ACS Publications. Most Trusted. Most Cited. Most Read
Excited-State Relaxation of Hydrated Thymine and Thymidine Measured by Liquid-Jet Photoelectron Spectroscopy: Experiment and Simulation
My Activity

Figure 1Loading Img
  • Open Access
Article

Excited-State Relaxation of Hydrated Thymine and Thymidine Measured by Liquid-Jet Photoelectron Spectroscopy: Experiment and Simulation
Click to copy article linkArticle link copied!

View Author Information
Max-Born-Institut für nichtlineare Optik und Kurzzeitspektroskopie, Max-Born-Straße 2A, 12489 Berlin, Germany
Catalysis Research Center, Hokkaido University, Sapporo 001-0021, Japan
Department of Frontier Materials Chemistry, Graduate School of Science and Technology, Hirosaki University, Hirosaki 036-8561, Japan
Open PDFSupporting Information (1)

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2015, 137, 8, 2931–2938
Click to copy citationCitation copied!
https://doi.org/10.1021/ja511108u
Published February 11, 2015

Copyright © 2015 American Chemical Society. This publication is licensed under CC-BY.

Abstract

Click to copy section linkSection link copied!

Time-resolved photoelectron spectroscopy is performed on thymine and thymidine in aqueous solution to study the excited-state relaxation dynamics of these molecules. We find two contributions with sub-ps lifetimes in line with recent excited-state QM/MM molecular dynamics simulations (J. Chem. Phys.2013, 139, 214304). The temporal evolution of ionization energies for the excited ππ* state along the QM/MM molecular dynamics trajectories were calculated and are compatible with experimental results, where the two contributions correspond to the relaxation paths in the ππ* state involving different conical intersections with the ground state. Theoretical calculations also show that ionization from the nπ* state is possible at the given photon energies, but we have not found any experimental indication for signal from the nπ* state. In contrast to currently accepted relaxation mechanisms, we suggest that the nπ* state is not involved in the relaxation process of thymine in aqueous solution.

Copyright © 2015 American Chemical Society

1 Introduction

Click to copy section linkSection link copied!

DNA bases strongly absorb in the ultraviolet (UV) range and, yet, are astonishingly photostable. (1) The photostability of the DNA bases is discussed to play a key role for the photostability of DNA, itself. The biological relevance of DNA and its photoprotection mechanism triggered extensive research during the last ten years. (1-5) It is nowadays widely accepted that photostability of DNA bases is due to very efficient electronic relaxation via conical intersections with the ground state. Nevertheless, the detailed relaxation paths are still not fully resolved, in particular for the pyrimidine bases (thymine, cytosine, and uracil) there is still a lot of controversy.
Figure 1 shows the low-lying electronic states of isolated and hydrated thymine at the ground-state equilibrium geometry. In the gas phase an intense absorption band corresponding to excitation to the bright ππ* state centered at 250 nm (4.95 eV) (6) is observed.
In isolated thymine, excited-state decay times of <50 fs, <1 ps, and ∼2.4···6.4 ps are observed. (7) Different interpretations have been suggested. (i) Dynamics on the S2(ππ*) surface, where the fs dynamics is due to relaxation to an S2 minimum, while barrier crossing to the S2/S1 conical intersection (CI) is responsible for the ps decay. (8, 9) (ii) A barrierless pathway on the ππ* state from the Franck–Condon (FC) region to an ethylene-like CI with the ground state and trapping of part of the excited-state population in a ππ* state planar minimum. (10) (iii) A fast transition from S2 → S1 followed by a slower decay S1 → S0. (11) A recent time-resolved Auger spectroscopic study is in support of the last interpretation. (12)
In aqueous solution, due to the interactions with solvent water molecules, the ππ* state is stabilized, while the nπ* state is strongly destabilized with respect to the ground state (see Figure 1)
Around the FC region, the vertical ionization energy in aqueous solution is 8.1 eV (13) associated with a π hole (D0–1)), according to Koopmans’ theorem. The vertical ionization energy for ionizing into D1(n–1) is 9.6 eV. (13) In our experiments the total absorbed photon energy is 9.86 eV, indicating that ionization from both excited states is possible in the FC region.
Fluorescence upconversion (FU) experiments in aqueous solution revealed a biexponential sub-ps fluorescence decay, while excited-state absorption decays monoexponentially. (14-16) Transient absorption (TA) studies have also revealed that the vibronic ground state is repopulated with two distinct rates. (17) The faster vibronic ground-state repopulation occurs within a few ps. However, a significant fraction of the vibronic ground state is only populated on a time scale of several tens of ps. The picture that has emerged from experimental and theoretical findings for the solution phase so far is the excited ππ* state population very quickly bifurcates: Part of it evolves toward the CI with the ground state, quickly undergoing internal conversion. A significant fraction however is trapped in a long-lived low-lying state, which was suggested to be a close-lying nπ* state. Based on its lifetime and the fact that this state directly populates the ground state, it was excluded that this dark state is a triplet state. (17) Vibrational cooling of the hot ground state takes about few ps. (17-19)
However, a recent excited-state quantum mechanics/molecular mechanics (QM/MM) molecular dynamics (MD) simulation study suggests that the excited ππ* state population evolves along two directions: either toward the (ππ*(C5–C6) /S0)CI involving a twisting of the C5–C6 double bond with a strong puckering of the C6 atom or toward the (ππ*(C4–O8) /S0)CI which involves an out-of-plane displacement of the C4–O8 carbonyl group (20) (for labeling of the atoms see Figure 1). They also suggested that the nπ* state is not involved in the excited-state relaxation of thymine in aqueous solution. Very recently, Minezawa investigated the deactivation pathways of thymine in aqueous solution by the combined method of linear-response free energy and collinear spin-flip TD-DFT and found that the free energy of the deactivation pathway through (ππ*(C4–O8)/S0)CI is significantly stabilized by the solute–solvent interaction (21) supporting the findings of the QM/MM-MD simulation.
In this work, we present the first time-resolved photoelectron studies of thymine (Thy) and its nucleoside thymidine (Thd) in aqueous solution. In Thd, the hydrogen at the N1 position is replaced by a deoxyribose moiety.

2 Experimental Section

Click to copy section linkSection link copied!

The experimental setup is described in detail in ref 23. Briefly, the experiment is carried out at a regeneratively amplified Ti:sapphire laser system delivering 40 fs pulses at 800 nm with a 1 kHz repetition rate. Part of the laser power is used to pump an optical parametric amplifier (TOPAS, light conversion). By frequency doubling of the TOPAS output UV pulses in the range of (238···248) nm (5.21···5.00 eV, respectively) are generated. The remaining part is frequency tripled by second harmonic and sum frequency generation in α-BaB2O4 (BBO) crystals. UV pulses are attenuated to about 35–40 nJ and focused onto a ∼10 μm thin liquid jet. Both pump and probe beams are chopped allowing for a separate measurement of the one-color photoelectron signals and its subtraction from the two-color signal on a pulse-to-pulse basis. Photoelectrons are collected by a magnetic bottle type time-of-flight (TOF) spectrometer.

Figure 1

Figure 1. Electronic potential energies of low-lying states of isolated and hydrated thymine around the FC region. Excited-state energies for the gas phase and solvatochromic shift of the nπ* state are taken from ref 20, and the ππ* energy of the hydrated molecule is taken from experimental values. (15) Ionization energies for gas phase are taken from ref 22, and ionization energies of hydrated molecules are taken from ref 13. The absorbed photon energy, i.e., the sum of pump photon energy hνpu and probe photon energy hνpr, is given by the horizontal line. The molecular structure and the numbering of the atoms is shown in the right part of the figure.

The sample is a 0.5 mM aqueous solution of Thy or a 1 mM aqueous solution of Thd, buffered at pH 8 (1 mM tris(hydroxymethyl)-aminomethane (TRIS) and hydrochloric acid (HCl)). In addition, the sample contains 30 mM of sodium chloride (NaCl) to prevent electrokinetic charging of the jet. (24) NaCl does not significantly contribute to the observed one- and two-color signals at the given pulse intensities. (25) DNA bases, nucleosides and TRIS were purchased from Sigma-Aldrich Co., sodium chloride from Merck. All substances were used without further purification. The sample solution was prepared using demineralized water (residual resistivity: 0.25 μS/cm).

3 Computational Section

Click to copy section linkSection link copied!

