Chey M.
Jones
ab,
Nanna H.
List
ab and
Todd J.
Martínez
*ab
aDepartment of Chemistry and the PULSE Institute, Stanford University, Stanford, CA 94305, USA. E-mail: toddjmartinez@gmail.com
bSLAC National Accelerator Laboratory, 2575 Sand Hill Road, Menlo Park, CA 94025, USA
First published on 13th July 2021
The chromophore of the green fluorescent protein (GFP) is critical for probing environmental influences on fluorescent protein behavior. Using the aqueous system as a bridge between the unconfined vacuum system and a constricting protein scaffold, we investigate the steric and electronic effects of the environment on the photodynamical behavior of the chromophore. Specifically, we apply ab initio multiple spawning to simulate five picoseconds of nonadiabatic dynamics after photoexcitation, resolving the excited-state pathways responsible for internal conversion in the aqueous chromophore. We identify an ultrafast pathway that proceeds through a short-lived (sub-picosecond) imidazolinone-twisted (I-twisted) species and a slower (several picoseconds) channel that proceeds through a long-lived phenolate-twisted (P-twisted) intermediate. The molecule navigates the non-equilibrium energy landscape via an aborted hula-twist-like motion toward the one-bond-flip dominated conical intersection seams, as opposed to following the pure one-bond-flip paths proposed by the excited-state equilibrium picture. We interpret our simulations in the context of time-resolved fluorescence experiments, which use short- and long-time components to describe the fluorescence decay of the aqueous GFP chromophore. Our results suggest that the longer time component is caused by an energetically uphill approach to the P-twisted intersection seam rather than an excited-state barrier to reach the twisted intramolecular charge-transfer species. Irrespective of the location of the nonadiabatic population events, the twisted intersection seams are inefficient at facilitating isomerization in aqueous solution. The disordered and homogeneous nature of the aqueous solvent environment facilitates non-selective stabilization with respect to I- and P-twisted species, offering an important foundation for understanding the consequences of selective stabilization in heterogeneous and rigid protein environments.
The overwhelming number of variables in a complex protein environment creates a fundamental roadblock to the rational design of FPs. Due to the rigidity of its 11-stranded β-barrel,22,23 GFP envelops its chromophore in a highly ordered pocket of amino acids. Stiff protein backbones foster heterogeneous electronic and steric fingerprints that can significantly impact the chromophore's behavior.8 Because amino acids in GFP form complex networks of covalent and non-covalent interactions, it is difficult to predict the net effect of various residues acting on the chromophore. To filter out the effects that dictate chromophore behavior, we use HBDI− to monitor the molecule's response to incremental perturbations of the environment. Using the gas-phase dynamics of HBDI− as a baseline,24 we identify the influence of steric and electronic effects on chromophore behavior in solution, focusing on the homogeneous and disordered case of aqueous HBDI−.
Solvated HBDI− straddles the line between the freedom permitted by the gas-phase and the confinement imposed by a protein. In water, HBDI− experiences steric and electronic effects from the solvent, which are absent in vacuo, while in an environment that is less constricting than a protein scaffold. Time-resolved spectroscopies reveal that GFP has an excited-state lifetime on the order of nanoseconds20,25,26 and a fluorescence quantum yield near unity.8 Yet, HBDI− experiences ultrafast internal conversion in vacuum and in solution at ambient temperature.27–31 Twisting around the bridge bonds connecting the imidazolinone (I) and phenolate (P) moieties of the isolated chromophore leads to ultrafast deactivation that decreases fluorescence quantum yield.31–33 As highlighted in previous computational studies,34,35 non-methylated HBDI− (pHBI−) undergoes twisted intramolecular charge-transfer (TICT) as rotations of the I- and P-moieties become sufficiently large. The direction of charge-transfer (CT) on the excited state depends on whether rotation occurs along the I- or P-dihedral (ϕI/ϕP dihedrals illustrated in Fig. 1, respectively). Whereas rotation of ϕI induces CT from the I-ring to the P-ring in the gas-phase, rotation of ϕP induces CT from the P-ring to the I-ring.
