Intersystem Crossing Pathways in the ... - ACS Publications

May 11, 2016 - Intersystem Crossing Pathways in the Noncanonical Nucleobase 2-Thiouracil: A Time-Dependent Picture. Sebastian Mai, Philipp Marquetand,...
0 downloads 12 Views 480KB Size
Subscriber access provided by University of Pennsylvania Libraries

Letter

Intersystem Crossing Pathways in the Noncanonical Nucleobase 2-Thiouracil: A Time-Dependent Picture Sebastian Mai, Philipp Marquetand, and Leticia González J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.jpclett.6b00616 • Publication Date (Web): 11 May 2016 Downloaded from http://pubs.acs.org on May 14, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Intersystem Crossing Pathways in the Noncanonical Nucleobase 2-Thiouracil: A Time-Dependent Picture Sebastian Mai, Philipp Marquetand,∗ and Leticia González∗ Institute of Theoretical Chemistry, University of Vienna, Währinger Straße 17, 1090 Vienna, Austria E-mail: [email protected]; [email protected]

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract The deactivation mechanism after UV irradiation of 2-thiouracil has been investigated using nonadiabatic dynamics simulations at the MS-CASPT2 level of theory. It is found that after excitation the S2 quickly relaxes to S1 and from there intersystem crossing takes place to both T2 and T1 with a time constant of 400 fs and a triplet yield above 80%, in very good agreement with recent femtosecond experiments in solution. Both indirect S1 → T2 → T1 and direct S1 → T1 pathways contribute to intersystem crossing, with the former being predominant. The results contribute to the understanding of how some noncanonical nucleobases respond to harmful ultraviolet light, which could be relevant for prospective photo-chemotherapeutic applications.

Graphical TOC Entry

2

ACS Paragon Plus Environment

Page 2 of 20

Page 3 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Noncanonical nucleobases, i.e., DNA or RNA nucleobases with chemical modifications, are gaining increased attention as researchers are realizing that these modifications provide additional possibilities to encode genetic information. 1 One important chemical alteration is thionation, where the oxygen atom of a carbonyl group is substituted by a sulfur atom, yielding a thiobase. Such thiobases have been isolated from transfer-RNAs 2–4 from all domains of life, 5 but they are also commonly found in artificial DNA and RNA systems. They prevent mispairing in DNA-based nanotechnology, 6,7 are employed as fluorescent markers, 8–10 and— given their excellent binding affinity to biological targets—they are also used clinically as cytotoxic agents, 11,12 as anti-inflammatory drugs, 13 or as antithyroid drugs. 14,15 Particularly due to their medicinal relevance in photo-chemotherapeutic applications, 16–21 unravelling the photophysical and photochemical properties of thiobases is decisive. The ability of producing high triplet yields is pivotal to produce singlet oxygen and other reactive species responsible for the cytotoxicity. This ability seems to be linked to the heavy-atom effect of the sulfur atom, since canonical nucleobases exhibit ultrafast (i.e., few picosecond) internal conversion back to the electronic ground state, 22,23 while thiobases do not. 24 Understanding the photochemistry of noncanonical nucleobases in general is also important in the context of molecular evolution in prebiotic settings, where only the most photostable molecules survived the harsh ultraviolet irradiation on the early Earth. 22,25,26 A number of recent studies has focused on the excited-state dynamics of thiouracils but a consensus regarding the mechanism to populate triplet states has not been reached yet. Time-resolved experimental studies show ultrafast near-unity triplet yield for 2-thiouracil (2TU), 2-thiothymine, 2-thiouridine, and 2,4-dithiouracil. 21,27–31 Stationary quantum-chemical calculations are only available for 2TU, but different relaxation pathways are proposed to explain intersystem crossing (ISC). 32–34 After populating the lowest bright state S2 , all studies agree that internal conversion to the S1 should take place. As a subsequent deactivation pathway, two routes for ISC have been proposed: one direct S1 → T1 and involving mostly planar geometries 32,33 and one indirect S1 → T2 → T1 with geometries showing a large degree