The QM/MM-MD simulations are performed both in the ground and excited states of a thymine molecule solvated in water. The QM region is composed of the thymine molecule and treated by ab initio calculations, while the MM region consists of surrounding solvent water molecules that are represented by the SPC/Fw model. (26) The electrostatic interaction of electrons in the QM region with the MM effective point charges are taken into account through one-electron integrals. The periodic supercell with a length of 16 Å is employed and the thymine molecule is solvated by 122 water molecules, where the number of water molecules is determined to reproduce the density of water at ambient conditions. The long-range electrostatic interactions between the MM molecules are treated by the Ewald sum and the cutoff distance of 8 Å (half the length of the supercell) is employed for electrostatic interactions between the QM (thymine) and MM molecules. All quantum chemical calculations are performed by the MOLPRO2010.1 package, (27) and it is interfaced with the in-house MD program for the QM/MM-MD simulations.
The QM/MM-MD simulations in the electronic ground state are performed to estimate the vertical excitation and ionization energies in aqueous solution. The Møller–Plesset second-order perturbation (MP2) method with the DZP quality of basis set are employed in the MD simulations. The Newtonian equation of motion for nuclei is integrated by the velocity Verlet algorithm with a time step of 0.5 fs and the temperature is controlled at 300 K by the Nosé–Hoover thermostat. The configurations are picked up at every 250 fs from the MD trajectories, and the complete active space second-order perturbation theory (CASPT2) method is employed to calculate the vertical excitation and ionization energies at these configurations. Then, the averages over 100 configurations are taken to estimate the excitation and ionization energies in aqueous solution, which are used to calculate the constant shift for excited-state ionization energy (see below).
The excited-state QM/MM-MD simulations are performed at the CASPT2 level of theory. The initial coordinates and velocities are taken from the above-mentioned ground-state QM/MM-MD run by picking up coordinates and velocities at every 500 fs. The excited-state MD simulations are performed without the thermostats, and they are initiated from the lowest ππ* state. When the energy difference between the ground and ππ* states becomes less than 0.2 eV, we assume that the molecule reaches the CI region.
Here we describe the details of the CASPT2 calculations. The DZP quality of basis sets are also employed, and a level shift with a value of 0.2 a.u. (∼5.4 eV) is applied. The notation of CASPT2(m,n) is occasionally used, in which case the active space for a reference state-averaged complete active space self-consistent field (SA-CASSCF) wave function is composed of m electrons and n orbitals (SA-CASSCF(m,n)). The excited-state QM/MM-MD simulations in the ππ* state are carried out by CASPT2(4,4), where the active space is comprised of only π orbitals and the lowest two states are averaged with equal weights in the SA-CASSCF calculations. In the previous paper, it was confirmed that the CASPT2(4,4) method provides a reliable description of the ππ* state. (20) Along the excited-state MD trajectories, the CASPT2(12,9) calculations are performed at fixed intervals to obtain the energies of the ground, lowest ππ*, and * states. Here, the active space is comprised of eight π orbitals (five π orbitals are doubly occupied and three are unoccupied in the closed-shell configuration) and one lone-pair orbital that belongs to the O8 atom. The ground, ππ*, and * states (a total of three states) are averaged with equal weights in the SA-CASSCF(12,9) calculations. Obviously, our simulations focus on the deactivation process occurring only in the ππ* state, and the transition from the ππ* to * state is not taken into account. However, the relative energies of the * state are obtained along the trajectories, and the role of the * state is discussed in section 5.
The cationic ground and first excited states (D0 and D1), which correspond to either π-hole or n-hole states, are also calculated by the CASPT2 method along the QM/MM-MD trajectories. The active space is the same as that used in the calculations of the neutral molecule, except that the number of electrons in the active space is 11 in this case, and therefore it can be denoted by CASPT2(11,9). Here, two states, D0 and D1, are averaged with equal weights in the reference SA-CASSCF(11,9) calculations.
At this point we should comment on the relative stability of the lowest ππ* and * states around the FC region. In the gas phase, theoretical calculations predicted that the lowest ππ* and * states lie close to each other and the order of these two state changes depending on the computational levels. The literature values before 2008 are compiled in ref 28, and the more recent values are found in our previous paper. (20) In aqueous solution, the ππ* state is stabilized, while the * state is strongly destabilized due to the hydrogen bond with solvent water molecules. This destabilization of the * state is rationalized by recognizing that the electronic transition involves the transfer of an electron in the lone-pair orbital of the O8 atom, which acts as a hydrogen-bond acceptor, toward the diffusive π* orbital, leading to a decrease in the solute–solvent interaction. Because of this, for both thymine and uracil, it is generally accepted that the * state is less stable than the ππ* state around the FC region in aqueous solution (see a recent review by Improta and Barone (5)). Specifically, for thymine, the time-dependent density functional theory (TD-DFT) calculations employing explicit four water molecules and polarizable continuum models predicted that the vertical excitation energies are 4.97 (−0.09) and 5.27 (+0.41) eV for the lowest ππ* and * states, respectively (the numbers in parentheses indicate the shifts from the gas-phase values). (15) Also, Etinski and Marian employed six water molecules in conjunction with the conductor-like screening model (COSMO) and the excitation energies obtained by the coupled cluster with approximate treatment of doubles (CC2) method were 5.08 (−0.18) and 5.48 (+0.51) eV for the ππ* and * excitations, respectively. (29) Our calculated values by the QM(CASPT2)/MM-MD approach are 4.47 (−0.30) and 5.41 (+0.54) eV for the ππ* and * excitations, respectively, and it is seen that the ππ* excitation energy is closer to the experimental value of 4.68 eV (15) than other theoretical works. Also, the shift from the gas-phase value for the ππ* excitation is in quite good agreement with the experiment (−0.27 eV). (6, 15)
As often pointed out before, the accuracy of the ionization energy and the excitation energy is not similar. Therefore, the excited-state ionization potential energy is corrected using the experimental value, as has been done before. (8, 30) Here, the constant shift Δ is introduced, which is adjusted to ensure that the calculated excited-state ionization energy matches with the experimental value, and is given by(1)where S0 is the neutral ground state and D0 is the cationic ground state. When the thymine molecule is equilibrated in aqueous solution, D0 is associated with the π-hole cation. The averages of the S1(ππ*) vertical excitation and ionization energies in aqueous solution are calculated to be ⟨ES1(ππ*)CASPT2ES0CASPT2aq = 4.47 eV and ⟨ED0–1)CASPT2ES0CASPT2aq = 8.74 eV, respectively, where the computational details are given above. By using the experimental values of IES0→D0–1)vertical,expt,aq = 8.1 eV (13) and ΔES0→S1(ππ*)vertical,expt,aq = 4.68 eV, (15) we find Δ = 0.85 eV for the correction of the excited-state ionization potential energy in aqueous solution.
In order to examine the possibility of detecting the signal if the molecule is trapped in the * state, we have also run the QM/MM-MD simulations in the * state by employing the SA-CASSCF(12,9) method. In the CASSCF method, the potential energy of the ππ* state is usually highly overestimated, and therefore even in aqueous solution, the S1 state is characterized by the * state. After running several picoseconds in the * state, the molecule is equilibrated in the * state. Then we further run additional 5 ps, and we use CASPT2 calculations along these trajectories and evaluate the ionization energies from the * state.
We note that the electron correlation effects are different between the ππ* and * states, and therefore when evaluating the excited-state ionization energy, it is advisible to use different shifts depending on which excited state the ionization occurs. In obtaining a shift for the ionization from the * state, however, no experimental values of the S2(*) vertical excitation energy ΔES0→S2(*)vertical,expt,aq are available. The experimental value of the ionization energy to the D1(n–1) state is given as IES0→D1(n–1)vertical,expt,aq = 9.6 eV, (13) and our theoretical values of the ionization energy to the D1(n–1) state and the S2(*) vertical excitation energy are ⟨ED1(n–1)CASPT2ES0CASPT2aq = 10.55 eV and ⟨ES2(*)CASPT2ES0CASPT2aq = 5.41 eV, respectively. Therefore, the shift is given as Δ = ΔES0→S2(nπ*)vertical,expt,aq – 4.46 eV. It is known that the CASPT2 method tends to overestimate the electron correlation of the ππ* state for DNA bases (e.g., see Table 1 in ref 31) and this fact is reflected in the lower ππ* excitation energy than the experiment: 4.95 eV (experiment) and 4.77 eV (CASPT2) in the gas phase. Since the electron correlation in the ππ* state is generally stronger than that in the * state, we speculate that the calculated * excitation energy is closer to the experimental value (if available) than the ππ* state. Therefore, if we assume that ΔES0→S2(*)vertical,expt,aq is the same as the calculated value, the shift is estimated as Δ = 0.95 eV, which is slightly larger than the shift for the ππ* state of Δ = 0.85 eV. We use the shift of Δ = 0.85 eV throughout this work, but given this small difference of the shifts (∼0.1 eV), the discussion of the ionization from the * state given below will not be affected.

4 Results

Click to copy section linkSection link copied!

Figure 2 shows the time-dependent photoelectron spectra and the global fit results for Thy. Results for Thd are shown in Figure S4 in the Supporting Information. (32) Photoelectrons are generated, if the probe pulse photon energy is larger than the time-dependent ionization energy of the excited molecule. Therefore, the signal decay can in principle reflect the excited state population decay, an increase of the energy gap between excited neutral state and final ionic state exceeding the probe photon energy, and time-dependent ionization cross sections. This has to be kept in mind when discussing the observed lifetimes.
We observe photoelectron signal from the excited molecules for kinetic energies up to about 2.5 eV and for a delay range Δτ of −0.5 ps ≤ Δτ ≤ 1 ps. Signal at positive (negative) delays originates from photoreaction triggered by the pulse at 266 nm (238 nm) and probed by 238 nm (266 nm). A global analysis of the data was performed by fitting with the following function:(2)Ai(Ekin) is the decay associated spectrum and Pi(t) is the excited-state population of the ith contribution, while G(t) is the Gaussian cross-correlation function.
In a first attempt, the data were analyzed in terms of only a single exponential excited-state population decay for each delay direction. These results are shown in Figures S3 and S4 in the Supporting Information. Although this model fits the data quite reasonably, there are systematic residuals. Therefore, the experimental results were analyzed based on a model that assumes two parallel relaxation paths:(3)Sm and Sn may be different excited states or refer to different trajectories on the same excited-state potential energy surface. Please note that by parallel relaxation path we mean the temporal evolution of wave packets generated at two different points in phase space. For negative delays we used only one relaxation path (see below for the discussion). For comparison, the Supporting Information shows also results, in which two components in each delay direction are considered. Following the analysis of Ullrich et al., (7) we assume in our global analysis that the spectra of the two components do not change in time. The cross-correlation width (220 fs fwhm) and the temporal origin t0 were independently determined from nonresonant two-color photoelectron signal of gaseous nitric oxide (NO) and buffer solution. At long delay times we observe a very small negative pump probe signal (the same at positive and negative delays), which we attribute to dead time effects of the multi channel plate in single electron counting mode. This is included in the fit by a temporally constant negative contribution. The fit result together with the data are shown in Figure 2a, while Figure 2c shows the residuals of the global analysis. Fit parameters are summarized in Table 1.
At positive delays, the residuals mainly reflect statistical deviations of the data and the model fits the experiment very well. However, at negative delays close to the temporal overlap, residuals are systematically positive (in particular for Thd, see Figure S4 in the Supporting Information (32)), indicating that the model underestimates the excited-state signal. It becomes evident from Figures 2d and S4k in the Supporting Information (32) that the deviations seem to arise from a very tiny temporal shift of the excited-state population decay which may simply reflect the limitations of an exponential decay model. Instead, the prepared wave packet needs some time to evolve and to reach the CI; that is, it is reasonable to expect a small delay in excited-state depopulation.
The decay associated spectra are shown in Figure 2b and the population dynamics of the individual contributions in Figure 2d. In positive delay direction we observe contributions with lifetimes of 70 fs (120 fs) and 410 fs (390 fs) for Thy (Thd). These lifetimes are shorter than those obtained in FU or TA experiments (cf. Table 1). As mentioned above, possible reasons are the energy gap betweeen the excited and ionic state exceeding the probe photon energy or time-dependent ionization cross sections. However, also the observed lifetimes in TA or FU may be influenced from time-dependent cross sections, overlapping features (e.g., hot ground state absorption), and a direct comparison is difficult. The lifetimes we have retrieved for Thy are slightly shorter than for Thd and thus seem to follow the opposite trend than observed in TA and FU. However, we stress that in all cases the given errors are from the fit, only. Due to the rather long cross-correlation or instrumental response functions, the real errors should be expected much larger in either case.