Fig. 1 Structure of the HBDI− model chromophore illustrating key geometric parameters. Cyan atoms represent carbon, red atoms represent oxygen, blue atoms represent nitrogen, and white atoms represent hydrogen. Blue, red, and orange bonds connect atoms belonging to the P-ring, I-ring, and methine bridge, respectively. ϕI is defined as the dihedral angle formed by NI and the carbon atoms of the methine bridge. ϕP is defined as the dihedral angle formed by the carbon atoms of the methine bridge and CP. The pyramidalization of the central methine carbon atom, θpyr, is calculated as and uses the bridge bond vectors in a method similar to Radhakrishnan et al.71 For reference, the pyramidalization angles of ideal sp2 and sp3 hybridized carbon atoms using this definition are 0° and 55°, respectively. |
Using ultrafast fluorescence and polarization spectroscopy experiments,27,36,37 Meech and co-workers identified similar short (hundreds of femtoseconds) and long (picoseconds) decay components for HBDI− across a variety of solvents. In water, time constants of 210 fs and 1.1 ps were reported for fluorescence decay—significantly faster than the 1.3/11.5 ps excited-state lifetimes reported in vacuo.21,30 Early quantum chemical calculations on gas-phase pHBI−38 suggested that a volume-conserving “hula-twist” deactivation mechanism39,40 is responsible for the limited sensitivity of the chromophore's excited-state lifetime on solvent viscosity. However, more recent calculations implicate drastic torsional motion along either of the methine bridge bonds as responsible for the ultrafast deactivation of the anionic chromophore.34,41 This behavior is corroborated by high-level quantum mechanics/molecular mechanics (QM/MM) simulations in water using extended multistate multireference second-order perturbation theory (XMS-CASPT2)42 and spin-flip time-dependent density functional theory.43 These studies found that the dominant decay pathway of pHBI− is through the I-twisted channel. However, the limited simulation time considered (1 ps) did not allow identification of the sources responsible for the multi-timescale decay in solution. To resolve this, we simulate multiple picoseconds of the aqueous HBDI− photodynamics while monitoring important geometric and electronic features of the system.
Ultrafast dispersed pump-dump-probe experiments of HBDI− in water present excited-state decay lifetimes in the 620–940 fs range for twisted excited-state intermediates.44 Although the imperfect selectivity of a dump pulse may make it difficult to attribute time constants to specific processes, the Z/E isomerization quantum yield derived from these experiments was below the detection threshold of ∼5%. Solution-phase studies on the chromophore's protonated counterpart, HBDI, demonstrate that the solvent can have a noticeable impact on isomerization quantum yield.45,46 In aprotic solvents, HBDI recorded Z/E photoisomerization quantum yields of ∼50%. This yield drops to ∼10–20% in protic solvents, where excited-state proton transfer is a competing process. While torsion along ϕI can result in the E-isomer (Z/E isomerization), torsion along ϕP will regenerate the Z-isomer due to the symmetry of the P-ring (Z/Z isomerization). Thus, Z/Z isomerization products are experimentally indistinguishable from decay pathways that do not involve isomerization. For this reason, it remains unclear what ratio of excited-state species proceeds toward photoproduct formation (Z/E isomerization), undergoes P-ring flip (Z/Z isomerization), or re-populates the original ground state (Z-isomer).
A key ingredient of QM/MM (or fully QM) simulations of photochemistry is the proper description of the potential energy surface topography around conical intersections (CIs). Many electronic structure methods are not capable of correctly describing S0/S1 CI regions47–49 (e.g. time-dependent density functional theory and other single-reference methods). Following our recent gas-phase study,24 we use the α-CASSCF method,50 which provides a correct and efficient description of the exact degeneracy and relative energetics along key deactivation coordinates. This enables us to extend our simulations to multiple picoseconds. Consequently, we are able to monitor the chromophore's approach toward the CI seams and provide predictions for the photoproduct branching. The shape of CIs (e.g. peaked vs. sloped) can also be critically important in determining the photoproducts.51–54 Although it appears that the CIs become more peaked upon solvation for the neutral GFP chromophore55,56 (suggesting a shortening of the excited-state lifetime and a larger isomerization quantum yield), the shape of the CIs that govern nonadiabatic transitions for HBDI− in water has been largely unexplored.
