1 to the ΔT/T spectrum of the FAD chromophore in its oxidized form. The most obvious dynamic feature of the ΔT/T map for the WBF mutant (Figure 2b) is the complete disappearance of the π3 → π2 SE band around 550 nm (decay time, 0.5 ps), while the amplitude of the GSB bands remain unchanged. This indicates that the decay of the excited state π3 population is not due to the refilling of the FADox ground state orbital, π2, but reflects the formation of the negatively charged FAD·– radical, most likely due to the transfer of an electron from TrpA. The map in Figure 2b also shows more subtle features. After the first electron transfer has been completed, the ΔT/T spectrum, now given by EADS2 in Figure 2f, has changed substantially. Not only has the characteristic SE band around 550 nm disappeared completely, but also the sharp peak in the FADox* absorption around 360 nm and the broad ESA tail (red line in Figure 2f) have vanished in EADS2. While the GSB and ESA around 450 nm do not change when switching from EADS1 to EADS2, a new narrow ESA band around 370 nm with a side peak at 400 nm and a broad ESA band in the 500–700 nm range emerge in EADS2. After subtracting the GSB contribution from EADS2, a ΔT/T spectrum is obtained that quantitatively matches the sum of the absorption spectra of the FAD•– radical anion and the TrpH•+ radical cation known from the literature.29,45 The corresponding spectra deduced from EADS2 are shown as solid lines in Figure 3b.The signatures of the spectrum of TrpH•+ (purple line in Figure 3c) are the broad absorption centered around 560 nm and the tail of a UV absorption band at 335 nm at the edge of the probe window.29,39,50 Theoretical studies of the FAD•– spectrum show a transition from the singly occupied π3 to the unoccupied π5 orbital,47 experimentally seen in the peak in Figure 3b (circles) at 370 nm, red-shifted by ca. 10 nm with respect to the corresponding transition in FADox. A weak peak at 400 nm in EADS2 is assigned47 to the transition π1 → π3 and a broad resonance between 400 and 470 nm (green line in Figure 3b) is assigned to π2 → π3. The decomposed spectra agree well with those reported for other types of cryptochrome, suggesting that the disappearance of FADox leads to the formation of the [FAD•–TrpAH•+] radical pair. The associated decay time of τ1 = 1/k1= 0.51 ps is therefore the time constant for the first electron transfer in ErCry4a activation, corresponding to the [FADox* TrpAH] → [FAD•– TrpAH•+] process. Since the characteristic markers for FADox*, i.e., the SE emission band around 550 nm and the ESA peak around 360 nm, are completely absent in EADS2, we conclude that the yield for this first electron transfer, η1, is close to unity.
Our data analysis reveals some dynamics associated with a second time constant, τv = 1.8 ps. This process, however, has only a minor effect on the ΔT/T spectra (Figure 2f). Close inspection reveals a slight narrowing of the TrpH•+ band around 560 nm and a minor reshaping of the FAD•– peaks at 370 and 400 nm. This has recently been associated with vibrational cooling of the optically formed radicals29 or, alternatively, to a second electron transfer along the Trp chain.45 The residual ΔT/T signal decays almost entirely with a time constant of τr = 89 ps. This is the decay of the photo-induced radical pairs by geminate recombination which refills the ground state, FADox. The residual, weak EADS4 spectrum in Figure 2f, with a shape matching that of EADS3, but with 20-fold smaller amplitude, suggests that a small concentration of radical pairs still exists for tW ≫ τr. Importantly, the ΔT/T spectra of WBF show no clear signatures of other electron transfer processes except for k1.
2.2.2. Transient Absorption of ErCry4a WAF
The substitution of TrpA by Phe in the WAF mutant is expected to block the first electron transfer. Initially, the ΔT/T spectra in Figure 2a (EADS1, blue line in Figure 2e) are very similar to those of WBF. The fast dynamics on the sub-picosecond time scale, however, are no longer observed, which directly confirms the assignment of the first electron transfer in the dynamics of WBF. The fastest electron transfer dynamics in WAF occur with a time constant of 3.3 ps and result (EADS2, red line in Figure 2e) in a partial reduction of the SE and ESA signals of FADox*. Since electron transfer processes involving TrpA cannot be the origin of this decay, the observed dynamics are most likely due to the FAD. Indeed, photo-induced electron transfer from the adenine to the flavin group in the FAD cofactor is known to occur in solution on a 5 ps time scale for the stacked conformation of the two aromatic rings.27,31 Although the exact time scale for the back electron transfer from flavin to adenine is somewhat uncertain for free FAD in solution,27,31 it is generally thought to happen on a time scale similar to the forward electron transfer, giving rise to an equilibrium between neutral and charge-separated states. The existence of such an equilibrium for the intra-FAD electron transfer in the WAF mutant may be the reason why the recorded data show only a partial decay of FADox with a 3.3 ps time constant. More surprising are the ΔT/T dynamics with a larger, 12 ps, time constant. After its completion, the resulting EADS3 (green line in Figure 2a) matches that seen in the WBF mutant, indicating the formation of a [FAD•–TrpXH•+] radical pair. MD simulations of the protein structure (Figure S9a) indicate that two additional Trp residues, Trp290 and Trp350, are present at edge-to-edge distances of 4 Å from the adenine and 7.5 Å from the flavin, respectively. These residues could potentially be involved in the formation of a [FAD•–TrpH•+] radical pair. The ΔT/T changes in Figure 2a disappear almost completely on a 100 ps time scale, and we associate this time constant with an effective geminate recombination of the [FAD•–TrpH•+] radical pair. The experiments prove the absence of fast electron transfer with sub-picosecond time constant in the WAF mutant.
2.2.3. Transient Absorption of ErCry4a WCF and WDF
The above analysis puts us in an excellent position to explore the electron transfer along the Trp chain in the ErCry4a WCF and WDF mutants (Figure 2c,d). In both cases, the ΔT/T spectra at early time delays (EADS1 in Figure 2g,h) are very similar to those discussed before for the WAF and WBF mutants. In the case of WCF and WDF, the initial dynamics point to a first electron transfer from TrpA to FAD with close to 100% efficiency and time constants of τ1 = 0.49 ps (WCF) and 0.4 ps (WDF), respectively. Also, the rapid decay of EADS2, seen for the WBF mutant, occurs with time constants of τv = 4.1 ps (WCF) and 2.2 ps (WDF). In contrast to the WBF case, the shape of the spectrum of the Trp radical cation changes slightly when going from EADS2 to EADS3, most obviously around 500 nm and in the far-red region. Since EADS2 is not affected by an extension of the Trp chain, our data do not support the assignment of τ2 to an electron transfer process. Instead, this observation presents strong arguments in favor of vibrational cooling as suggested earlier.29 The spectra recorded for both WCF and WDF mutants show a third EADS3 component with a spectral lineshape that is very similar to that observed for the WBF mutant. The EADS3 spectra decay with characteristic times of τ2 = 58 ps for WCF and 26 ps for WDF, respectively. In stark contrast to WBF, however, a fourth EADS4 component now appears after the third relaxation step with an amplitude approximately 2/3 that of EADS3, and with an almost identical lineshape. This finding implies that even after the third relaxation step, i.e. after about 100 ps, the excited ErCry4a still contains a sizeable concentration of radical pairs and that geminate recombination is not yet complete. This observation thus points to a secondary electron transfer step, [FAD•–TrpAH•+] →k2 [FAD•–TrpBH•+], in which an electron is transferred from TrpB to TrpA with a rate constant k2 = 1/τ2. The data in Figure 2g,h indicate that this secondary electron transfer occurs with a yield of η2 ≃ 0.77 for WCF and 0.71 for WDF. In the case of WCF, the resulting radical pairs then recombine very slowly with τr ≃ 1.6 ns. In WDF, both the GSB and the radical pair absorption decay more rapidly than for the other mutants. The analysis in Figure 2h reveals the presence of a new decay channel with a time constant τ3 = 144 ps. This relaxation process has little effect on the shape of the ΔT/T spectra: the spectra EADS3, EADS4 and EADS5 differ only in amplitude. We assign the observed decay to the third electron transfer step, [FAD•–TrpBH•+] → [FAD•–TrpCH•+], in which an electron moves from TrpC to TrpB with a rate constant k3 = 1/τ3 and a yield η3 ≃ 0.52. This yield is significantly lower than that of the second step. Figure 3d summarizes the population dynamics of the different neutral and charged radical states in WDF predicted by the analysis of the ΔT/T spectra. Only a small fraction, about 20–25%, of the initially generated radical pairs is still present after the first three electron transfers along the Trp chain (yellow line in Figure 3d). These radical pairs then undergo geminate recombination on a time scale τr = 2.0 ns (EADS5). Comparative studies of several samples of the same mutant indicate that the number of decay components needed to explain the experimental data does not depend on the specific sample studied. Also, the decay times and yields are quite similar.