Figure 2

Figure 2. Time-resolved photoelectron spectrum of thymine, globally analyzed by three spectral contributions. (a) Comparison between data (color) and global fit (contour lines), (b) decay associated spectra, (c) residuals (color) and global fit (contour lines), and (d) population dynamics of individual contributions. The color scale of the residuals covers a range of ±20% of the maximum photoelectron signal. Vertical lines in panel b mark the maximum kinetic energies for the individual contributions.

Table 1. Summary of Fit Parameters and Comparison with Fluorescence Upconversion Results (14, 15)a
  τ1/fsτ2/fsτt/psτ3/fsαβ
this workThy70 ± 10410 ± 40 320 ± 400.690.08
 Thd120 ± 10390 ± 10 290 ± 200.540.09
FUThy (15)195 ± 17633 ± 18  0.56 ± 0.02 
 Thd (33)150 ± 20720 ± 30  0.70 ± 0,02 
TAThy (16, 17) 540 ± 40b30 ± 13 0.77 
a

τ1-3 are derived lifetimes, where the superscript “–” refers to the value obtained for negative delay direction. Ai = ∫0Ai(Ekin) dEkin is the energy integrated signal associated with decay i at delay = 0. α = (A1+)/(A1+ + A2+) is the fraction of the signal of the faster decaying component A1+ in the total signal A1+ + A2+ in positive delay direction. β = (A)/(A1+ + A2+) is the ratio between signal in negative delay direction A to signal in positive delay direction A1+ + A2+. Uncertainties given correspond to one standard deviation as obtained from the fit.

b

Thd.

The average kinetic energies(4)for Thy (Thd) are 1.42 eV (1.43 eV) for the faster decaying component and 1.26 eV (1.19 eV) for the second component. For negative delays the excited-state lifetime is 320 fs (290 fs) and the AKE is 1.08 eV (1.00 eV). Apparently, the kinetic energy difference between the contributions at positive delays and the one at negative delays does not correspond to the probe photon energy difference (5.2 – 4.66 eV = 0.54 eV). We have previously observed, that the photoelectron signal at very low kinetic energies in liquid jet experiments is in general strongly reduced, possibly as an effect of a cloud of charged water clusters surrounding the liquid jet. (24, 25) This would lead to an overestimation of AKE values, in particular for the lower probe photon energy case of negative delay directions. Similar behavior was observed in case of adenine and adenosine, where additional photoelectron band broadening effects were explained by wave packet dynamics within the pulse duration. In case of adenine/adenosine we have shown, that the difference of maximum kinetic energies is in good agreement with the probe photon energy difference. (25) For Thy (Thd), the maximum kinetic energies of the individual contributions are 2.8 eV (2.9 eV) and 2.6 eV (2.6 eV) at positive delays and 2.0 eV (2.0 eV) at negative delays (cf. Figures 2b and S4h in the Supporting Information (32)).
The difference between the maximum kinetic energy for the slower decaying component at positive delays and the component in negative direction fits very well with the difference in probe photon energy which indicates that both contributions originate from the same relaxation path. This, in turn, would indicate the absence of the faster decay in the negative direction. However, the low signal at negative delays makes the deconvolution very difficult, and we cannot exclude that both components contribute to the signal here.
We have also performed a global analysis assuming four contributions, i.e., two in each direction. This analysis gives similarly good results as with three components only (see the Supporting Information (32)). For Thy we find very similar decay rates at negative and positive delays, although the additional, quickly decaying component has only a very small share in the entire signal. In Thd, the contribution is even smaller, and its retrieved lifetime (3 fs) is unrealistically small. In conlusion, considering three contributions for the global analysis of our data is reasonable, in particular if the main focus is on the dynamics in positive delays. The overall similarity between the spectra and observed decay times obtained for Thy and Thd leads to the conclusion that the excited-state relaxation occurs along the same coordinates in both molecules.
If the molecules are excited at 238 nm (negative delays), the observed excited-state signal decays with a lifetime of 300 fs, i.e., 100 fs faster than in case of 266 nm excitation. This may reflect both accelerated excited-state dynamics due to higher excess energy or a smaller FC window for ionization due to lower probe photon energy, i.e., the accessible phase space for observation of the excited state is smaller at negative than at positive delays.

5 Assignment of the Relaxation Paths

Click to copy section linkSection link copied!

Our data show that the excited-state dynamics of Thy and Thd after optical excitation at 4.66 eV involve two components and occur on sub-ps time scales. No experimental evidence for a trapped excited-state population persisting as long as 30 ps is observed. Given the electron effective attenuation length in liquid water of 2–5 nm at low kinetic energies, (34-37) an assignment of the two observed contributions to dynamics occurring in molecules in the bulk or at the surface may be conceivable. However, we discard this assignment. A detailed discussion is found in the Supporting Information. (32)
Can the two observed pathways be assigned to the different pathways observed in recent QM/MM-MD simulations? (20) What is the role of the nπ* state for the excited-state dynamics?
We have performed QM/MM-MD simulations as described in ref 20 but extended the calculations to 30 trajectories. We found 6 trajectories reaching (ππ*(C4–O8)/S0)CI and 24 reaching ethylene-like (ππ*(C5–C6)/S0)CI regions. For the ethylene-like relaxation paths, the nπ* state is always above the ππ* state (see the Supporting Information, Figure S7(left) (32)). For the trajectories that reach (ππ*(C4–O8) /S0)CI regions, the nπ* state sometimes gets close to the ππ* state (see the Supporting Information, Figure S7(right) (32)). In our simulation, we cannot treat the hopping from the ππ* to the nπ* state. But if the transition to the nπ* state takes place, then the molecule will relax to the equilibrium structure in the nπ* state. We will show below that the photon energy of the probe pulse is large enough to ionize the equilibrated molecule in the nπ* state. We notice that in Figure S7(right) the S0, ππ*, and nπ* states come close to each other at some points, suggesting the existence of the three-state conical intersections around that region. The involvement of the three-state conical intersections in the relaxation dynamics of pyrimidine bases has been previously proposed. (38-40)
In Figure 3 we show the temporal evolution of the potential energy difference between the ππ* state and the ground state S0, the nπ* excited state as well as the ionic ground state D0–1) along representative trajectories toward either conical intersection region.
In addition, we show also the energetic range that can be overcome by absorption of a probe pulse photon for the case that the excess energy is completely and instantaneously distributed over different modes. We find that the molecule can be ionized along a significant fraction of the relaxation paths. Given the typically very broad photoelectron bands, we do expect to detect photoelectrons during the entire residence of the molecule in the ππ* state.
It is interesting to compare the decay associated spectra with the electron kinetic energies calculated for the relaxation along the two pathways. In Figure 4 (right panel) we show photoelectron kinetic energies along the individual trajectories.
For relaxation toward (ππ*(C5–C6)/S0)CI, we find an average kinetic energy of 1.8 eV at temporal overlap. The ionization energy very quickly increases as nuclear dynamics sets in. Already after 25 fs, the average kinetic energy is only about 1 eV. For comparison with experimental data we need to consider our limited temporal resolution. Averaging over the first 100 fs from simulations we find an average photoelectron kinetic energy of 1.14 eV. For the relaxation toward (ππ*(C4–O8)/S0)CI, the average kinetic energy at t = 0 is 1.4 eV, averaging over the first 100 fs, it amounts to 1.05 eV. The temporally averaged values are slightly smaller than derived from our experiments. However, as discussed before, we expect to measure larger average kinetic energies due to reduced photoelectron signal at low kinetic energies. Considering this aspect, the difference between the average kinetic energies found from the simulation and experiment are in good agreement. Also, the amplitude ratios of the two relaxation paths found in experiment and simulation are in good agreement: In the simulation, 20% of the trajectories proceed toward (ππ*(C4–O8)/S0)CI, while 80% reach (ππ*(C5–C6)/S0)CI. In the experiment we find an amplitude ratio of about 30:70 for the two components. Figure 4 (left panel) shows the excited-state population which can be ionized by the probe photon energy as retrieved from the simulations. We find that the ethylene-like path leads to faster internal conversion than the path toward (ππ*(C4–O8)/S0)CI; nevertheless, both relaxation paths take place on a sub-ps time scale. Based on the qualitative agreement of the photoelectron spectrum, the relative amplitudes and the excited-state lifetime, between QM/MM-MD simulations and experiment we assign the short-lived component to a wave packet following the ethylene-like relaxation path and the longer-lived component to a wave packet which proceeds toward (ππ*(C4–O8)/S0)CI.