Recent femtosecond stimulated Raman spectroscopy (FSRS) and transient absorption experiments of aqueous HBDI−31 fill in some of the gaps left by ultrafast fluorescence experiments. While time-resolved fluorescence experiments elucidate the temporal signatures of highly fluorescent HBDI− species, i.e., closer to the Franck–Condon (FC) point, FSRS is able to track the progression of HBDI− along vibrational modes of dark, non-fluorescent species. FSRS studies yield three time constants for the excited-state dynamics of HBDI−: ∼350 fs, 2–3 ps, and ∼250 ps.31 These were tentatively assigned to the formation of a charge-separated (CS) species, the transition from the CS state to a TICT state, and ground-state recovery of the Z-isomer, respectively.
In this work, we explore the internal conversion dynamics following photoexcitation of explicitly-solvated HBDI−. Using ab initio multiple spawning57–61 (AIMS), we provide insight into the primary deactivation pathways in aqueous solution and each pathway's relative accessibility to CIs. In particular, we investigate the characteristics of HBDI− (i) at the FC region, (ii) out of the FC region as the molecule approaches CI seams, (iii) upon encountering the CI seam, and (iv) away from the CI seam as it returns to the ground state. This non-equilibrium picture (chromophore and solvent response) is compared to the corresponding excited-state equilibrium solvation picture, defined by the free energy surface of aqueous HBDI−. Our results are interpreted in the context of previous experiments performed in silico, in vacuo, and in solution. Using this approach, we discuss the impact of solvation on HBDI− and develop a more complete picture of GFP chromophore photophysics.
Fig. 2 Active space orbitals of HBDI− from α(0.67)-SA3-CASSCF(4,3) at the FC point: the anionic, allylic (a) bonding, (b) non-bonding, and (c) anti-bonding orbitals. |
The α parameter was chosen to reproduce the S0 → S1 vertical excitation energy of the planar chromophore in solution, computed with the SA3-XMS-CASPT2(4,3) method72–74 (1s orbitals of heavy atoms were excluded from the perturbation treatment and a level shift of 0.3 a. u. was used). This led to an α value of 0.67 for hydrated HBDI− (Fig. S1†). This is very similar to the α value of 0.64 which we found for isolated HBDI− (used in the gas-phase calculations reported herein).24 As shown in Fig. S1,† this α-CASSCF method reproduces XMS-CASPT2 potential energy profiles along the two bridge torsional coordinates (ϕI and ϕP). Unless otherwise stated, all calculations were performed with α(0.67)-SA3-CASSCF(4,3)/6-31G* in explicit solvent. Relative energies, geometric parameters, and transition dipole moments of representative optimized geometries are provided in the ESI† (Fig. S1–S3 and Tables S1–S4). Fractional occupation number Hartree–Fock (FON–HF) was used to determine initial guess orbitals for the α-CASSCF calculations (using Gaussian broadening with an electronic temperature of 0.3 a. u.).75–78
Critical points were obtained using the DL-FIND geometry optimization library.79 Electronic structure calculations were performed with TeraChem.80–84
Following the initial solvent equilibration, the geometric constraints on the chromophore were lifted and a 10 ps ground-state QM/MM molecular dynamics (QM/MM-MD) trajectory, within the NVT ensemble, was run to obtain ICs for the nonadiabatic dynamics simulations. The chromophore was treated at the QM level (described above), whereas the waters and sodium ion were treated at the MM level. A D3 dispersion correction was included for the QM region, using Hartree–Fock parameters.92 Spherical boundary conditions were again employed to avoid evaporation of water molecules. QM/MM-MD simulations were performed with a Langevin thermostat at 300 K, using the time-reversible Niklasson integrator94 and a time step of 0.5 fs. After an equilibration period of 3 ps, a total of 500 temporally-equidistant samples were extracted from the remaining 7 ps of QM/MM-MD to generate an absorption spectrum. The stick spectra were convolved with a Gaussian lineshape function, using a full width at half maximum (FWHM) of 0.40 eV based on solution-phase experiments.95