2.2.4. Transient Absorption of Wild Type ErCry4a
We now compare the results in Figure 2 to those for wild type ErCry4a studied under the same experimental conditions. Qualitatively, the ΔT/T spectra in Figure 4a look almost identical to those recorded for the WDF mutant. Spectra at selected tW are shown in Figure 4b and the dynamics at different probe wavelengths are compared to the results of a global fit analysis in Figure 4c. As in the case of the WDF mutant, this analysis reveals five distinct EADS spectra with features that closely resemble those of the WDF mutant. Also, the decay times associated with those five components are almost the same as for WDF. This strongly indicates that the sequential electron transfer model developed from the spectra of the mutants can be used to explain the wild type data. As such, we assign the time constant τv = 3 ps to vibrational cooling. The decay time of EADS5 of τr = 2.2 ns is identified with a geminate recombination process.
Figure 4.

Transient absorption spectroscopy of wild type ErCry4a. (a) Differential transmission ΔT/T spectra of the wild type protein for delays been 0.2 ps and 1.5 ns (logarithmic time axis). (b) ΔT/T spectra at selected delays. (c) Dynamics of the ΔT/T spectra at selected wavelengths. The results of the global analysis are shown as black dashed lines. (d) EADS spectra as obtained from the global data analysis. The resulting spectra are very similar to those of the WDF mutant. Based on the analysis of the mutant spectra, we assign the components with decay times of 0.39, 30, and 141 ps to the first, second, and third sequential electron transfer along the tryptophan chain of the protein. The 3 ps decay component is associated with vibrational cooling of the optically generated radical pairs, while the 2.2 ns component reflects radical pair recombination.
Our experimental results provide clear evidence for three sequential electron transfers in wild type ErCry4a, with decay constants k1 = 1/0.39 ps–1, k2 = 1/30 ps–1, and k3 = 1/141 ps–1. The associated yields of the individual steps are η1 = 1.0, η2 = 0.78, and η3 = 0.77. In these measurements, the yield of the third step is somewhat higher than that for the WDF mutant. Thus, the concentration of radical pairs for the wild type protein is more than half the initial concentration of FADox*. The electron transfer rate constants and quantum yields deduced for the wild type protein and all mutants are summarized in Table 1. Importantly, our analysis of the transient absorption spectra of all four mutants gives no clear indication of back electron transfers within the FAD-Trp chain. All transients, except for the geminate recombination (τr) process, can be well understood using a simple rate equation model considering only the forward steps. We therefore conclude that the back transfers are much slower than the forward steps and in fact too slow to be revealed by our experiments.
Table 1. Experimentally Obtained Electron Transfer Time Constants and Yields for Mutant and Wild Type Forms of ErCry4a.
| protein | ET1 (η1) | ET2 (η2) | ET3 (η3) |
|---|---|---|---|
| WBF | 0.51 ps (100%) | ||
| WCF | 0.49 ps (100%) | 58 ps (77%) | |
| WDF | 0.40 ps (100%) | 26 ps (71%) | 144 ps (52%) |
| WT | 0.39 ps (100%) | 30 ps (78%) | 141 ps (77%) |
2.3. Quantum Mechanical/Molecular Mechanical Simulations
2.3.1. Occupancy of Tryptophan Sites
Our measured electron transfer time constants can now be compared to theoretical predictions from real-time quantum mechanical/molecular mechanical (QM/MM) simulations.41,51 The entire cryptochrome molecule is too large for a purely quantum-mechanical treatment, and therefore hybrid methods must be employed. Our simulations were performed for wild type ErCry4a to provide additional atomic-level insights into the nature of the charge separation process. This is the first attempt to carry out quantum simulations of electron transfer in ErCry4a for direct comparison with experimental measurements. Earlier phenomenological studies10 helped to estimate the kinetics of the charge transfer steps but did not take explicit account of the atomistic dynamics of the protein environment.
The electron transfer times within the Trp-tetrad are ca. 100 ps or faster (Table 1). A non-adiabatic, semi-classical electron transfer scheme based on a fragment-molecular-orbital implementation of the semi-empirical density-functional method DFTB41,51−54 is therefore suitable for computational modeling. For an accurate description of the electron transfer process, it is vital to take account of the motion of the biological scaffold of the protein. Moreover, this essentially classical motion should be coupled to the quantum dynamics of the transferring electron. The DFTB approach has been used previously to study real-time electron transfers in frog and plant cryptochromes41,51 and is now applied here to model the ET2–4 steps (see Figure 1a) in ErCry4a. The ErCry4a structure required for this study was obtained by homology modeling, equilibrated using extended MD simulations, and validated by several prior analyses.10,55,56 The starting point for simulating the ET2–4 steps is the [FAD•–TrpAH•+] radical pair, produced by the first electron transfer, ET1. Here, an electron vacancy (hole) resides on the TrpA residue and starts to propagate along the tryptophan chain, facilitated by the motion of the protein.41 As the hole progresses along the tetrad, one can calculate the probability that it is located on each of the tryptophans (the occupancies) and hence determine the electron transfer kinetics from their time dependence. Note that ET1 was not considered in our computational study because this step, which involves excited flavin and tryptophan, requires a special treatment within DFTB, while ET2–4 involve only tryptophan residues. See the Methods section for a description of the DFTB method.
The time dependence of the occupancies of the four tryptophan sites is shown in Figure 5a (averaged over 50 1-ns simulations) and Figure 5b (averaged over 27 630-ps simulations). By 630 ps, the electron hole had reached TrpCH•+ or TrpDH•+ with a probability of at least 0.3. Figure 5a includes the results of all trajectories, independent of whether the electron transfer reached the end of the Trp-tetrad. The solid lines in Figure 5a,b correspond to the numerical fits to the simulated data, which were obtained from the solution of the set of coupled rate equations
![]() |
1 |
Here, [Trpm] denotes the occupancy of the m-th tryptophan site and kn is the rate constant for the n-th electron transfer. This analysis gives estimates of the characteristic times of each electron transfer step; the results are summarized in Table 2. Based on our experimental results, we assume that all back transfer steps in the rate equation model are much slower than the corresponding forward rates and can therefore be neglected. Given the limited number of simulation runs currently available and the resulting statistical uncertainty, we cannot exclude the existence of small contributions from back electron transfer.