Figure 3

Figure 3. Potential energy differences between the ππ* state and the neutral ground state, the nπ* state and the ionic ground state along representative trajectories taken from ref 20 that reach (ππ*(C5–C6)/S0)CI (left) and (ππ*(C4–O8)/S0)CI (right) regions. Gray-shaded areas (between red and black solid lines) show ionization windows of ππ* state for instantaneous and complete redistribution of excess energy (E(ππ*) + ℏωpr) in case of 5.2 eV probe pulses. Diamonds mark the total absorbed energy at temporal overlap (ℏωpu + ℏωpr + E(S0; t = 0)).

Figure 4

Figure 4. (left) Average population of trajectories in the ππ* state that can be ionized for the given pump–probe time delay and photon energy (ℏωpr = 5.2 eV) (black line). The trajectories are divided into two groups: trajectories that reach the (ππ*(C5–C6)/S0)CI and (ππ*(C4–O8)/S0)CI regions, and the average populations of each group are shown in blue and red dotted lines. (right) Kinetic energies of photoelectrons along the trajectories. The trajectories are again divided into two groups and shown in blue and red circles. The averages of kinetic energy at t = 0 over all trajectories are calculated to be 1.72 eV. The averages over trajectories that reach the (ππ*(C5–C6)/S0)CI and (ππ*(C4–O8)/S0)CI regions are 1.80 and 1.40 eV, respectively.

Is the * state involved in the relaxation dynamics?
As mentioned previously, our simulation can not treat the hopping from the ππ* to the nπ* state. If such a transition occurs, the molecule will relax to the nπ* minimum. We will show here, that the photon energy of the probe pulse (5.2 eV) is high enough to ionize the molecule equilibrated in the nπ* state. We have run QM/MM-MD simulations in the nπ* state in aqueous solution (see the Computational section) and found that the ionization energies from the nπ* state to D(n–1) or D(π–1) (n-hole is not always the cationic ground state along trajectories, n-hole and π-hole states switch often) along the trajectories are close to and on average slightly below the photon energy of the probe pulse (see Figure 5).
Ionizing the molecule in the nπ* state removes the electron from the π* orbital, i.e., due to the charater of the electronic configuration in the single-electron transition, we expect predominantly ionization into the D(n–1) ionic state, leaving an n-hole in the molecule. (41) Therefore, contributions from ionization into D(π–1) are expected to be small. Considering the typically very broad photoelectron bands of hydrated molecules, we would clearly expect to detect photoelectrons from molecules trapped in the nπ* state.

Figure 5

Figure 5. Potential energy difference between the D(π–1) and nπ* states (red line), and the D(n–1) and nπ* states (blue line) along the QM/MM-MD trajectories that are equilibrated in the nπ* state using the SA-CASSCF(12,9) calculations. The energies are recalculated by CASPT2(12,9) and CASPT2(11,9) for the neutral and cationic molecules, respectively. The energies of the cationic states are shifted using the correction described in the text, and the experimentally applied photon energy (5.2 eV) is shown by the dotted line.

This would have been visible in particular at longer delays, when the ππ* population has decayed. From the absence of any long-lived signal, we conclude that the nπ* state does not play an important role in the relaxation process. The biexponential repopulation of the vibronic ground state observed in transient absorption (17) might instead reflect ground state dynamics along different relaxation paths. We want to note that also in the case of cytosine and its nucleoside we do not observe any contribution from the nπ* state, although the ionization energies are significantly lower for the nπ* state; that is, the ionization window is significantly larger, and also in spite of the fact that at least for Cyt nucleotide the dark state yield was reported to be 41%. (17)

6 Conclusions

Click to copy section linkSection link copied!

We have followed the ππ* population excited by short UV pulses in thymine and thymidine. In agreement with recent excited-state QM/MM-MD simulations we find two different relaxation paths toward conical intersections with the ground state. (20) These relaxation paths are assigned to an ethylene-like relaxation toward the (ππ*(C5–C6)/S0)CI and toward (ππ*(C4–O8)/S0)CI regions. When molecules are excited by 266 nm light (4.66 eV) excited-state lifetimes of thymine (thymidine) are 70 fs (120 fs) and 410 fs (390 fs), respectively. These lifetimes are somewhat shorter than observed in fluorescence upconversion or in our QM/MM-MD simulations (see Figure 4, left panel).
No contribution from the nπ* excited state has been observed, although we have shown that the probe photon energy is sufficiently high to ionize the equilibrated molecule in the nπ* state. From the absence of any long-lived signal we conclude that the nπ* state is very likely not involved in the excited-state relaxation of Thy and Thd. We may speculate that in this case, the biexponential repopulation of the vibronic ground state observed in transient absorption (17) reflects ground-state dynamics along different relaxation paths.
Finally, we want to note that similar data are available for cytosine which will be discussed in a separate paper. Also for cytosine, the signal from the nπ* state is absent, although cytosine has a significantly larger ionization window due to lower ionization energies.
Our work shows the potential of liquid jet time-resolved photoelectron spectroscopy to study the excited-state dynamics of solvated molecules in dilute solutions. It provides important new information complementary to transient absorption or fluorescence upconversion and will thus contribute significantly to a better understanding of excited-state dynamics of solvated molecules and, in particular, of hydrated DNA bases.

Supporting Information

Click to copy section linkSection link copied!

Details on the detection efficiency at low kinetic energy, the role of inelastic processes, global analysis, time-resolved photoelectron spectra, global analysis results for thymine and thymidine, an extended time scan for thymine, and a discussion of alternative signal assignment. This material is available free of charge via the Internet at http://pubs.acs.org.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
    • Akira Nakayama - Catalysis Research Center, Hokkaido University, Sapporo 001-0021, Japan Email: [email protected]
    • Andrea Lübcke - Max-Born-Institut für nichtlineare Optik und Kurzzeitspektroskopie, Max-Born-Straße 2A, 12489 Berlin, Germany Email: [email protected]
  • Authors
    • Franziska Buchner - Max-Born-Institut für nichtlineare Optik und Kurzzeitspektroskopie, Max-Born-Straße 2A, 12489 Berlin, Germany
    • Shohei Yamazaki - Department of Frontier Materials Chemistry, Graduate School of Science and Technology, Hirosaki University, Hirosaki 036-8561, Japan
    • Hans-Hermann Ritze - Max-Born-Institut für nichtlineare Optik und Kurzzeitspektroskopie, Max-Born-Straße 2A, 12489 Berlin, Germany
  • Notes
    The authors declare no competing financial interest.

Acknowledgment

Click to copy section linkSection link copied!

The authors thank F. Noack for his support by providing the laser system in the femtosecond application laboratory of the Max-Born-Institut Berlin. This work was financially supported by Deutsche Forschungsgemeinschaft, Project LU 1638/1-1, and by KAKENHI (Grant-in-Aid for Scientific Research). Part of the calculations was performed on supercomputers at Research Center for Computational Science, Okazaki, Japan.

References

Click to copy section linkSection link copied!

This article references 41 other publications.