(1) |
Fig. 3 (a) Simulated absorption spectra of HBDI− in gas-phase (red) and water (blue) using 500 ICs from each system. Sticks represent the S0/S1 excitation energy and their relative oscillator strengths for the 50 solvated ICs used for AIMS simulations, at the α(0.67)-SA3-CASSCF(4,3)/6-31G* level. Individual sticks for gas-phase (FWHM of 0.07 eV) and aqueous (FWHM of 0.40 eV) systems were convolved with a Gaussian envelope based on experimental FWHM. Simulated gas-phase data was taken from List et al.,24 sampling from a 300 K harmonic Wigner distribution and using α(0.64)-SA3-CASSCF(4,3)/6-31G*. Experimental gas-phase and solvated chromophore spectra were obtained by digitizing data from Nielsen et al.104 and Andersen et al.,95 respectively. (b) Ground-state radial distribution functions of HBDI− heteroatoms (highlighted in each panel and identified in Fig. 1) in water. Shaded regions represent the first solvation shell. The dotted line represents a value of one, indicating that the solvent has reached bulk-like behavior. |
To understand the limited impact of geometric perturbations, we analyze the initial distribution of key geometric parameters to observe how the sampled configuration space changes upon solvation. Fig. 4a shows the bond length distributions for the bridge bonds, rI and rP, from the ground-state QM/MM-MD. In water, the bridge bonds of HBDI− sample a narrower region of configuration space when compared to vacuum. This difference is related to whether a quantum or classical sampling technique is used, as indicated by the similar configuration space sampled by gas-phase and solution-phase HBDI− using QM/MM-MD sampling in both instances (Fig. S7†). Distributions for the solvated ICs used for the AIMS dynamics are shown in Fig. S8.† The average rI-to-rP bond length ratio in water is consistent with a stabilization of the phenolate resonance structure, in line with the RDFs in Fig. 3b. Therefore, we verify that the phenolate resonance structure of HBDI− is favored throughout the ground-state dynamics in water, as suggested by calculations in a polarizable continuum water model.41 While the phenolate and quinoid resonance structures contribute almost equally in gas-phase,24,68 the decrease in the rI-to-rP ratio upon solvation indicates a preference toward the phenolate resonance structure in solution. Interestingly, the distributions along the ϕI and ϕP dihedrals (Fig. 4b) remain almost identical between systems. The striking similarity between these angular distributions indicates that the presence of mobile water molecules does not restrict the chromophore's motion along its twisting dihedrals on the ground state. The concerted rotation around the I- and P-bonds, which are connected by the methine bridge carbon atom and adjacent to one another, is reminiscent of hula-twist motion.39,40 However, we observe conrotatory rotation of ϕI and ϕP, as opposed to the volume-conserving disrotatory motion.105
The two-dimensional S1 umbrella sampling along the bridge dihedrals in Fig. 5 shows that single-bond twisting about either of the bridge bonds is favored. The resulting PMF highlights four energetic wells located near 90° I- and P-twisted structures on S1. Interestingly, the excited-state dynamics do not follow the minimum free energy path to reach these energetic wells (vide infra). Instead, the chromophore reaches highly-twisted configurations by following either the I- or P-channels illustrated in Fig. 5. Compared to the gas-phase photodynamics of HBDI−,24 the solvated chromophore exhibits more pronounced hula-twist-like (concerted single- and double-bond rotation) motion out of the FC region.