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

O8 H11

N3

6

H12

H10

4 5

2

N1

S7

H9

Figure 1: Structure of 2TU and atom numbering. of pyramidalization. 34 Since stationary calculations cannot predict the relative importance of these routes and might even overlook additional pathways, a time-dependent study is mandatory. Here, we investigate the dynamics of the excited states of 2TU (Figure 1), with the aim of providing a clear-cut deactivation mechanism that can also help to enlighten the general photophysics of thiobases and other noncanonical nucleobases, which is a timely topic. 21,24,30,35–42 Accurate CASPT2-based surface hopping molecular dynamics simulations including singlet and triplet states have been performed, employing a local development version of the Sharc (Surface Hopping including ARbitrary Couplings) software suite. 43–45 Ab initio methods regularly employed for excited-state dynamics simulations, such as CASSCF or TD-DFT, often give a qualitatively inaccurate description of the potential energy landscape. 46–50 Hence, here we rely on the MS-CASPT2(12,9)/cc-pVDZ level of theory (multi-state complete active space perturbation theory) for all underlying on-the-fly computations of energies, numerical gradients and relevant couplings between the electronic states. While this is not the first application of MS-CASPT2 in excited-state dynamics simulations (pioneered by the group of Martínez 51 ), here we employ the largest MS-CASPT2 setup so far, to the best of our knowledge. The usage of the MS variant of CASPT2 for dynamics is necessary, as the single-state variant fails to correctly describe excited-state crossings. 51,52 Although MS-CASPT2 can create state-oversplitting artifacts, 53,54 the good agreement between single-state and multi-state CASPT2 static calculations for 2TU 34 suggests that these artifacts are not present here. As will be shown below, this computationally expensive approach achieves very good agreement

4

ACS Paragon Plus Environment

Page 4 of 20

Page 5 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

with the available experimental ISC time scales. 21,30,42 The orbitals included in the active space (Figure S1) and other details of the calculations as well as a discussion regarding the quality and stability of the computations can be found in the Supporting Information (SI). To provide initial conditions for the dynamics simulations, first an absorption spectrum (see Figure S2 of SI) was computed from 2000 geometries sampled from the ground state Wigner distribution. The single absorption band obtained shows a maximum at 4.2 eV, almost exclusively due to excitation to the S2 state, 34 which is a ππ ∗ excitation from the sulfur atom to one π ∗ antibonding orbital of the pyrimidine ring, denoted as πS π6∗ henceforth (see orbital labels in Figure S1). The lower-lying S1 state is of nS π2∗ character and does not contribute much to the absorption. Higher-lying excited states were not computed for the absorption spectrum as we are focused on the dynamics after excitation to S2 . Accordingly, the dynamics simulations are initiated in the bright S2 state and consider the interaction with all states below, i.e., the S1 and S0 as well as the three lowest triplet states, T1 , T2 , and T3 (3 πS π2∗ , 3 nS π2∗ , 3 π5 π6∗ at the Franck-Condon geometry). A total of 44 trajectories with a simulation time of 1 ps have been considered; the nuclear dynamics was simulated with a 0.5 fs time step and the electronic wavefunction was propagated with a 0.02 fs time step using the “local diabatization” formalism. 55,56 Figure 2 (a) shows the time-evolution of the excited-state populations (thin lines) and the results of a global fit described below (thick lines). For all analysis purposes, the populations of T2 and T3 are merged, as the T3 population is only 2% on average. As can be seen, after excitation to the S2 state, internal conversion to the S1 state is basically completed within 200 fs. Simultaneously, ISC to all three triplet states proceeds on a timescale of a few hundred femtoseconds, whereas hardly any population of the ground state is observed within 1 ps. The triplet population at 1 ps is 80%, with most of the remaining population still residing in S1 . This dynamical behaviour is different from the canonical nucleobase uracil, where ISC is smaller and ground state recovery is ultrafast. 22,23 To identify the most relevant deactivation pathways and branching ratios, the net pop-