  1. 1
    Crespo-Hernández, C. E.; Cohen, B.; Hare, P. M.; Kohler, B. Chem. Rev. 2004, 104, 1977
  2. 2
    de Vries, M. S.; Hobza, P. Annu. Rev. Phys. Chem. 2007, 58, 585
  3. 3
    Middleton, C. T.; de La Harpe, K.; Su, C.; Law, Y. K.; Crespo-Hernández, C. E.; Kohler, B. Annu. Rev. Phys. Chem. 2009, 60, 217
  4. 4
    Kleinermanns, K.; Nachtigallová, D.; de Vries, M. S. Int. Rev. Phys. Chem. 2013, 32, 308
  5. 5
    Improta, R.; Barone, V. Top. Curr. Chem. 2014, 355, 329
  6. 6
    Abouaf, R.; Pommier, J.; Dunet, H. Chem. Phys. Lett. 2003, 381, 486
  7. 7
    Ullrich, S.; Schultz, T.; Zgierski, M. Z.; Stolow, A. Phys. Chem. Chem. Phys. 2004, 6, 2796
  8. 8
    Hudock, H. R.; Levine, B. G.; Thompson, A. L.; Satzger, H.; Townsend, D.; Gador, N.; Ullrich, S.; Stolow, A.; Martínez, T. J. J. Phys. Chem. A 2007, 111, 8500
  9. 9
    Szymczak, J. J.; Barbatti, M.; Hoo, J. T. S.; Adkins, J. A.; Windus, T. L.; Nachtigallová, D.; Lischka, H. J. Phys. Chem. A 2009, 113, 12686
  10. 10
    Merchán, M.; González-Luque, R.; Climent, T.; Serrano-Andrés, L.; Rodríguez, E.; Reguero, M.; Peláez, D. J. Phys. Chem. B 2006, 110, 26471
  11. 11
    Lan, Z.; Fabiano, E.; Thiel, W. J. Phys. Chem. B 2009, 113, 3548
  12. 12
    McFarland, B. K.; Farrell, J. P.; Miyabe, S.; Tarantelli, F.; Aguilar, A.; Berrah, N.; Bostedt, C.; Bozek, J. D.; Bucksbaum, P. H.; Castagna, J. C.; Coffee, R. N.; Cryan, J. P.; Fang, L.; Feifel, R.; Gaffney, K. J.; Glownia, J. M.; Martinez, T. J.; Mucke, M.; Murphy, B.; Natan, A.; Osipov, T.; Petrovíc, V. S.; Schorb, S.; Schultz, T.; Spector, L. S.; Swiggers, M.; Tenney, I.; Wang, S.; White, J. L.; White, W.; Gühr, M. Nat. Comm. 2014, 5, 4235
  13. 13
    Seidel, R. Electronic-Structure Interactions in Aqueous Solutions: A Liquid-Jet Photoelectron-Spectroscopy Study. Ph.D. Thesis; Technische Universitaet Berlin: Berlin, 2011.
  14. 14
    Gustavsson, T.; Sharonov, A.; Markovitsi, D. Chem. Phys. Lett. 2002, 351, 195
  15. 15
    Gustavsson, T.; Bányász, A.; Lazzarotto, E.; Markovitsi, D.; Scalmani, G.; Frisch, M. J.; Barone, V.; Improta, R. J. Am. Chem. Soc. 2006, 128, 607
  16. 16
    Pecourt, J.-M. L.; Peon, J.; Kohler, B. J. Am. Chem. Soc. 2001, 132, 10370
  17. 17
    Hare, P. M.; Crespo-Hernández, C. E.; Kohler, B. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 435
  18. 18
    Gustavsson, T.; Bányász, A.; Sarkar, N.; Markovitsi, D.; Improta, R. Chem. Phys. 2008, 350, 186
  19. 19
    Improta, R.; Barone, V.; Lami, A.; Santoro, F. J. Phys. Chem. B 2009, 113, 14491
  20. 20
    Nakayama, A.; Arai, G.; Yamazaki, S.; Taketsugu, T. J. Chem. Phys. 2013, 139, 214304
  21. 21
    Minezawa, N. J. Chem. Phys. 2014, 141, 164118
  22. 22
    Urano, S.; Yang, X.; LeBreton, P. R. J. Mol. Struct. 1989, 214, 315
  23. 23
    Buchner, F.; Lübcke, A.; Heine, N.; Schultz, T. Rev. Sci. Instrum. 2010, 81, 113107
  24. 24
    Preissler, N.; Buchner, F.; Schultz, T.; Lübcke, A. J. Phys. Chem. B 2013, 117, 2422
  25. 25
    Buchner, F.; Ritze, H.-H.; Lahl, J.; Lübcke, A. Phys. Chem. Chem. Phys. 2013, 15, 11402
  26. 26
    Wu, Y.; Tepper, H. L.; Voth, G. A. J. Chem. Phys. 2006, 124, 024503
  27. 27
    Werner, H. J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Celani, M. S. P.; Korona, T.; Lindh, R.; Mitrushenkov, A.; Rauhut, G.; Shamasundar, K. R.; Adler, T. B.; Amos, R. D.; Bernhardsson, A.; Berning, A.; Cooper, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.; Goll, E.; Hampel, C.; Hesselmann, A.; Hetzer, G.; Hrenar, T.; Jansen, G.; Köppl, C.; Liu, Y.; Lloyd, A. W.; Mata, R. A.; May, A. J.; McNicholas, S. J.; Meyer, W.; Mura, M. E.; Nicklaß, A.; O’Neill, D. P.; Palmieri, P.; Pflüger, K.; Pitzer, R.; Reiher, M.; Shiozaki, T.; Stoll, H.; Stone, A. J.; Tarroni, R.; Thorsteinsson, T.; Wang, M.; Wolf, A.MOLPRO version 2010.1 a package of ab initio programs; Birmingham, UK.
  28. 28
    Zechmann, G.; Barbatti, M. J. Phys. Chem. A 2008, 112, 8273
  29. 29
    Etinski, M.; Marian, C. M. Phys. Chem. Chem. Phys. 2010, 12, 4915
  30. 30
    Tao, H.; Levine, B. G.; Martínez, T. J. J. Phys. Chem. A 2009, 113, 13656
  31. 31
    Schreiber, M.; Silva-Junior, M. R.; Sauer, S. P. A.; Thiel, W. J. Chem. Phys. 2008, 128, 134110
  32. 32

    See the Supporting Information.

  33. 33
    Onidas, D.; Markovitsi, D.; Marguet, S.; Sharonov, A.; Gustavsson, T. J. Phys. Chem. B 2002, 106, 11367
  34. 34
    Konovalov, V. V.; Raitsimring, A. M.; Tsvetkov, Y. D. Int. J. Rad. Appl. Instr. C Rad. Phys. Chem. 1988, 32, 623
  35. 35
    Plante, I.; Cucinotta, F. A. New J. Phys. 2009, 11, 063047
  36. 36
    Buchner, F.; Schultz, T.; Lübcke, A. Phys. Chem. Chem. Phys. 2012, 14, 5837
  37. 37
    Suzuki, Y.-I.; Nishizawa, K.; Kurahashi, N.; Suzuki, T. Phys. Rev. E 2014, 90, 010302(R)
  38. 38
    Matsika, S. J. Phys. Chem. A 2005, 109, 7538
  39. 39
    Kistler, K. A.; Matsika, S. J. Chem. Phys. 2008, 128, 215102
  40. 40
    González-Vázquez, J.; González, L. Chem. Phys. Chem. 2010, 11, 3617
  41. 41
    Satzger, H.; Townsend, D.; Zgierski, M. Z.; Patchkovskii, S.; Ullrich, S.; Stolow, A. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 10196

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 61 publications.