Fig. 5 Relative free energy profile on S1 obtained from two-dimensional umbrella sampling along ϕI and ϕP. The zero point (marked with an “X”) corresponds to the planar structure. The observed deactivation pathways through I-twisting (red) and P-twisting (green) are depicted schematically. For reference, arrows with dashed lines represent deactivation pathways observed in gas-phase simulations.24 |
To separate out the electronic effects of equilibrium solvation, Fig. 6 displays the average S0/S1 energy gap across all windows in solvent and solvent-stripped configurations. In water (Fig. 6a), the observed minima along ϕI lie close to the CI seam, whereas rotation along ϕP leads to minima on S1 that are ∼1 eV removed from S0. This suggests that the I-twisted CI seam is more peaked than the P-twisted CI seam. The absence of electronic degeneracy associated with P-twisting hints that energetically and/or entropically unfavorable geometric changes are required to reach degeneracy along this pathway. Fig. 6b shows that I-twisted intermediates have lower S0/S1 energy gaps than those of P-twisted intermediates in the gas-phase, as observed in solution. This energetic asymmetry originates from the differences in electron affinity between the I- and P-rings.24 However, to conclude whether or not water stabilizes one twisted structure over the other, it is necessary to look at the energetic differences caused by solvation. The difference between water and gas-phase energy gaps (Fig. S9†) reveals that water introduces non-selective gap reduction (∼0.3 eV) for highly-twisted (∼90°) structures—the I- and P-twisted energy gaps decrease by approximately the same amount upon solvation. Yet, as previously discussed for Fig. 6a, the consequences of this gap reduction are significant. This equilibrium solvation picture suggests that (i) aqueous HBDI− proceeds along one-bond-flip-dominated coordinates toward the twisted CI seam, as in gas-phase,24 (ii) the I-twisted CI seam is more accessible than its P-twisted counterpart, and (iii) solvation promotes internal conversion through the I-twisted CI seam, similar to its neutral counterpart.55,56 Our nonadiabatic dynamics simulations of HBDI− show that (ii) and (iii) occur in water, while there is no evidence for (i). Rather than participating in a predominantly one-bond-flip mechanism (as in gas-phase), aqueous HBDI− reaches CI seams via an aborted hula-twist-like motion out of the FC region. Note that the concerted torsion is conrotatory as opposed to the more usual (for the hula-twist mechanism) disrotatory motion.
(2) |
Viewing the dynamics within defined time intervals reveals novel information regarding the time scales of each deactivation pathway available to aqueous HBDI−—two distinct sub-populations exist with different excited-state lifetimes (Fig. 8). In accordance with the decreased electronic density along the methine bridge upon excitation (see Fig. 2) facile rotation along ϕI and ϕP is permitted on the excited state. Monitoring the S1 population over time, with respect to the ϕI and ϕP dihedrals, illustrates (i) the accessibility and (ii) the population transfer efficiency of each internal conversion pathway. During the first 100 fs, the S1 wavepacket remains localized near the FC region. Rather than following the minimum energy path along one-bond-flip coordinates (Fig. 5), the conrotatory hula-twist-like motion exhibited during the ground-state dynamics (Fig. 4b) is preserved. This indicates a non-equilibrium solvation regime where the response of the surrounding waters is too slow to impede the concerted rotation of ϕI and ϕP out of the FC region, which is otherwise associated with a small free energy barrier (less than 0.2 eV). Once the free energy barrier along this path becomes too large to overcome, the wavepacket bifurcates via a process that involves a one-bond-flip-like rotation along either ϕI or ϕP – with one path leading to highly I-twisted structures (I-channel) and the other path leading to highly P-twisted structures (P-channel). The branching associated with this mechanism becomes apparent by 500 fs as exploration into the I- and P-channels enables population transfer to the ground state. By 1 ps, motion along the torsional angles is more prominent as the I- and P-channels arise as the exclusive avenues for population transfer. Essentially no TBFs occupy near-planar configurations after 1.75 ps of dynamics, as four distinct regions in configuration space emerge as dihedral basins. These basins coincide with the energetic wells identified in Fig. 5. Interestingly, 250 fs later, no evidence of the I-twisted species is present while the P-twisted species remains on S1 for several picoseconds. The disappearance of the I-twisted species on S1 is explained by an ultrafast population transfer to the ground state, permitted by the near-degeneracy highlighted in Fig. 6a. For this part of the wavepacket, twisting along ϕI is enough to induce population transfer to the ground state. The persistence of the P-twisted species for several picoseconds on the excited state signifies that dihedral torsion around ϕP alone is insufficient to facilitate efficient population transfer. Rather, additional geometric requirements are necessary for the P-species to participate in internal conversion.