Figure 5.
Characterization of electron transfers (ET2-ET4) in the Trp-tetrad by QM/MM DFTB simulations. (a) Average occupancy of the four tryptophan sites computed over 50 simulations (points), fitted using eq 1 (lines). (b) Average occupancy of the four tryptophan sites (points) computed over 27 simulations in which electron transfer is considered complete, i.e., the occupancy of TrpCH•+ or TrpDH•+ exceeds 0.3. The solid lines show a fit to eq 1. (c–e) Average electronic couplings computed over 50 simulations between two neighboring tryptophan sites as a function of simulation time (dark color) with the corresponding standard deviation (light color). (f–h) Average site energies of the different tryptophans divided into three electron transfer scenarios observed in the simulations: (f) average site energies for the completed electron transfers (N = 27); see Video S1; (g) average site energies for the transfers that do not go beyond the first tryptophan TrpA (N = 13); see Video S2; and (h) average site energies for the transfers that stop at TrpB (N = 60); see Video S3.
Table 2. Estimated Electron Transfer Times Obtained from Simulations of Electron Propagation along the Trp-Tetrada.
Results are shown for the numerical fits of the solution of eq 1 to the computed data for the 1-ns simulation and 630-ps simulations, averaged over 50 and 27 independent simulations, respectively. In the latter case only those simulations where TrpCH·+ or TrpDH·+ obtained an occupancy of at least 0.3 were considered. The values have been rounded to the nearest 10 ps.
This analysis demonstrates the possibility of electron transfer between the tryptophans in ErCry4a. However, Figure 5a shows some discrepancies between the simulated data and the fitted curves at longer times—differences that are less pronounced in Figure 5b. The poor fitting of the simulation data in Figure 5a is due to the large uncertainties in the average occupancy values of the four TrpA–D sites. In a significant fraction of the simulations, electron transfer was not complete within the 1-ns time span. In a subset of the trajectories, the hole in TrpA rarely jumps to TrpB. This is a natural effect which arises because different initial structures support the ET process to different extents, and some do so rather inefficiently. A much larger number of simulations, associated with a prohibitive computational effort, would be required to take a proper account of this dependence. The current protocol is a practical solution that results in relatively minor overestimates of the rate constants of the individual ET steps. To illustrate the influence on the overall kinetics of those simulations in which the hole does not get as far as the end of the Trp-chain, we have included in Figure 5b data only for those simulations in which the occupancy of TrpCH•+ or TrpDH•+ reaches at least 0.3 by the end of the simulation. This gives a much better correspondence between the average occupancies of the tryptophans and the solutions of eq 1. The results in Figure 5b show the electron transfers TrpBH → TrpAH•+, followed by TrpCH → TrpBH•+, both on a 200 ps time scale. The fourth electron transfer, TrpDH → TrpCH•+, is predicted to be somewhat faster (<100 ps). Thus, the occupancy of TrpCH•+ will always be low, making it difficult to distinguish the occupancies of TrpCH•+ and TrpDH•+ in an experiment. This agrees well with our experimental observations for the wild type and mutant proteins. The electron transfer times derived from the simulations are presented in Table 2.
2.3.2. Marcus Theory Analysis
The results of the DFTB simulations can be qualitatively understood if we consider weak coupling between the donor and acceptor sites and assume the high temperature limit. In this case, the electron transfer rate constant, kET, can be estimated using Marcus theory,57 as:
| 2 |
kET is governed by three parameters: the coupling between the donor and acceptor, Tij (ij = AB, BC, CD); the thermodynamic driving force, ΔG; and the reorganization energy λ. The driving force is related to the site energies.41 The electronic coupling between two tryptophan sites is defined as Tij = ⟨ϕi|Ĥ0|ϕj⟩, while the site energies can be identified as the ionization potentials of the respective tryptophans as ϵi = ⟨ϕi|Ĥ0|ϕi⟩. Here, ϕi denotes the molecular orbitals of the electrons localized on the respective tryptophan residues, while Ĥ0 is the Hamiltonian of the system, including contributions from the solvent. Thus, Coulomb corrections to ΔG58 are intrinsically included in the DFTB simulations and, in fact, create a major part of the driving force for sequential electron transfer.41
Figure 5c–e shows the time evolution of the electronic couplings Tij of adjacent tryptophans. These interactions are the fundamental results of the DFTB calculations and have been extensively discussed previously.41,51 In particular, it has been shown that despite the simplicity of the DFTB approach, it yields accurate coupling values for charge transfer between tryptophan residues in cryptochromes and photolyases,41,51,59 comparable with values obtained using higher order methods. Physically, the electronic couplings determine the probability of electron transfer between the sites (eq 2). The average coupling of TrpB and TrpC (TBC) is around 5 meV, while that of TrpC and TrpD is significantly higher, at 20–30 meV.
The average site energies of the four tryptophans are shown in Figure 5f–h for three cases: simulations containing completed electron transfers (Figure 5f and Video S1), simulations with no electron transfer from TrpBH to TrpAH•+ (Figure 5g and Video S2), and simulations with no electron transfer from TrpCH to TrpBH•+ (Figure 5h and Video S3). The three outcomes correspond to very different site energies. Note that while the electronic couplings Tij contribute quadratically to the overall electron transfer rate constant (see eq 2), ΔG and λ (which are determined by the site energies) contribute exponentially and so may have more of an effect on kET. The average site energy difference between TrpA and TrpB is −0.05 eV for the cases with completed charge transfers; +0.62 eV when there is no transfer from TrpBH to TrpAH•+; and −0.56 eV for the case where the hole gets stuck at TrpBH•+. The average difference between TrpB and TrpC is +0.68 eV for the simulations in which the TrpCH → TrpBH•+ step does not occur. Thus, in Figure 5f, ET is energetically downhill all the way to TrpD; in Figure 5g, it is uphill from the start; and in Figure 5h, the first step is downhill, but subsequent electron transfers are energetically unfavorable.
3. Discussion
The ability to express high purity Cry4a from night-migratory European robins and to replace the tryptophan residues individually by redox-inactive phenylalanines has allowed us to obtain new microscopic insight into the light-induced electron transfer dynamics along the Trp-tetrad chain that connects the flavin chromophore to the surface of the protein. Using ultrafast optical spectroscopy, in connection with a Marcus model for the electron transport inside the protein, we have unraveled several of the primary steps in the photocycle of this photoreceptor, which has been suggested to be crucial for the magnetic compass sense of migratory songbirds.10,11,14
When the tryptophan (TrpA) closest to the FAD chromophore is replaced by a phenylalanine, electron transfer along the Trp-tetrad is efficiently suppressed and slower electron transfer pathways are revealed, most likely involving tryptophan residues that are not part of the tetrad. In all the other Trp → Phe mutants we have studied and in the wild type protein, TrpA serves as an efficient donor, transferring an electron to the FAD within 0.5 ps of photoexcitation with a quantum yield of unity. The experiments reveal mutation-insensitive vibrational cooling dynamics within the manifold of radical pair states on a timescale of few picoseconds. Thesy also provide evidence for a cascade of sequential electron transfers in which the second (TrpB → TrpA) and third (TrpC → TrpB) steps occur with transfer times of 30 and 140 ps, respectively, and can be selectively blocked by mutation.