  1. Yuki Obara, Srijon Ghosh, Alexander Humeniuk, Shota Kamibashira, Shunsuke Adachi, Toshinori Suzuki. Formation of Ground-State Intermediate during Electronic Relaxation of Pyrimidine Nucleobases. Journal of the American Chemical Society 2025, Article ASAP.
  2. Alessandro Loreti, Victor Manuel Freixas, Davide Avagliano, Francesco Segatta, Huajing Song, Sergei Tretiak, Shaul Mukamel, Marco Garavelli, Niranjan Govind, Artur Nenov. WFOT: A Wave Function Overlap Tool between Single- and Multi-Reference Electronic Structure Methods for Spectroscopy Simulation. Journal of Chemical Theory and Computation 2024, 20 (11) , 4804-4819. https://doi.org/10.1021/acs.jctc.4c00310
  3. Juliana Cuéllar-Zuquin, Ana Julieta Pepino, Ignacio Fdez. Galván, Ivan Rivalta, Francesco Aquilante, Marco Garavelli, Roland Lindh, Javier Segarra-Martí. Characterizing Conical Intersections in DNA/RNA Nucleobases with Multiconfigurational Wave Functions of Varying Active Space Size. Journal of Chemical Theory and Computation 2023, 19 (22) , 8258-8272. https://doi.org/10.1021/acs.jctc.3c00577
  4. Yo-ichi Yamamoto, Toshinori Suzuki. Distortion Correction of Low-Energy Photoelectron Spectra of Liquids Using Spectroscopic Data for Solvated Electrons. The Journal of Physical Chemistry A 2023, 127 (11) , 2440-2452. https://doi.org/10.1021/acs.jpca.2c08046
  5. Natsumi Orimo, Yo-ichi Yamamoto, Shutaro Karashima, Alexie Boyer, Toshinori Suzuki. Ultrafast Electronic Relaxation in 6-Methyluracil and 5-Fluorouracil in Isolated and Aqueous Conditions: Substituent and Solvent Effects. The Journal of Physical Chemistry Letters 2023, 14 (11) , 2758-2763. https://doi.org/10.1021/acs.jpclett.3c00195
  6. Yuta Miura, Yo-ichi Yamamoto, Shutaro Karashima, Natsumi Orimo, Ayano Hara, Kanae Fukuoka, Tatsuya Ishiyama, Toshinori Suzuki. Formation of Long-Lived Dark States during Electronic Relaxation of Pyrimidine Nucleobases Studied Using Extreme Ultraviolet Time-Resolved Photoelectron Spectroscopy. Journal of the American Chemical Society 2023, 145 (6) , 3369-3381. https://doi.org/10.1021/jacs.2c09803
  7. William G. Fortune, Michael S. Scholz, Helen H. Fielding. UV Photoelectron Spectroscopy of Aqueous Solutions. Accounts of Chemical Research 2022, 55 (24) , 3631-3640. https://doi.org/10.1021/acs.accounts.2c00523
  8. Zachary N. Heim, Daniel M. Neumark. Nonadiabatic Dynamics Studied by Liquid-Jet Time-Resolved Photoelectron Spectroscopy. Accounts of Chemical Research 2022, 55 (24) , 3652-3662. https://doi.org/10.1021/acs.accounts.2c00609
  9. Qiushuang Xu, Daniel Aranda, Martha Yaghoubi Jouybari, Yanli Liu, Meishan Wang, Javier Cerezo, Roberto Improta, Fabrizio Santoro. Nonadiabatic Vibrational Resonance Raman Spectra from Quantum Dynamics Propagations with LVC Models. Application to Thymine. The Journal of Physical Chemistry A 2022, 126 (41) , 7468-7479. https://doi.org/10.1021/acs.jpca.2c05271
  10. Michael S. Scholz, William G. Fortune, Omri Tau, Helen H. Fielding. Accurate Vertical Ionization Energy of Water and Retrieval of True Ultraviolet Photoelectron Spectra of Aqueous Solutions. The Journal of Physical Chemistry Letters 2022, 13 (30) , 6889-6895. https://doi.org/10.1021/acs.jpclett.2c01768
  11. Yeonsig Nam, Francesco Montorsi, Daniel Keefer, Stefano M. Cavaletto, Jin Yong Lee, Artur Nenov, Marco Garavelli, Shaul Mukamel. Time-Resolved Optical Pump-Resonant X-ray Probe Spectroscopy of 4-Thiouracil: A Simulation Study. Journal of Chemical Theory and Computation 2022, 18 (5) , 3075-3088. https://doi.org/10.1021/acs.jctc.2c00064
  12. Shota Tsuru, Bikramjit Sharma, Masanari Nagasaka, Christof Hättig. Solvent Effects in the Ultraviolet and X-ray Absorption Spectra of Pyridazine in Aqueous Solution. The Journal of Physical Chemistry A 2021, 125 (33) , 7198-7206. https://doi.org/10.1021/acs.jpca.1c05183
  13. Conaill F. Perry, Inga Jordan, Pengju Zhang, Aaron von Conta, Fernanda B. Nunes, Hans Jakob Wörner. Photoelectron Spectroscopy of Liquid Water with Tunable Extreme-Ultraviolet Radiation: Effects of Electron Scattering. The Journal of Physical Chemistry Letters 2021, 12 (11) , 2990-2996. https://doi.org/10.1021/acs.jpclett.0c03424
  14. Blake A. Erickson, Zachary N. Heim, Elisa Pieri, Erica Liu, Todd J. Martinez, Daniel M. Neumark. Relaxation Dynamics of Hydrated Thymine, Thymidine, and Thymidine Monophosphate Probed by Liquid Jet Time-Resolved Photoelectron Spectroscopy. The Journal of Physical Chemistry A 2019, 123 (50) , 10676-10684. https://doi.org/10.1021/acs.jpca.9b08258
  15. Johan Hummert, Geert Reitsma, Nicola Mayer, Evgenii Ikonnikov, Martin Eckstein, Oleg Kornilov. Femtosecond Extreme Ultraviolet Photoelectron Spectroscopy of Organic Molecules in Aqueous Solution. The Journal of Physical Chemistry Letters 2018, 9 (22) , 6649-6655. https://doi.org/10.1021/acs.jpclett.8b02937
  16. Bert M. Pilles, Benjamin Maerz, Jinquan Chen, Dominik B. Bucher, Peter Gilch, Bern Kohler, Wolfgang Zinth, Benjamin P. Fingerhut, Wolfgang J. Schreier. Decay Pathways of Thymine Revisited. The Journal of Physical Chemistry A 2018, 122 (21) , 4819-4828. https://doi.org/10.1021/acs.jpca.8b02050
  17. Jamie W. Riley, Bingxing Wang, Joanne L. Woodhouse, Mariana Assmann, Graham A. Worth, and Helen H. Fielding . Unravelling the Role of an Aqueous Environment on the Electronic Structure and Ionization of Phenol Using Photoelectron Spectroscopy. The Journal of Physical Chemistry Letters 2018, 9 (4) , 678-682. https://doi.org/10.1021/acs.jpclett.7b03310
  18. Javier Cerezo, Yanli Liu, Na Lin, Xian Zhao, Roberto Improta, and Fabrizio Santoro . Mixed Quantum/Classical Method for Nonadiabatic Quantum Dynamics in Explicit Solvent Models: The ππ*/nπ* Decay of Thymine in Water as a Test Case. Journal of Chemical Theory and Computation 2018, 14 (2) , 820-832. https://doi.org/10.1021/acs.jctc.7b01015
  19. Markus Kowalewski, Benjamin P. Fingerhut, Konstantin E. Dorfman, Kochise Bennett, and Shaul Mukamel . Simulating Coherent Multidimensional Spectroscopy of Nonadiabatic Molecular Processes: From the Infrared to the X-ray Regime. Chemical Reviews 2017, 117 (19) , 12165-12226. https://doi.org/10.1021/acs.chemrev.7b00081
  20. Ana Julieta Pepino, Javier Segarra-Martí, Artur Nenov, Roberto Improta, and Marco Garavelli . Resolving Ultrafast Photoinduced Deactivations in Water-Solvated Pyrimidine Nucleosides. The Journal of Physical Chemistry Letters 2017, 8 (8) , 1777-1783. https://doi.org/10.1021/acs.jpclett.7b00316
  21. Valentyn I. Prokhorenko, Alessandra Picchiotti, Martina Pola, Arend G. Dijkstra, and R. J. Dwayne Miller . New Insights into the Photophysics of DNA Nucleobases. The Journal of Physical Chemistry Letters 2016, 7 (22) , 4445-4450. https://doi.org/10.1021/acs.jpclett.6b02085
  22. Wilmer Arbelo-González, Rachel Crespo-Otero, and Mario Barbatti . Steady and Time-Resolved Photoelectron Spectra Based on Nuclear Ensembles. Journal of Chemical Theory and Computation 2016, 12 (10) , 5037-5049. https://doi.org/10.1021/acs.jctc.6b00704
  23. Roberto Improta, Fabrizio Santoro, and Lluís Blancafort . Quantum Mechanical Studies on the Photophysics and the Photochemistry of Nucleic Acids and Nucleobases. Chemical Reviews 2016, 116 (6) , 3540-3593. https://doi.org/10.1021/acs.chemrev.5b00444
  24. Javier Segarra-Martí, Marco Garavelli, and Francesco Aquilante . Multiconfigurational Second-Order Perturbation Theory with Frozen Natural Orbitals Extended to the Treatment of Photochemical Problems. Journal of Chemical Theory and Computation 2015, 11 (8) , 3772-3784. https://doi.org/10.1021/acs.jctc.5b00479
  25. Shota Tsuru, Bikramjit Sharma, Christof Hättig, Dominik Marx. Nuclear Quantum Effects Have a Significant Impact on UV/Vis Absorption Spectra of Chromophores in Water. Angewandte Chemie International Edition 2025, 64 (3) https://doi.org/10.1002/anie.202416058
  26. Shota Tsuru, Bikramjit Sharma, Christof Hättig, Dominik Marx. Nuclear Quantum Effects Have a Significant Impact on UV/Vis Absorption Spectra of Chromophores in Water. Angewandte Chemie 2025, 137 (3) https://doi.org/10.1002/ange.202416058
  27. Danillo Valverde, Sean J. Hoehn, Eduardo D. Koyanagui, Sarah E. Krul, Carlos E. Crespo‐Hernández, Antonio Carlos Borin. Theoretical and Experimental Evaluation of the Electronic Relaxation Mechanisms of 2‐Pyrimidinone: The Primary UVA Absorbing Moiety of the DNA and RNA (6–4) Photolesion. ChemPhotoChem 2024, 8 (12) https://doi.org/10.1002/cptc.202400070
  28. Piao Xu, Dongdong Wang, Duoduo Li, Jinyou Long, Song Zhang, Bing Zhang. UV wavelength-dependent photoionization quantum yields for the dark 1 nπ* state of aqueous thymidine. Physical Chemistry Chemical Physics 2024, 26 (41) , 26251-26257. https://doi.org/10.1039/D4CP02594F
  29. Do Hyung Kang, Masafumi Koga, Neal Haldar, Daniel M. Neumark. Dynamics of photoexcited 5-bromouracil and 5-bromo-2′-deoxyuridine studied by extreme ultraviolet time-resolved photoelectron spectroscopy in liquid flat jets. Chemical Science 2024, 15 (41) , 17245-17255. https://doi.org/10.1039/D4SC03920C
  30. Piao Xu, Jianduo Lu, Dongdong Wang, Duoduo Li, Jinyou Long, Song Zhang, Bing Zhang. Ultraviolet photoemission spectra of aqueous thymidine by liquid-microjet photoelectron spectroscopy: A combined experimental error assessment. Chinese Journal of Chemical Physics 2024, 37 (3) , 439-448. https://doi.org/10.1063/1674-0068/cjcp2402024
  31. Dennis Mayer, Fabiano Lever, Markus Gühr. Time‐resolved x‐ray spectroscopy of nucleobases and their thionated analogs. Photochemistry and Photobiology 2024, 100 (2) , 275-290. https://doi.org/10.1111/php.13903
  32. Davide Avagliano, Irene Conti, Mohsen M.T. El-Tahawy, Vishal K. Jaiswal, Artur Nenov, Marco Garavelli. Hybrid QM/MM Approach for the Calculation of Excited States in Complex Environments. 2024, 158-187. https://doi.org/10.1016/B978-0-12-821978-2.00059-3
  33. Yusuke Tsutsui. Fundamental Principles of Impedance Spectroscopy and its Biological Applications. 2022, 101-123. https://doi.org/10.1002/9781119886358.ch4
  34. Rocío Borrego-Varillas, Artur Nenov, Piotr Kabaciński, Irene Conti, Lucia Ganzer, Aurelio Oriana, Vishal Kumar Jaiswal, Ines Delfino, Oliver Weingart, Cristian Manzoni, Ivan Rivalta, Marco Garavelli, Giulio Cerullo. Tracking excited state decay mechanisms of pyrimidine nucleosides in real time. Nature Communications 2021, 12 (1) https://doi.org/10.1038/s41467-021-27535-7
  35. Irene Conti, Matteo Bonfanti, Artur Nenov, Ivan Rivalta, Marco Garavelli. Photo-Active Biological Molecular Materials: From Photoinduced Dynamics to Transient Electronic Spectroscopies. 2021, 77-142. https://doi.org/10.1007/978-3-030-57721-6_2
  36. Vishal K. Jaiswal, Javier Segarra-Martí, Marco Marazzi, Elena Zvereva, Xavier Assfeld, Antonio Monari, Marco Garavelli, Ivan Rivalta. First-principles characterization of the singlet excited state manifold in DNA/RNA nucleobases. Physical Chemistry Chemical Physics 2020, 22 (27) , 15496-15508. https://doi.org/10.1039/D0CP01823F
  37. Mohammadhassan Valadan, Enrico Pomarico, Bartolomeo Della Ventura, Felice Gesuele, Raffaele Velotta, Angela Amoresano, Gabriella Pinto, Majed Chergui, Roberto Improta, Carlo Altucci. A multi-scale time-resolved study of photoactivated dynamics in 5-benzyl uracil, a model for DNA/protein interactions. Physical Chemistry Chemical Physics 2019, 21 (48) , 26301-26310. https://doi.org/10.1039/C9CP03839F
  38. Toshinori Suzuki. Ultrafast photoelectron spectroscopy of aqueous solutions. The Journal of Chemical Physics 2019, 151 (9) https://doi.org/10.1063/1.5098402
  39. Jamie W. Riley, Bingxing Wang, Michael A. Parkes, Helen H. Fielding. Design and characterization of a recirculating liquid-microjet photoelectron spectrometer for multiphoton ultraviolet photoelectron spectroscopy. Review of Scientific Instruments 2019, 90 (8) https://doi.org/10.1063/1.5099040
  40. Alice Kunin, Daniel M. Neumark. Time-resolved radiation chemistry: femtosecond photoelectron spectroscopy of electron attachment and photodissociation dynamics in iodide–nucleobase clusters. Physical Chemistry Chemical Physics 2019, 21 (14) , 7239-7255. https://doi.org/10.1039/C8CP07831A
  41. Pengli Li, Jiadan Xue, Xuming Zheng. Decay dynamics of 6‐azauracil from the light absorbing S 2 (ππ*) state. Journal of Raman Spectroscopy 2019, 50 (3) , 345-359. https://doi.org/10.1002/jrs.5518
  42. Alice Kunin, Daniel M. Neumark. Femtosecond Time-Resolved Photoelectron Spectroscopy of Molecular Anions. 2019, 307-335. https://doi.org/10.1007/978-981-13-9371-6_11
  43. Gaurav Kumar, Anirban Roy, Ryan S. McMullen, Shanmukh Kutagulla, Stephen E. Bradforth. The influence of aqueous solvent on the electronic structure and non-adiabatic dynamics of indole explored by liquid-jet photoelectron spectroscopy. Faraday Discussions 2018, 212 , 359-381. https://doi.org/10.1039/C8FD00123E
  44. Inga Jordan, Arohi Jain, Thomas Gaumnitz, Jun Ma, Hans Jakob Wörner. Photoelectron spectrometer for liquid and gas-phase attosecond spectroscopy with field-free and magnetic bottle operation modes. Review of Scientific Instruments 2018, 89 (5) https://doi.org/10.1063/1.5011657
  45. Ana Julieta Pepino, Javier Segarra-Martí, Artur Nenov, Ivan Rivalta, Roberto Improta, Marco Garavelli. UV-induced long-lived decays in solvated pyrimidine nucleosides resolved at the MS-CASPT2/MM level. Physical Chemistry Chemical Physics 2018, 20 (10) , 6877-6890. https://doi.org/10.1039/C7CP08235E
  46. L. Martinez-Fernandez, T. Fahleson, P. Norman, F. Santoro, S. Coriani, R. Improta. Optical absorption and magnetic circular dichroism spectra of thiouracils: a quantum mechanical study in solution. Photochemical & Photobiological Sciences 2017, 16 (9) , 1415-1423. https://doi.org/10.1039/c7pp00105c
  47. François Dehez, Hugo Gattuso, Emmanuelle Bignon, Christophe Morell, Elise Dumont, Antonio Monari. Conformational polymorphism or structural invariance in DNA photoinduced lesions: implications for repair rates. Nucleic Acids Research 2017, 45 (7) , 3654-3662. https://doi.org/10.1093/nar/gkx148
  48. Sebastian Mai, Martin Richter, Philipp Marquetand, Leticia González. The DNA nucleobase thymine in motion – Intersystem crossing simulated with surface hopping. Chemical Physics 2017, 482 , 9-15. https://doi.org/10.1016/j.chemphys.2016.10.003
  49. Shuai Sun, Alex Brown. Effects of hydrogen bonding with H2O on the resonance Raman spectra of uracil and thymine. Computational and Theoretical Chemistry 2017, 1100 , 70-82. https://doi.org/10.1016/j.comptc.2016.12.014
  50. Hongjuan Wang, Xuebo Chen. How Does Thymine DNA Survive Ultrafast Dimerization Damage?. Molecules 2017, 22 (1) , 60. https://doi.org/10.3390/molecules22010060
  51. Javier Segarra-Martí, Antonio Francés-Monerris, Daniel Roca-Sanjuán, Manuela Merchán. Assessment of the Potential Energy Hypersurfaces in Thymine within Multiconfigurational Theory: CASSCF vs. CASPT2. Molecules 2016, 21 (12) , 1666. https://doi.org/10.3390/molecules21121666
  52. D. A. Horke, H. M. Watts, A. D. Smith, E. Jager, E. Springate, O. Alexander, C. Cacho, R. T. Chapman, R. S. Minns. Hydrogen Bonds in Excited State Proton Transfer. Physical Review Letters 2016, 117 (16) https://doi.org/10.1103/PhysRevLett.117.163002
  53. Vasilios G. Stavros, Jan R.R. Verlet. Gas-Phase Femtosecond Particle Spectroscopy: A Bottom-Up Approach to Nucleotide Dynamics. Annual Review of Physical Chemistry 2016, 67 (1) , 211-232. https://doi.org/10.1146/annurev-physchem-040215-112428
  54. Angelo Giussani, Javier Segarra-Martí, Artur Nenov, Ivan Rivalta, Alessandra Tolomelli, Shaul Mukamel, Marco Garavelli. Spectroscopic fingerprints of DNA/RNA pyrimidine nucleobases in third-order nonlinear electronic spectra. Theoretical Chemistry Accounts 2016, 135 (5) https://doi.org/10.1007/s00214-016-1867-z
  55. Bing Xue, Atsushi Yabushita, Takayoshi Kobayashi. Ultrafast dynamics of uracil and thymine studied using a sub-10 fs deep ultraviolet laser. Physical Chemistry Chemical Physics 2016, 18 (25) , 17044-17053. https://doi.org/10.1039/C5CP07861J
  56. Barbara Marchetti, Tolga N. V. Karsili, Michael N. R. Ashfold, Wolfgang Domcke. A ‘bottom up’, ab initio computational approach to understanding fundamental photophysical processes in nitrogen containing heterocycles, DNA bases and base pairs. Physical Chemistry Chemical Physics 2016, 18 (30) , 20007-20027. https://doi.org/10.1039/C6CP00165C
  57. XinZhong Hua, LinQiang Hua, XiaoJun Liu. The methyl- and aza-substituent effects on nonradiative decay mechanisms of uracil in water: a transient absorption study in the UV region. Physical Chemistry Chemical Physics 2016, 18 (20) , 13904-13911. https://doi.org/10.1039/C6CP00732E
  58. Katharina Röttger, Hugo J. B. Marroux, Hendrik Böhnke, David T. J. Morris, Angus T. Voice, Friedrich Temps, Gareth M. Roberts, Andrew J. Orr-Ewing. Probing the excited state relaxation dynamics of pyrimidine nucleosides in chloroform solution. Faraday Discussions 2016, 194 , 683-708. https://doi.org/10.1039/C6FD00068A
  59. Akira Nakayama. A QM/MM study of absorption spectra of uracil derivatives in aqueous solution. 2016, 020018. https://doi.org/10.1063/1.4968644
  60. I. Jordan, M. Huppert, M. A. Brown, J. A. van Bokhoven, H. J. Wörner. Photoelectron spectrometer for attosecond spectroscopy of liquids and gases. Review of Scientific Instruments 2015, 86 (12) https://doi.org/10.1063/1.4938175
  61. Franziska Buchner, Berit Heggen, Hans-Hermann Ritze, Walter Thiel, Andrea Lübcke. Excited-state dynamics of guanosine in aqueous solution revealed by time-resolved photoelectron spectroscopy: experiment and theory. Physical Chemistry Chemical Physics 2015, 17 (47) , 31978-31987. https://doi.org/10.1039/C5CP04394H