The time scale differences established in Fig. 8 were refined using the S1 population profiles for the I- and P-deactivating species (Fig. 9). In the case of I-twisted deactivation, the entire excited-state population is transferred to the ground state within ∼1.5 ps, with decay initiating at ∼175 fs. This rapid initiation and conclusion of population transfer for the I-species is due to an easily accessible I-twisted CI seam that efficiently transfers population to the ground state, in accordance with Fig. 6. On the other hand, the P-deactivation mechanism is associated with a species that remains on S1 for ∼540 fs before population transfer begins. Furthermore, this mode of population transfer is much slower and less efficient, as ∼10% of the population remains on the excited state after 5 ps. A similar analysis compares the deactivation of red-shifted and blue-shifted ICs (Fig. S10, and Table S6†). Data from this partitioning approach suggests that blue-shifted ICs deactivate faster than red-shifted ICs, but further investigations are required to reach a definitive conclusion.
After identifying the characteristic twisting pathways, the dynamics out of the FC region were probed using various electronic and geometric parameters (Fig. 10, and S11†). Here, the parent TBFs are partitioned into I- and P-twisted species based on whether the I- or P-channel was taken, respectively. Our results corroborate that the dominant mode of deactivation (66%) is through torsion along ϕI, in agreement with nonadiabatic dynamics simulations at the XMS-CASPT2 (ref. 42) and spin-flip time-dependent density functional theory43 levels. The violin plots in Fig. 10 present the distribution of key electronic and geometric observables sampled across all parent TBFs of each species during the first 500 fs. Characteristic activity along the ϕI and ϕP twisting dihedrals is observed within the first 500 fs. As expected, the I-species samples along ϕI more aggressively than it does along ϕP, and the opposite behavior is true for the P-species. Worth noting, however, is the time scale at which both species reach ∼90° rotations of their respective dihedrals. Between 200 and 400 fs, both species begin to explore and populate these highly twisted configurations. As in the gas-phase,35 we observe that twisting along ϕI facilitates CT from the I-ring to the P-ring while twisting along ϕP facilitates CT from the P-ring to the I-ring. Furthermore, the molecule's rapid approach toward highly I- and P-twisted structures indicates that access to their respective TICT states is downhill in solution.
In the gas-phase, bond length alternation and bridge pyramidalization are the main coordinates used by HBDI− to reach the CI seam from the respective twisted S1 minimum.24 As shown in Fig. 10 (300–500 fs), displacement along the pyramidalization coordinate is attenuated for the I- and P-twisted species in solution relative to gas-phase (|θpyr| of 30° and 48° for I- and P-twisted MECIs,24 respectively). Because solvation essentially closes the energy gap of the I-twisted species but not completely for the P-twisted, the bridge pyramidalization requirement for reaching the I-twisted CI seam essentially vanishes. Compared to species along the I-channel, larger-amplitude displacements of the pyramidalization coordinate are required to trigger population transfer along the P-channel. This suggests that the path to the I-twisted CI seam is essentially energetically downhill, while it is uphill for the P-twisted case.
In addition to monitoring the geometric and electronic changes that occur during the first few hundred femtoseconds following photoexcitation, the response of the solvent during the first 1 ps was investigated. Fig. 11 shows the evolution of the first solvation peak for the parts of the wavepacket following the I- and P-twisted pathways. In the case of the I-deactivating species in Fig. 11a, there are no significant changes to the RDF profiles of OP, OI, or NI during the first 1 ps. This contrasts the evolution of RDFs for the same atoms in the P-deactivating species in Fig. 11b. Over the first 1 ps, the maximum of the first solvation peak for OP and OI gradually decreases and increases, respectively, toward the RDFs associated with equilibrated ϕP-twisted structures. This trend is consistent with the TICT behavior identified in Fig. 10. Upon torsion along ϕP, CT occurs from the P-ring to the I-ring. Consequently, a P-twisted intermediate will have (i) a more negative I-ring and (ii) a more positive P-ring than its pre-twisted counterpart. Thus, it is expected that the surrounding solvent would rearrange to compensate for this CT process by shifting increased solvent density from the OP atom toward the OI atom. Interestingly, the opposite trend is absent for the I-species. We attribute this difference to the disparate deactivation time scales along ϕI and ϕP that are presented in Fig. 9. Because deactivation through the I-channel occurs so rapidly, water does not have time to relax fully around I-twisted structures before internal conversion. Furthermore, we suspect that the ground-state solvent arrangement (Fig. 3b) is beneficial for facilitating ultrafast deactivation via I-twisting, since water is already preferentially localized around the P-ring prior to TICT. Such behavior explains the absence of large shifts of excited-state RDF peaks in Fig. 11a. On the other hand, deactivation through the P-channel is more gradual than decay through the I-channel, due to the barrier between the P-twisted CI seam and the corresponding TICT intermediate. Consequently, the solvent has time to reorient itself around P-twisted structures on the excited state. As a result, we observe gradual shifts of excited-state RDF peaks in Fig. 11b.