The yields of the final step in generating the charge-separated states in the wild type protein and the WDF mutant are >75% and ∼50%, respectively. Our measurements, which probe the dynamics over the first 1.5 ns after photoexcitation, do not provide a clear indication for a fourth transfer step, from TrpD to TrpC, in wild type ErCry4a, either because of the difficulty of distinguishing these distant tryptophans or because this step is too fast or too slow to be resolved. A comparison of the transient absorption data from the wild type protein and the WDF mutant suggests that the presence of the fourth tryptophan in the chain has very little effect on the dynamics within the first 1.5 ns, beyond an increase in the concentration of radical pairs at the end of the measurement window. Such a change in concentration could be explained, for example, by the fourth electron transfer, from TrpD → TrpC, being faster than the third and the formation of an equilibrium between the associated radical pairs.10,18 The present data indeed point to the formation of such an equilibrium and suggest more refined studies of the functional role of the terminal Trp.
While the time constants we have deduced for the first and second electron transfers agree well with those recently reported by Kutta et al. for fruit fly cryptochrome,29 the kinetics measured here are somewhat slower than recent predictions10 based on MD simulations and Moser–Dutton theory. They are also faster than those predicted by the DFTB-based QM/MM simulations presented here. In the QM/MM simulations, the first ET1 could not be modeled and will remain an open question until reliable DFTB parameters become available for the excited flavin chromophore. The simulations give atomistic insight into the sequential electron transfer cascade along the Trp-tetrad with typical transfer times in the 100 ps range. Interestingly, the simulations suggest that for certain ErCry4a configurations, electron transfer does not proceed to the end of the tetrad but gets stuck at TrpAH•+. In such cases, the calculations suggest an increase in average site energy on going from TrpAH•+ to TrpBH•+ of 0.62 eV. In contrast, when this electron transfer does take place, it is because there is a decrease in the average site energy. The increases in site energy found for some of the simulations cannot be attributed to a single amino acid residue but are a collective effect of the protein environment that, overall, appears slightly different in the various starting configurations used in the QM/MM simulations.
The characteristic electron transfer times for the ET2 and ET3 processes obtained from the simulations in which the hole reaches the end of the Trp-tetrad agree, within an order-of-magnitude, with experiment. Comparing Tables 1 and 2, the agreement is better for ET3 than for ET2. Computationally, we could also establish a characteristic time for ET4 (60 ps), which has hitherto been inaccessible to experimental measurement. This step is somewhat faster than the preceding electron transfers, which would explain why it has been difficult to separate ET3 and ET4 experimentally. The theoretical electron transfer times follow from a simple kinetic model, eq 1, which was reasonably successful in describing the calculated average time dependence of the occupancies of the four tryptophan sites. The model did not include back electron transfer steps, which are expected to be significantly slower than the forward electron transfers as found in earlier studies40,41,51 and in the present measurements.
To minimize photodegradation of the light-sensitive proteins, our pump–probe experiments have all been performed at 1 °C and pH 8.0, under which conditions the samples are photostable for several hours. The conditions in a cell in a living bird are somewhat different: 40–43 °C and pH ∼7.3.60,61Thermodynamically, the main change in the charge states of proteins that results from a reduction in temperature is a shift in the acid–base equilibrium of histidine side chains, which can be reversed by an increase in pH. This suggests that, in terms of the protein structure, 1 °C and pH 8 should resemble the physiological conditions.61,62
Turning from thermodynamics to dynamics, the internal mobility of the protein will be strongly temperature-dependent. The effect this has on the electron transfer kinetics can be estimated from the Marcus model, as shown in eq 2, which suggests that the electron transfer times would decrease by only ca. 50% when the temperature is increased from 1 to 42 °C, the body temperature of a European robin. As such, we expect the main conclusions of the present work, which deals exclusively with the first nanosecond after photo-excitation, to remain valid. A preliminary study indeed suggests that the electron transfer rates and yields do not change greatly between 1 and 30 °C. This temperature dependence will be discussed in more detail in a forthcoming publication.
Finally, it should be noted that the magnetically sensitive steps in the photochemistry of cryptochromes all take place on timescales from hundreds of nanoseconds to seconds, minutes and possibly hours.10 Many of these processes are likely to be strongly temperature-dependent. Examples include the protonation and deprotonation of the flavin and tryptophan radical ions,63 electron spin relaxation,17 and the conformational changes that are necessary to form the state of the protein that initiates signal transduction.11 None of these processes is amenable to the ultrafast methods used here, and their temperature dependence will need to be investigated using other techniques.
4. Materials and Methods
4.1. Sample Preparation
Wild type ErCry4a (GenBank: KX890129.1) and its four mutants WXF (X = A–D) were cloned, expressed, and purified according to the protocol described by Xu et al.10 Briefly, the tryptophan mutants were generated by replacing the DNA codon for tryptophan (TGG) at amino acid position 395 (WAF), 372 (WBF), 318 (WCF), or 369 (WDF) in the ErCry4a gene by a phenylalanine codon (TTT) in a polymerase chain reaction using the Q5 site-directed mutagenesis kit (New England Biolabs). Plasmids were confirmed by Sanger sequencing (LGC Genomics). Proteins were expressed in BL21(DE3) E. coli cells in the dark and purified by Ni-NTA agarose columns, followed by anion exchange chromatography. Purified protein samples were concentrated to 5–6 mg/mL in an aqueous buffer solution (20 mM Tris, 250 mM NaCl, and 20% glycerol) along with 10 mM of the reducing agent 2-mercaptoethanol (BME) to avoid dimerization of the protein. Samples were snap-frozen in liquid nitrogen and stored at −80 °C for 3–5 days until the measurements were made. Since the photocycle starts from the fully oxidized state FADox, the reducing agent was removed and the sample was fully oxidized prior to the optical measurements. For this, the protein sample was washed with BME-free buffer solution in a Millipore centrifugation filter (Amicon Ultra, 30 kDa) using a temperature-controlled microcentrifuge (4 °C, 14,000 rpm). This step also removed free FAD from the sample. Addition of 1.5 mM potassium ferricyanide (PFC), followed by centrifugation for 1 h to remove aggregated proteins, was repeated until the sample was fully oxidized as confirmed by absorbance measurements. A remaining concentration of 1.7–2.6 mM PFC prevented photoreduction during the optical experiments. The PFC-corrected absorbance was used to determine the final concentration of the samples to be 120–220 μM using the molar extinction coefficient of FADox.31 All pump–probe experiments were performed at a temperature of 1 °C and in a buffer solution at pH 8.0 to avoid photodegradation of the light-sensitive protein. Transient absorption measurements on a buffer solution containing 1 mM PFC showed only a very weak nonlinear signal with a lifetime of a few picoseconds when pumped at 450 nm (see Figure S4), and no detectable PL was observed.
4.2. PL Experiments
PL measurements were performed using broadband (400–680 nm) 1.25 pJ pulses from a 40 MHz fiber laser source (SC400, Fianium). Emitted light was collected with a microscope objective with a numerical aperture of 0.3 and detected using a single-photon avalanche photodiode and a time-correlated single-photon counting unit (PicoHarp 300, PicoQuant). The instrument response time of the detection system at 550 nm is ∼60 ps. To obtain correlated excitation and emission spectra, a Fourier transform approach using two phase-stable common-path interferometers (Translating Wedge-based Identical pulse eNcoding System, TWINS64) was used,65 yielding a spectral resolution of ∼6 nm in excitation and emission wavelength. A more detailed description of the setup and data analysis can be found in the Supporting Information.