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2015, 137, 8, 2931–2938
Click to copy citationCitation copied!
https://doi.org/10.1021/ja511108u
Published February 11, 2015

Copyright © 2015 American Chemical Society. This publication is licensed under CC-BY.

Article Views

2810

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Electronic potential energies of low-lying states of isolated and hydrated thymine around the FC region. Excited-state energies for the gas phase and solvatochromic shift of the nπ* state are taken from ref 20, and the ππ* energy of the hydrated molecule is taken from experimental values. (15) Ionization energies for gas phase are taken from ref 22, and ionization energies of hydrated molecules are taken from ref 13. The absorbed photon energy, i.e., the sum of pump photon energy hνpu and probe photon energy hνpr, is given by the horizontal line. The molecular structure and the numbering of the atoms is shown in the right part of the figure.

    Figure 2

    Figure 2. Time-resolved photoelectron spectrum of thymine, globally analyzed by three spectral contributions. (a) Comparison between data (color) and global fit (contour lines), (b) decay associated spectra, (c) residuals (color) and global fit (contour lines), and (d) population dynamics of individual contributions. The color scale of the residuals covers a range of ±20% of the maximum photoelectron signal. Vertical lines in panel b mark the maximum kinetic energies for the individual contributions.

    Figure 3

    Figure 3. Potential energy differences between the ππ* state and the neutral ground state, the nπ* state and the ionic ground state along representative trajectories taken from ref 20 that reach (ππ*(C5–C6)/S0)CI (left) and (ππ*(C4–O8)/S0)CI (right) regions. Gray-shaded areas (between red and black solid lines) show ionization windows of ππ* state for instantaneous and complete redistribution of excess energy (E(ππ*) + ℏωpr) in case of 5.2 eV probe pulses. Diamonds mark the total absorbed energy at temporal overlap (ℏωpu + ℏωpr + E(S0; t = 0)).