A geometric analysis at the points of population transfer elucidates the difference between the deactivation profiles of the I- and P-species. For I-deactivation, the spawning events that lead to population transfer generally occur near geometries with ϕI angles near 90° and ϕP angles that lie relatively close to planarity, and vice versa for P-deactivation (Fig. 12). The relationship between key dihedral and pyramidalization angles reveals that, in addition to lasting longer on S1, P-twisted TBFs demand larger pyramidalization angles than I-twisted TBFs to reach their respective CI seams. To investigate the implications of this difference, the absolute population transfer was calculated for each spawning event as a function of the pyramidalization and dihedral angle values upon entering the spawning region (Fig. 12a). Movie S1† provides a time lapse of these population and angle parameters to clarify this point. For a direct comparison of the spawned geometries associated with each pathway, the pyramidalization angle was measured for each spawned geometry of I-deactivating (Fig. 12b) and P-deactivating (Fig. 12c) TBFs. From these distributions, it is clear that pyramidalization requirements for P-deactivating TBFs are more demanding (peaked at ∼30°) than the requirements for I-deactivating TBFs (peaked at ∼10°). The larger pyramidalization angles needed to reach the P-twisted CI gate access to the P-twisted region of the CI seam. Therefore, we conclude that limited access to an uphill CI seam is responsible for the lingering S1 population of the P-species.
Fig. 12 (a) Absolute population transfer and pyramidalization angle associated with the bridge carbon atom for spawned geometries of aqueous HBDI− from parent TBFs. Absolute population transfer is defined by the fraction of original S1 population transferred to S0. Plotted values of ϕI, ϕP, and θpyr characterize geometries that initiate population transfer. A histogram (using 5° bins) of pyramidalization angle magnitudes is provided for spawned TBFs that transfer population through torsion dominated by (b) ϕI or (c) ϕP twisting. Pyramidalization angles measured for I- and P-twisted MECIs in gas-phase24 are represented with vertical lines in (b) and (c). |
In addition to being more accessible than the P-twisted CI seam, the I-twisted CI seam appears to be more efficient at transferring population to the ground state. On average, population transfer events mediated by the I- and P-channels transfer 56 ± 32% and 35 ± 26% of the incoming S1 population to S0 per spawning event, respectively (Fig. S12†). This is expected based on the more direct approach toward the I-twisted CI seam, than for its P-twisted counterpart, as supported by Fig. 5 and 6. The uphill pyramidalization requirements and lack of efficiency explain the more gradual deactivation of the P-twisted species.
Relative to gas-phase HBDI−,24 aqueous HBDI− generally demands less pyramidalization to reach I- and P-twisted CI seams (Fig. 12b and c). This difference between systems becomes most apparent when comparing internal conversion tendencies associated with the I-channel. In gas-phase, I-twisted configurations transfer population to S0 while sampling along a larger range of ϕP values than in water. Based on our analysis regarding pyramidalization gating, we conclude that this behavior is due to the topology differences of the I-twisted CI seam across systems. Because solvation reduces the S0/S1 energy gap for I-twisted structures, aqueous HBDI− has smaller pyramidalization requirements to reach the I-twisted CI seam than does gas-phase HBDI−. Thus, I-twisted structures of the solvated chromophore spend less time sampling ϕP torsion than their gas-phase counterparts, as they undergo internal conversion through a near-barrierless process. The I-twisted spawning geometries of aqueous HBDI−, which are associated with ultrafast population transfer, are consistent with rapid accessibility out of the FC region via an aborted hula-twist-like mechanism (as presented in Fig. 8). Because reaching the aqueous P-twisted CI seam from a P-twisted TICT intermediate is an energetically uphill process, P-twisted intermediates significantly sample configurations where ϕP is above 90°. The enhanced sampling along this coordinate may explain the higher proportion of spawning geometries skewed toward the E-isomer, relative to this proportion for ϕI from I-twisted spawns (Fig. S13†).