4.3. Transient Absorption Experiments
The measurements were performed using ∼30 fs pump pulses at 450 nm on a 15 μL volume of the sample in a quartz microcuvette (Hellma). The pulses are generated in an optical parametric amplifier (Topas, Light Conversion) pumped by regeneratively amplified 25-fs pulses at 800 nm (Legend Elite, Coherent) at a repetition rate of 10 kHz. Differential transmission spectra ΔT(λ, tW)/T = (Ton(λ, tW) – Toff(λ))/Toff(λ) were recorded as a function of probe wavelength λ and time delay tW using a broad supercontinuum probe, generated in a CaF2 crystal. Here, Ton/off denotes the probe transmission in the presence/absence of the pump, respectively. The spectra were recorded with parallel and crossed polarizations of the pump and probe beams. Scattering corrections to the ΔT spectra were made as described in the Supporting Information. The experiments were performed with 20 nJ pump pulses focused to a spot size of ∼50 μm, and it was ensured that the ΔT/T signals are well in the linear regime (see Figure S2). No change in signal due to sample degradation could be observed during the measurements (see Figure S3).
4.4. Data Analysis
Transient absorption scans recorded with parallel and crossed pump and probe polarizations were averaged, and isotropic, magic angle spectra were calculated. The few-ps chirp of the probe continuum was corrected by extracting the wavelength-dependent time delay zero (tW(λ) = 0) from the cross-phase modulation artifact of a transient absorption measurement of plain buffer solution. The first 200 fs of the corrected dynamics were discarded due to residual coherent signal contributions from the solvent. The datasets were subjected to a global analysis using a multi-exponential decay model.66 This decomposes the data into a set of n decay associated difference spectra (DADSi) with corresponding decay times τi, ΔT/T(λ, tW) = ∑ni = 1DADSi(λ)e–tW/τi. The lowest number of decays necessary to simultaneously reproduce the data at all wavelengths was taken. For more details, see the Supporting Information. The DADS spectra were then used to obtain evolution associated difference spectra EADSk(λ) = ∑ni = kDADSi(λ).
4.4.1. MD Simulations
The structure of ErCry4a in its dark state was adapted from earlier studies,10,46,56 where it was simulated in a 94 Å × 106 Å × 102 Å water box neutralized with 0.15 M NaCl, resulting in a total of 100,518 atoms. The structure was minimized for 500 conjugate gradient steps using a steepest descent minimization algorithm, followed by a second, 2 ns equilibration using the leap-frog integrator with a temperature of 300 K, kept constant by a Berendsen thermostat. The equilibration was followed by a 100 ns dynamic equilibration and 100 ns production simulation. Both equilibration and production simulations utilized a 2-fs timestep, and the LINCS algorithm67 was used to keep the lengths of bonds involving hydrogen atoms fixed at their equilibrium values. Periodic boundary conditions were adopted for all stages, and the particle-mesh-Ewald summation method was employed for evaluating the Coulomb forces. Van der Waals forces were calculated using a smooth cut-off of 12 Å with a switching distance of 10 Å. All calculations were carried out utilizing the GROMACS package.52,68 The MD calculation utilized the Amber99SB forcefields for proteins69,70 with earlier parametrized forcefields for the FAD cofactor.41,51
4.5. Non-adiabatic Simulation of Electron Transfer
The method for the real-time electron transfer has been carried out using a hybrid QM/MM DFTB protocol from two earlier studies on frog and plant cryptochromes.41,51 The protocol employed molecular fragments participating in the electron transfer as a defined segment of the system that was subjected to the quantum mechanical description (QM region), with the rest of the system being described using classical molecular mechanics force fields (MM region). The computational scheme utilizes a separation of frontier orbitals similar to the Hückel and Pariser-Parr-Pople models71,72 although in this case the orbitals were based on molecular fragments instead of atoms. For the DFTB calculations, a total of 100 structures (snapshots) were sampled from the production MD simulation, and each of the snapshots was taken at a 1-ns time interval. The QM region was for all QM/MM simulations selected as the four tryptophans. In this region, TrpAH•+ was assumed to have a missing electron, as the initial structure and FAD was assumed to have an extra electron, that is, being in its FAD•– state. The snapshots were each simulated for 1 ns using a time step of 1 fs, though due to computational limitations, only 50 of the snapshots completed the 1 ns calculation, while the other 50 calculations resulted in a simulation length of 630 ps. The MM region was simulated using the Amber99SB forcefields for proteins69,70 with earlier parametrized forcefields for the FAD cofactor,41,51 while the QM region and electron transfer was estimated using earlier parameters.73,74
Acknowledgments
Financial support by the Deutsche Forschungsgemeinschaft (project number 395940726- SFB1372: Magnetoreception and Navigation in Vertebrates: From biophysics to brain and behavior, GRK1885 “Molecular basis of sensory biology”, INST 184/163-1 FUGG, Li 580/16-1 and DE 3578/3-1), by the European Research Council (under the European Union’s Horizon 2020 research and innovation program, grant agreement no. 810002 (Synergy Grant: “QuantumBirds”)) and by the Volkswagen Foundation (SMART; Lichtenberg professorship) is gratefully acknowledged. Computational resources for the simulations were provided by the CARL Cluster at the Carl-von-Ossietzky University, Oldenburg, supported by the DFG and the Ministry for Science and Culture of Lower Saxony. This work was also supported by the North-German Supercomputing Alliance (HLRN). We thank P.J. Hore and K.B. Henbest for fruitful discussions. We express our gratitude to P.J. Hore for a critical reading of our manuscript.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c00442.
Additional experimental details, materials and methods, PL spectra, effect of pump pulse energy, stability of differential transmission spectra, reference measurements, DADS spectra, and spectral assignment (PDF)
Simulations for the completed electron transfer (AVI)
Simulations with no electron transfer from TrpBH to TrpAH·+ (AVI)
Simulations with no electron transfer from TrpCH to TrpBH·+ (AVI)
The authors declare no competing financial interest.