    Figure 4

    Figure 4. (left) Average population of trajectories in the ππ* state that can be ionized for the given pump–probe time delay and photon energy (ℏωpr = 5.2 eV) (black line). The trajectories are divided into two groups: trajectories that reach the (ππ*(C5–C6)/S0)CI and (ππ*(C4–O8)/S0)CI regions, and the average populations of each group are shown in blue and red dotted lines. (right) Kinetic energies of photoelectrons along the trajectories. The trajectories are again divided into two groups and shown in blue and red circles. The averages of kinetic energy at t = 0 over all trajectories are calculated to be 1.72 eV. The averages over trajectories that reach the (ππ*(C5–C6)/S0)CI and (ππ*(C4–O8)/S0)CI regions are 1.80 and 1.40 eV, respectively.

    Figure 5

    Figure 5. Potential energy difference between the D(π–1) and nπ* states (red line), and the D(n–1) and nπ* states (blue line) along the QM/MM-MD trajectories that are equilibrated in the nπ* state using the SA-CASSCF(12,9) calculations. The energies are recalculated by CASPT2(12,9) and CASPT2(11,9) for the neutral and cationic molecules, respectively. The energies of the cationic states are shifted using the correction described in the text, and the experimentally applied photon energy (5.2 eV) is shown by the dotted line.

  • References


    This article references 41 other publications.

    1. 1
      Crespo-Hernández, C. E.; Cohen, B.; Hare, P. M.; Kohler, B. Chem. Rev. 2004, 104, 1977
    2. 2
      de Vries, M. S.; Hobza, P. Annu. Rev. Phys. Chem. 2007, 58, 585
    3. 3
      Middleton, C. T.; de La Harpe, K.; Su, C.; Law, Y. K.; Crespo-Hernández, C. E.; Kohler, B. Annu. Rev. Phys. Chem. 2009, 60, 217
    4. 4
      Kleinermanns, K.; Nachtigallová, D.; de Vries, M. S. Int. Rev. Phys. Chem. 2013, 32, 308
    5. 5
      Improta, R.; Barone, V. Top. Curr. Chem. 2014, 355, 329
    6. 6
      Abouaf, R.; Pommier, J.; Dunet, H. Chem. Phys. Lett. 2003, 381, 486
    7. 7
      Ullrich, S.; Schultz, T.; Zgierski, M. Z.; Stolow, A. Phys. Chem. Chem. Phys. 2004, 6, 2796
    8. 8
      Hudock, H. R.; Levine, B. G.; Thompson, A. L.; Satzger, H.; Townsend, D.; Gador, N.; Ullrich, S.; Stolow, A.; Martínez, T. J. J. Phys. Chem. A 2007, 111, 8500
    9. 9
      Szymczak, J. J.; Barbatti, M.; Hoo, J. T. S.; Adkins, J. A.; Windus, T. L.; Nachtigallová, D.; Lischka, H. J. Phys. Chem. A 2009, 113, 12686
    10. 10
      Merchán, M.; González-Luque, R.; Climent, T.; Serrano-Andrés, L.; Rodríguez, E.; Reguero, M.; Peláez, D. J. Phys. Chem. B 2006, 110, 26471
    11. 11
      Lan, Z.; Fabiano, E.; Thiel, W. J. Phys. Chem. B 2009, 113, 3548
    12. 12
      McFarland, B. K.; Farrell, J. P.; Miyabe, S.; Tarantelli, F.; Aguilar, A.; Berrah, N.; Bostedt, C.; Bozek, J. D.; Bucksbaum, P. H.; Castagna, J. C.; Coffee, R. N.; Cryan, J. P.; Fang, L.; Feifel, R.; Gaffney, K. J.; Glownia, J. M.; Martinez, T. J.; Mucke, M.; Murphy, B.; Natan, A.; Osipov, T.; Petrovíc, V. S.; Schorb, S.; Schultz, T.; Spector, L. S.; Swiggers, M.; Tenney, I.; Wang, S.; White, J. L.; White, W.; Gühr, M. Nat. Comm. 2014, 5, 4235
    13. 13
      Seidel, R. Electronic-Structure Interactions in Aqueous Solutions: A Liquid-Jet Photoelectron-Spectroscopy Study. Ph.D. Thesis; Technische Universitaet Berlin: Berlin, 2011.
    14. 14
      Gustavsson, T.; Sharonov, A.; Markovitsi, D. Chem. Phys. Lett. 2002, 351, 195
    15. 15
      Gustavsson, T.; Bányász, A.; Lazzarotto, E.; Markovitsi, D.; Scalmani, G.; Frisch, M. J.; Barone, V.; Improta, R. J. Am. Chem. Soc. 2006, 128, 607
    16. 16
      Pecourt, J.-M. L.; Peon, J.; Kohler, B. J. Am. Chem. Soc. 2001, 132, 10370
    17. 17
      Hare, P. M.; Crespo-Hernández, C. E.; Kohler, B. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 435
    18. 18
      Gustavsson, T.; Bányász, A.; Sarkar, N.; Markovitsi, D.; Improta, R. Chem. Phys. 2008, 350, 186
    19. 19
      Improta, R.; Barone, V.; Lami, A.; Santoro, F. J. Phys. Chem. B 2009, 113, 14491
    20. 20
      Nakayama, A.; Arai, G.; Yamazaki, S.; Taketsugu, T. J. Chem. Phys. 2013, 139, 214304
    21. 21
      Minezawa, N. J. Chem. Phys. 2014, 141, 164118
    22. 22
      Urano, S.; Yang, X.; LeBreton, P. R. J. Mol. Struct. 1989, 214, 315
    23. 23
      Buchner, F.; Lübcke, A.; Heine, N.; Schultz, T. Rev. Sci. Instrum. 2010, 81, 113107
    24. 24
      Preissler, N.; Buchner, F.; Schultz, T.; Lübcke, A. J. Phys. Chem. B 2013, 117, 2422
    25. 25
      Buchner, F.; Ritze, H.-H.; Lahl, J.; Lübcke, A. Phys. Chem. Chem. Phys. 2013, 15, 11402
    26. 26
      Wu, Y.; Tepper, H. L.; Voth, G. A. J. Chem. Phys. 2006, 124, 024503
    27. 27
      Werner, H. J.; Knowles, P. J.; Knizia, G.; Manby, F. R.; Celani, M. S. P.; Korona, T.; Lindh, R.; Mitrushenkov, A.; Rauhut, G.; Shamasundar, K. R.; Adler, T. B.; Amos, R. D.; Bernhardsson, A.; Berning, A.; Cooper, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.; Goll, E.; Hampel, C.; Hesselmann, A.; Hetzer, G.; Hrenar, T.; Jansen, G.; Köppl, C.; Liu, Y.; Lloyd, A. W.; Mata, R. A.; May, A. J.; McNicholas, S. J.; Meyer, W.; Mura, M. E.; Nicklaß, A.; O’Neill, D. P.; Palmieri, P.; Pflüger, K.; Pitzer, R.; Reiher, M.; Shiozaki, T.; Stoll, H.; Stone, A. J.; Tarroni, R.; Thorsteinsson, T.; Wang, M.; Wolf, A.MOLPRO version 2010.1 a package of ab initio programs; Birmingham, UK.
    28. 28
      Zechmann, G.; Barbatti, M. J. Phys. Chem. A 2008, 112, 8273
    29. 29
      Etinski, M.; Marian, C. M. Phys. Chem. Chem. Phys. 2010, 12, 4915
    30. 30
      Tao, H.; Levine, B. G.; Martínez, T. J. J. Phys. Chem. A 2009, 113, 13656
    31. 31
      Schreiber, M.; Silva-Junior, M. R.; Sauer, S. P. A.; Thiel, W. J. Chem. Phys. 2008, 128, 134110
    32. 32

      See the Supporting Information.

    33. 33
      Onidas, D.; Markovitsi, D.; Marguet, S.; Sharonov, A.; Gustavsson, T. J. Phys. Chem. B 2002, 106, 11367
    34. 34
      Konovalov, V. V.; Raitsimring, A. M.; Tsvetkov, Y. D. Int. J. Rad. Appl. Instr. C Rad. Phys. Chem. 1988, 32, 623
    35. 35
      Plante, I.; Cucinotta, F. A. New J. Phys. 2009, 11, 063047
    36. 36
      Buchner, F.; Schultz, T.; Lübcke, A. Phys. Chem. Chem. Phys. 2012, 14, 5837
    37. 37
      Suzuki, Y.-I.; Nishizawa, K.; Kurahashi, N.; Suzuki, T. Phys. Rev. E 2014, 90, 010302(R)
    38. 38
      Matsika, S. J. Phys. Chem. A 2005, 109, 7538
    39. 39
      Kistler, K. A.; Matsika, S. J. Chem. Phys. 2008, 128, 215102
    40. 40
      González-Vázquez, J.; González, L. Chem. Phys. Chem. 2010, 11, 3617
    41. 41
      Satzger, H.; Townsend, D.; Zgierski, M. Z.; Patchkovskii, S.; Ullrich, S.; Stolow, A. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 10196
  • Supporting Information

    Supporting Information


    Details on the detection efficiency at low kinetic energy, the role of inelastic processes, global analysis, time-resolved photoelectron spectra, global analysis results for thymine and thymidine, an extended time scan for thymine, and a discussion of alternative signal assignment. This material is available free of charge via the Internet at http://pubs.acs.org.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.