Based on vibrational signatures obtained from FSRS experiments, Taylor et al. proposed that the formation of a CS species and the transition from this CS state to a TICT state were responsible for measured short (∼350 fs) and long (∼2–3 ps) time components, respectively.31 Our data suggests a different origin of these time constants. As supported by the substantial torsional displacements sampled within the first ∼200 fs (Fig. 10), the formation of TICT states is essentially barrierless. In other words, we do not see evidence of a non-twisted CS intermediate in water. Rather, we observe two competing processes: (i) an I-twisting process that easily reaches a downhill CI seam and (ii) a P-twisting process that encounters a barrier as it approaches an uphill CI seam. As an alternative, we propose that the accessibility of the I- and P-twisted CI seams, from the corresponding TICT states, is the primary factor responsible for the short (caused by mainly I-species) and long (caused by mainly P-species) time components reported from the FSRS experiments. However, rigorously establishing this interpretation of the FSRS fingerprints requires calculation of the vibrational signatures of I- and P-twisted structures and remains a task for future work.
Fig. 14 Two-dimensional fluorescence spectrum, probed at 500 nm, for (a) all TBFs, (b) I-deactivating TBFs, and (c) P-deactivating TBFs. (d) Time-resolved fluorescence signals of HBDI− calculated from AIMS simulations in water. The maximum fluorescence intensities of I- and P-deactivating species are scaled based on the relative proportion of ICs that deactivate through the I- and P-channels. Experimental data, following 400 nm excitation, for HBDI− in water was provided by Meech and co-workers.27,103 To align the experimental31 and theoretical fluorescence wavelength maxima, a shift of 0.13 eV was applied to the theory spectra (Fig. S15†). Transparent lines represent the error obtained from bootstrapping analysis. |
Our AIMS simulations provide novel mechanistic and electronic insight into the behavior that governs the photodynamics of the GFP chromophore. QM/MM-MD simulations indicate that surrounding HBDI− with mobile solvent molecules has minimal impact on the configuration space sampled by the chromophore in water. Yet, the electronic effects of water blue-shift the absorption spectrum. During gas- and solution-phase dynamics on S0, preferential twisting along ϕI and ϕP in a hula-twist-like (conrotatory) fashion is observed. The subsequent internal conversion mechanisms predominantly occur through rotation of ϕI in both systems.24 The absence of rigid electronic and steric constraints on the chromophore likely explains the ultrafast deactivation tendencies reported by experiments for both systems.21,27,30,31 However, the stabilization of highly-twisted structures and the promotion of electronic degeneracy, especially with respect to ϕI which results in a downhill CI seam, are responsible for shorter lifetimes in water when compared to gas-phase. Furthermore, the fitted time constants for the S1 population decay (609 ± 181 fs and 3.1 ± 1.2 ps) are associated with internal conversion through the I- and P-channels, respectively. Through a direct comparison of AIMS simulations and experimental fluorescence spectra, we provide mechanistic insight to explain the observed fluorescence profile of HBDI− in water.27 We attribute the short experimental time constant (210 fs) describing HBDI− fluorescence in water to the departure from the FC region (of both I- and P-deactivating species), while the longer experimental fluorescence time constant (1.1 ps) corresponds to the longer-lived, P-deactivating species.
Although our simulations do not include an explicit protein environment, our findings imply that the mobility and electronic distribution of residues surrounding the chromophore in a protein can tune the chromophore's behavior on the excited state. From our simulations, we predict that modifications to the protein environment can lead to internal conversion pathways with notably different population transfer rates, fluorescence profiles, and isomerization quantum yields. Using Dronpa2 variants, Romei and co-workers110 have demonstrated that chromophore modifications can bias the preferred directionality of the CT process and modify observables of interest in the protein. Our direct observation of TICT processes in solution provides evidence that steering the chromophore toward a given pathway can significantly alter its properties. To characterize the photochemical processes that drive the dynamics of the chromophore in protein scaffolds, we will extend our simulations and analyses to fluorescent protein systems.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1sc02508b |
This journal is © The Royal Society of Chemistry 2021 |