Supplementary Material
References
- Chaves I.; Pokorny R.; Byrdin M.; Hoang N.; Ritz T.; Brettel K.; Essen L. O.; van der Horst G. T. J.; Batschauer A.; Ahmad M. The Cryptochromes: Blue Light Photoreceptors in Plants and Animals. Annu. Rev. Plant Biol. 2011, 62, 335–364. 10.1146/annurev-arplant-042110-103759. [DOI] [PubMed] [Google Scholar]
- Griffin E. A.; Staknis D.; Weitz C. J. Light-Independent Role of CRY1 and CRY2 in the Mammalian Circadian Clock. Science 1999, 286, 768–771. 10.1126/science.286.5440.768. [DOI] [PubMed] [Google Scholar]
- Hsu D. S.; Zhao X.; Zhao S.; Kazantsev A.; Wang R.-P.; Todo T.; Wei Y.-F.; Sancar A. Putative Human Blue-Light Photoreceptors hCRY1 and hCRY2 Are Flavoproteins. Biochemistry 1996, 35, 13871–13877. 10.1021/bi962209o. [DOI] [PubMed] [Google Scholar]
- Zhu H.; Sauman I.; Yuan Q.; Casselman A.; Emery-Le M.; Emery P.; Reppert S. M. Cryptochromes Define a Novel Circadian Clock Mechanism in Monarch Butterflies That May Underlie Sun Compass Navigation. PLoS Biol. 2008, 6, e4 10.1371/journal.pbio.0060004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stanewsky R.; Kaneko M.; Emery P.; Beretta B.; Wager-Smith K.; Kay S. A.; Rosbash M.; Hall J. C. The cry(b) mutation identifies cryptochrome as a circadian photoreceptor in Drosophila. Cell 1998, 95, 681–692. 10.1016/S0092-8674(00)81638-4. [DOI] [PubMed] [Google Scholar]
- van der Horst G. T. J.; Muijtjens M.; Kobayashi K.; Takano R.; Kanno S.; Takao M.; de Wit J.; Verkerk A.; Eker A. P. M.; van Leenen D.; Buijs R.; Bootsma D.; Hoeijmakers J. H. J.; Yasui A. Mammalian Cry1 and Cry2 are essential for maintenance of circadian rhythms. Nature 1999, 398, 627–630. 10.1038/19323. [DOI] [PubMed] [Google Scholar]
- Somers D. E.; Devlin P. F.; Kay S. A. Phytochromes and cryptochromes in the entrainment of the Arabidopsis circadian clock. Science 1998, 282, 1488–1490. 10.1126/science.282.5393.1488. [DOI] [PubMed] [Google Scholar]
- Lin C. T.; Yang H. Y.; Guo H. W.; Mockler T.; Chen J.; Cashmore A. R. Enhancement of blue-light sensitivity of Arabidopsis seedlings by a blue light receptor cryptochrome 2. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 2686–2690. 10.1073/pnas.95.5.2686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo H. W.; Yang W. Y.; Mockler T. C.; Lin C. T. Regulations of flowering time by Arabidopsis photoreceptors. Science 1998, 279, 1360–1363. 10.1126/science.279.5355.1360. [DOI] [PubMed] [Google Scholar]
- Xu J. J.; Jarocha L. E.; Zollitsch T.; Konowalczyk M.; Henbest K. B.; Richert S.; Golesworthy M. J.; Schmidt J.; Dejean V.; Sowood D. J. C.; Bassetto M.; Luo J. T.; Walton J. R.; Fleming J.; Wei Y. J.; Pitcher T. L.; Moise G.; Herrmann M.; Yin H.; Wu H. J.; Bartolke R.; Kasehagen S. J.; Horst S.; Dautaj G.; Murton P. D. F.; Gehrckens A. S.; Chelliah Y.; Takahashi J. S.; Koch K. W.; Weber S.; Solov’yov I. A.; Xie C.; Mackenzie S. R.; Timmel C. R.; Mouritsen H.; Hore P. J. Magnetic sensitivity of cryptochrome 4 from a migratory songbird. Nature 2021, 594, 535–540. 10.1038/s41586-021-03618-9. [DOI] [PubMed] [Google Scholar]
- Hore P. J.; Mouritsen H. The Radical-Pair Mechanism of Magnetoreception. Annu. Rev. Biophys. 2016, 45, 299–344. 10.1146/annurev-biophys-032116-094545. [DOI] [PubMed] [Google Scholar]
- Solovyov I. A.; Chandler D. E.; Schulten K. Magnetic Field Effects in Arabidopsis thaliana Cryptochrome-1. Biophys. J. 2007, 92, 2711–2726. 10.1529/biophysj.106.097139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pinzon-Rodriguez A.; Bensch S.; Muheim R. Expression patterns of cryptochrome genes in avian retina suggest involvement of Cry4 in light-dependent magnetoreception. J. R. Soc. Interface 2018, 15, 20180058 10.1098/rsif.2018.0058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mouritsen H. Long-distance navigation and magnetoreception in migratory animals. Nature 2018, 558, 50–59. 10.1038/s41586-018-0176-1. [DOI] [PubMed] [Google Scholar]
- Kutta R. J.; Archipowa N.; Johannissen L. O.; Jones A. R.; Scrutton N. S. Vertebrate Cryptochromes are Vestigial Flavoproteins. Sci. Rep. 2017, 7, 44906. 10.1038/srep44906. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zoltowski B. D.; Chelliah Y.; Wickramaratne A.; Jarocha L.; Karki N.; Xu W.; Mouritsen H.; Hore P. J.; Hibbs R. E.; Green C. B.; Takahashi J. S. Chemical and structural analysis of a photoactive vertebrate cryptochrome from pigeon. Proc. Natl. Acad. Sci. U. S. A. 2019, 116, 19449–19457. 10.1073/pnas.1907875116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kattnig D. R.; Sowa J. K.; Solov’yov I. A.; Hore P. J. Electron spin relaxation can enhance the performance of a cryptochrome-based magnetic compass sensor. New J. Phys. 2016, 18, 063007 10.1088/1367-2630/18/6/063007. [DOI] [PubMed] [Google Scholar]
- Wong S. Y.; Wei Y. J.; Mouritsen H.; Solov’yov I. A.; Hore P. J. Cryptochrome magnetoreception: four tryptophans could be better than three. J. R. Soc. Interface 2021, 18, 20210601 10.1098/rsif.2021.0601. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kao Y. T.; Tan C.; Song S. H.; Ozturk N.; Li J.; Wang L. J.; Sancar A.; Zhong D. P. Ultrafast dynamics and anionic active states of the flavin cofactor in cryptochrome and photolyase. J. Am. Chem. Soc. 2008, 130, 7695–7701. 10.1021/ja801152h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lukacs A.; Eker A. P. M.; Byrdin M.; Brettel K.; Vos M. H. Electron Hopping through the 15 angstrom Triple Tryptophan Molecular Wire in DNA Photolyase Occurs within 30 ps. J. Am. Chem. Soc. 2008, 130, 14394–14395. 10.1021/ja805261m. [DOI] [PubMed] [Google Scholar]
- Paulus B.; Bajzath C.; Melin F.; Heidinger L.; Kromm V.; Herkersdorf C.; Benz U.; Mann L.; Stehle P.; Hellwig P.; Weber S.; Schleicher E. Spectroscopic characterization of radicals and radical pairs in fruit fly cryptochrome-protonated and nonprotonated flavin radical-states. FEBS J. 2015, 282, 3175–3189. 10.1111/febs.13299. [DOI] [PubMed] [Google Scholar]
- Schulten K.; Swenberg C. E.; Weller A. Biomagnetic sensory mechanism based on magnetic-field modulated coherent electron-spin motion. Z. Phys. Chem. 1978, 111, 1–5. 10.1524/zpch.1978.111.1.001. [DOI] [Google Scholar]
- Ritz T.; Adem S.; Schulten K. A model for photoreceptor-based magnetoreception in birds. Biophys. J. 2000, 78, 707–718. 10.1016/S0006-3495(00)76629-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Solov’yov I. A.; Chandler D. E.; Schulten K. Exploring the possibilities for radical pair effects in cryptochrome. Plant Signaling Behav. 2008, 3, 676–677. 10.4161/psb.3.9.5809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Solov’yov I. A.; Mouritsen H.; Schulten K. Acuity of a Cryptochrome and Vision-Based Magnetoreception System in Birds. Biophys. J. 2010, 99, 40–49. 10.1016/j.bpj.2010.03.053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhong D. P.; Zewail A. H. Femtosecond dynamics of flavoproteins: Charge separation and recombination in riboflavine (vitamin B-2)-binding protein and in glucose oxidase enzyme. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 11867–11872. 10.1073/pnas.211440398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kao Y. T.; Saxena C.; He T. F.; Guo L. J.; Wang L. J.; Sancar A.; Zhong D. P. Ultrafast dynamics of flavins in five redox states. J. Am. Chem. Soc. 2008, 130, 13132–13139. 10.1021/ja8045469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Immeln D.; Weigel A.; Kottke T.; Lustres J. L. P. Primary Events in the Blue Light Sensor Plant Cryptochrome: Intraprotein Electron and Proton Transfer Revealed by Femtosecond Spectroscopy. J. Am. Chem. Soc. 2012, 134, 12536–12546. 10.1021/ja302121z. [DOI] [PubMed] [Google Scholar]
- Kutta R. J.; Archipowa N.; Scrutton N. S. The sacrificial inactivation of the blue-light photosensor cryptochrome from Drosophila melanogaster. Phys. Chem. Chem. Phys. 2018, 20, 28767–28776. 10.1039/C8CP04671A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chosrowjan H.; Taniguchi S.; Mataga N.; Tanaka F.; Visser A. J. W. G. The stacked flavin adenine dinucleotide conformation in water is fluorescent on picosecond timescale. Chem. Phys. Lett. 2003, 378, 354–358. 10.1016/S0009-2614(03)01339-3. [DOI] [Google Scholar]
- Brazard J.; Usman A.; Lacombat F.; Ley C.; Martin M. M.; Plaza P. New Insights into the Ultrafast Photophysics of Oxidized and Reduced FAD ins Solution. J. Phys. Chem. A 2011, 115, 3251–3262. 10.1021/jp110741y. [DOI] [PubMed] [Google Scholar]
- Sancar A. Structure and Function of DNA Photolyase and Cryptochrome Blue-Light Photoreceptors. Chem. Rev. 2003, 103, 2203–2238. 10.1021/cr0204348. [DOI] [PubMed] [Google Scholar]
- Brazard J.; Ley C.; Lacombat F.; Plaza P.; Martin M. M.; Checcucci G.; Lenci F. Primary Photoprocesses Involved in the Sensory Protein for the Photophobic Response of Blepharisma japonicum. J. Phys. Chem. B 2008, 112, 15182–15194. 10.1021/jp805815e. [DOI] [PubMed] [Google Scholar]
- Bouly J.-P.; Schleicher E.; Dionisio-Sese M.; Vandenbussche F.; Van Der Straeten D.; Bakrim N.; Meier S.; Batschauer A.; Galland P.; Bittl R.; Ahmad M. Cryptochrome Blue Light Photoreceptors Are Activated through Interconversion of Flavin Redox States *. J. Biol. Chem. 2007, 282, 9383–9391. 10.1074/jbc.M609842200. [DOI] [PubMed] [Google Scholar]
- Yoneda Y.; Kudisch B.; Rafiq S.; Maiuri M.; Nagasawa Y.; Scholes G. D.; Miyasaka H. Vibrational Dephasing along the Reaction Coordinate of an Electron Transfer Reaction. J. Am. Chem. Soc. 2021, 143, 14511. 10.1021/jacs.1c01863. [DOI] [PubMed] [Google Scholar]
- Kudisch B.; Oblinsky D. G.; Black M. J.; Zieleniewska A.; Emmanuel M. A.; Rumbles G.; Hyster T. K.; Scholes G. D. Active-Site Environmental Factors Customize the Photophysics of Photoenzymatic Old Yellow Enzymes. J. Phys. Chem. B 2020, 124, 11236–11249. 10.1021/acs.jpcb.0c09523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shirdel J.; Zirak P.; Penzkofer A.; Breitkreuz H.; Wolf E. Absorption and fluorescence spectroscopic characterisation of the circadian blue-light photoreceptor cryptochrome from Drosophila melanogaster (dCry). Chem. Phys. 2008, 352, 35–47. 10.1016/j.chemphys.2008.06.006. [DOI] [Google Scholar]
- Liu Z. Y.; Tan C.; Guo X. M.; Li J.; Wang L. J.; Sancar A.; Zhong D. P. Determining complete electron flow in the cofactor photoreduction of oxidized photolyase. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 12966–12971. 10.1073/pnas.1311073110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martin R.; Lacombat F.; Espagne A.; Dozova N.; Plaza P.; Yamamoto J.; Müller P.; Brettel K.; de la Lande A. Ultrafast flavin photoreduction in an oxidized animal (6-4) photolyase through an unconventional tryptophan tetrad. Phys. Chem. Chem. Phys. 2017, 19, 24493–24504. 10.1039/C7CP04555G. [DOI] [PubMed] [Google Scholar]
- Solovyov I. A.; Domratcheva T.; Moughal Shahi A. R.; Schulten K. Decrypting Cryptochrome: Revealing the Molecular Identity of the Photoactivation Reaction. J. Am. Chem. Soc. 2012, 134, 18046–18052. 10.1021/ja3074819. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lüdemann G.; Solovyov I. A.; Kubař T.; Elstner M. Solvent Driving Force Ensures Fast Formation of a Persistent and Well-Separated Radical Pair in Plant Cryptochrome. J. Am. Chem. Soc. 2015, 137, 1147–1156. 10.1021/ja510550g. [DOI] [PubMed] [Google Scholar]
- Qin S. Y.; Yin H.; Yang C. L.; Dou Y. F.; Liu Z. M.; Zhang P.; Yu H.; Huang Y. L.; Feng J.; Hao J. F.; Hao J.; Deng L. Z.; Yan X. Y.; Dong X. L.; Zhao Z. X.; Jiang T. J.; Wang H. W.; Luo S. J.; Xie C. A magnetic protein biocompass. Nat. Mater. 2016, 15, 217–226. 10.1038/nmat4484. [DOI] [PubMed] [Google Scholar]
- Mims D.; Herpich J.; Lukzen N. N.; Steiner U. E.; Lambert C. Readout of spin quantum beats in a charge-separated radical pair by pump-push spectroscopy. Science 2021, 374, 1470–1474. 10.1126/science.abl4254. [DOI] [PubMed] [Google Scholar]
- Müller P.; Yamamoto J.; Martin R.; Iwai S.; Brettel K. Discovery and functional analysis of a 4th electron-transferring tryptophan conserved exclusively in animal cryptochromes and (6-4) photolyases. Chem. Commun. 2015, 51, 15502–15505. 10.1039/C5CC06276D. [DOI] [PubMed] [Google Scholar]
- Lacombat F.; Espagne A.; Dozova N.; Plaza P.; Müller P.; Brettel K.; Franz-Badur S.; Essen L. O. Ultrafast Oxidation of a Tyrosine by Proton-Coupled Electron Transfer Promotes Light Activation of an Animal-like Cryptochrome. J. Am. Chem. Soc. 2019, 141, 13394–13409. 10.1021/jacs.9b03680. [DOI] [PubMed] [Google Scholar]
- Günther A.; Einwich A.; Sjulstok E.; Feederle R.; Bolte P.; Koch K. W.; Solov’yov I. A.; Mouritsen H. Double-Cone Localization and Seasonal Expression Pattern Suggest a Role in Magnetoreception for European Robin Cryptochrome 4. Curr. Biol. 2018, 28, 211–223.e4. 10.1016/j.cub.2017.12.003. [DOI] [PubMed] [Google Scholar]
- Schwinn K.; Ferre N.; Huix-Rotllant M. UV-visible absorption spectrum of FAD and its reduced forms embedded in a cryptochrome protein. Phys. Chem. Chem. Phys. 2020, 22, 12447–12455. 10.1039/D0CP01714K. [DOI] [PubMed] [Google Scholar]
- Mondal P.; Schwinn K.; Huix-Rotllant M. Impact of the redox state of flavin chromophores on the UV-vis spectra, redox and acidity constants and electron affinities. J. Photochem. Photobiol., A 2020, 387, 112164 10.1016/j.jphotochem.2019.112164. [DOI] [Google Scholar]
- van Stokkum I. H. M.; Larsen D. S.; van Grondelle R. Global and target analysis of time-resolved spectra. Biochim. Biophys. Acta, Bioenerg. 2004, 1657, 82–104. 10.1016/j.bbabio.2004.04.011. [DOI] [PubMed] [Google Scholar]
- Solar S.; Getoff N.; Surdhar P. S.; Armstrong D. A.; Singh A. Oxidation of Tryptophan and N-Methylindole by N3., Br2-, and (Scn)2- Radicals in Light-Water and Heavy-Water Solutions - a Pulse-Radiolysis Study. J. Phys. Chem. 1991, 95, 3639–3643. 10.1021/j100162a038. [DOI] [Google Scholar]
- Sjulstok E.; Lüdemann G.; Kubar T.; Elstner M.; Solov’yov I. A. Molecular Insights into Variable Electron Transfer in Amphibian Cryptochrome. Biophys. J. 2018, 114, 2563–2572. 10.1016/j.bpj.2018.04.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hess B.; Kutzner C.; van der Spoel D.; Lindahl E. GROMACS 4: Algorithms for highly efficient, load-balanced, and scalable molecular simulation. J. Chem. Theory Comput. 2008, 4, 435–447. 10.1021/ct700301q. [DOI] [PubMed] [Google Scholar]
- Kubar T.; Elstner M. A hybrid approach to simulation of electron transfer in complex molecular systems. J. R. Soc. Interface 2013, 10, 20130415 10.1098/rsif.2013.0415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kubar T.; Elstner M. What governs the charge transfer in DNA? The role of DNA conformation and environment. J. Phys. Chem. B 2008, 112, 8788–8798. 10.1021/jp803661f. [DOI] [PubMed] [Google Scholar]
- Hanić M.; Frederiksen A.; Schuhmann F.; Solov’yov I. A. On the energetic differences of avian cryptochromes 4 from selected species. Eur. Phys. J. D 2022, 76, 198. 10.1140/epjd/s10053-022-00520-5. [DOI] [Google Scholar]
- Hanić M.; Schuhmann F.; Frederiksen A.; Langebrake C.; Manthey G.; Liedvogel M.; Xu J. J.; Mouritsen H.; Solov’yov I. A. Computational Reconstruction and Analysis of Structural Models of Avian Cryptochrome 4. J. Phys. Chem. B 2022, 126, 4623–4635. 10.1021/acs.jpcb.2c00878. [DOI] [PubMed] [Google Scholar]
- Marcus R. A.; Sutin N. Electron Transfers in Chemistry and Biology. Biochim. Biophys. Acta 1985, 811, 265–322. 10.1016/0304-4173(85)90014-X. [DOI] [Google Scholar]
- Rehm D.; Weller A. Kinetics and mechanics of electron transfer during fluorescence quenching in acetonitrile. Ber. Bunsenges. Phys. Chem. 1969, 73, 834. [Google Scholar]
- Holub D.; Ma H. J.; Krauss N.; Lamparter T.; Elstner M.; Gillet N. Functional role of an unusual tyrosine residue in the electron transfer chain of a prokaryotic (6-4) photolyase. Chem. Sci. 2018, 9, 1259–1272. 10.1039/C7SC03386A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roos A.; Boron W. F. Intracellular pH. Physiol. Rev. 1981, 61, 296–434. 10.1152/physrev.1981.61.2.296. [DOI] [PubMed] [Google Scholar]
- Reeves R. B. Interaction of body-temperature and acid-base-balance in ectothermic vertebrates. Annu. Rev. Physiol. 1977, 39, 559–586. 10.1146/annurev.ph.39.030177.003015. [DOI] [PubMed] [Google Scholar]
- Weber S.; Biskup T.; Okafuji A.; Marino A. R.; Berthold T.; Link G.; Hitomi K.; Getzoff E. D.; Schleicher E.; Norris J. R. Jr. Origin of Light-Induced Spin-Correlated Radical Pairs in Cryptochrome. J. Phys. Chem. B 2010, 114, 14745–14754. 10.1021/jp103401u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maeda K.; Robinson A. J.; Henbest K. B.; Hogben H. J.; Biskup T.; Ahmad M.; Schleicher E.; Weber S.; Timmel C. R.; Hore P. J. Magnetically sensitive light-induced reactions in cryptochrome are consistent with its proposed role as a magnetoreceptor. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 4774–4779. 10.1073/pnas.1118959109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brida D.; Manzoni C.; Cerullo G. Phase-locked pulses for two-dimensional spectroscopy by a birefringent delay line. Opt. Lett. 2012, 37, 3027–3029. 10.1364/OL.37.003027. [DOI] [PubMed] [Google Scholar]
- Lünemann D. C.; Thomas A. R.; Xu J. J.; Bartolke R.; Mouritsen H.; De Sio A.; Lienau C. Distinguishing between coherent and incoherent signals in excitation-emission spectroscopy. Opt. Express 2021, 29, 24326–24337. 10.1364/OE.428850. [DOI] [PubMed] [Google Scholar]
- Rabe M. Spectram: A MATLAB® and GNU Octave Toolbox for Transition Model Guided Deconvolution of Dynamic Spectroscopic Data. J. Open Res. Softw. 2020, 8, 13. 10.5334/jors.323. [DOI] [Google Scholar]
- Hess B.; Bekker H.; Berendsen H. J. C.; Fraaije J. G. E. M. LINCS: A linear constraint solver for molecular simulations. J. Comput. Chem. 1997, 18, 1463–1472. . [DOI] [Google Scholar]
- Bauer P.; Hess B.; Lindahl E. GROMACS 2022.3 Manual. Zenodo 2022, 2022, 3. [Google Scholar]
- Wang J. M.; Wolf R. M.; Caldwell J. W.; Kollman P. A.; Case D. A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 1157–1174. 10.1002/jcc.20035. [DOI] [PubMed] [Google Scholar]
- Lindorff-Larsen K.; Piana S.; Palmo K.; Maragakis P.; Klepeis J. L.; Dror R. O.; Shaw D. E. Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins: Struct., Funct., Bioinf. 2010, 78, 1950–1958. 10.1002/prot.22711. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pariser R.; Parr R. G. A Semi-Empirical Theory of the Electronic Spectra and Electronic Structure of Complex Unsaturated Molecules. II. J. Chem. Phys. 1953, 21, 767–776. 10.1063/1.1699030. [DOI] [Google Scholar]
- Pople J. A. Electron Interaction in Unsaturated Hydrocarbons. Trans. Faraday Soc. 1953, 49, 1375–1385. 10.1039/tf9534901375. [DOI] [Google Scholar]
- Kubar T.; Elstner M. Coarse-Grained Time-Dependent Density Functional Simulation of Charge Transfer in Complex Systems: Application to Hole Transfer in DNA. J. Phys. Chem. B 2010, 114, 11221–11240. 10.1021/jp102814p. [DOI] [PubMed] [Google Scholar]
- Kubar T.; Elstner M. Efficient algorithms for the simulation of non-adiabatic electron transfer in complex molecular systems: application to DNA. Phys. Chem. Chem. Phys. 2013, 15, 5794–5813. 10.1039/c3cp44619k. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