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2: (a) Excited-state populations from 44 trajectories (thin lines) and global fits (thick lines) to the populations. (b) Net population transfer among the electronic states. (c) Fitted time constants (in fs) from the global fit. The T3 population was absorbed into the T2 population in all panels. The thickness of the lines in (b) and (c) relates to the amount of transferred population. ulation transfer among the excited states is plotted in Figure 2 (b). Three different ISC routes are revealed: S2 → T2,3 , S1 → T2,3 and S1 → T1 . The S1 → T2 route (ignoring the T3 population henceforth), proposed in Ref. 34 , accounts for about 40% of the ISC events, and is thus the main ISC pathway. However, the routes S1 → T1 and S2 → T2 , suggested in earlier works, 32,33 account for 35% and 25% of ISC, respectively. Time scales associated to the excited-state populations in Figure 2 (a) could in principle be derived from a global fit to a kinetic model based on the transition graph in Figure 2 (b). However, the graph is relatively complex as it contains parallel pathways. In terms of chemical reaction network theory, 57 the stochiometry matrix corresponding to the transfer

6

ACS Paragon Plus Environment

Page 6 of 20

Page 7 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

graph exhibits a two-dimensional nullspace. Intuitively, this means that several different combinations of rate constants can lead to the same overall population dynamics. Stated differently, for the given transition graph the populations do not contain enough information to uniquely estimate all individual rate constants. Therefore, attempting to extract the individual transition rates for the three ISC routes by fitting will not be successful as the different ISC rate constants would be strongly correlated and large fitting errors and essentially arbitrary ISC rates would be obtained. Instead, for the purpose of obtaining time scales, a simplified kinetic model was devised, where the S1 → T2 → T1 pathway is merged into the competing S1 → T1 pathway. A global fit based on this model (see the SI for more details) yields the time constants presented in Figure 2 (c) and the corresponding fitted functions shown as thick lines in Figure 2 (a), which closely follow the actual populations. The error estimates given in Figure 2 (c) were obtained with the bootstrapping method. 58 They describe mainly the uncertainty in the populations due to incomplete sampling (i.e., due to the small number of trajectories); systematic errors due to the level of theory and the semi-classical nature of the dynamics method are not included. Consequently, the errors serve as a check as to whether the number of trajectories is sufficient. According to the simplified kinetic model, internal conversion from S2 → S1 proceeds with a time constant of 59 fs. The S1 → S0 deactivation is a much slower process that is not well described here because there are too few hops to the ground state. In addition, two ISC time constants were obtained, one of 540 ± 100 fs for S1 → T1 and another of 250 ± 60 fs for S2 → T2 . Both values can be combined to obtain an average ISC time constant of about 400 fs, which is in very good agreement with the values of 340 to 360 fs reported in aqueous and acetonitrile solution using transient absorption spectroscopy, 21,30 and also with the 330 fs obtained for 2-thiothymine. 28 Our ISC time constant also agrees reasonably well with the value of 775 fs from very recent multiphoton time-resolved photoelectron spectroscopy experiments, 42 considering that our simulations do not explicitly account for the multiphoton ionization probe. It has been

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

Trajectory Density (arb.u.) max

a721 (◦ )

p7513 (◦ )

p7213 (◦ )

rC=C (˚ A)

0

µ (D)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 20

(a)

1.50 1.45 1.40 1.35 60 30 0 −30 −60 40 20 0 −20 −40

S C N

N

(b)

C

rCC C

O

C

S O p

N

C

N

C

C

(c)

C

S O C

N p

C

N

C

C

(d)

130 120 110 100 90 7 6 5 4 3

O C

N

C S

C a

N

C

(e) µ(S0 ) 0

200 400 600 800 Propagation time (fs)

1000

Figure 3: Time-dependent evolution of selected internal coordinates in the ensemble of 44 trajectories. The definition of the chosen internal coordinates is schematically depicted on the right (see also SI). Panel (e) shows the evolution of the dipole moment of the active state (contour plot) and the ground state dipole moment µ(S0 ) (dashed line). shown that the inclusion of the time-dependent ionization probabilities in the simulations can easily change the timescales by a factor of two. 59 The kinetic model can also be used to extrapolate the populations to times longer than 1 ps. Including the errors associated to the time constants, a total triplet yield of 90 ± 10% has been obtained, in agreement with the experimental value of 75 ± 20%, 30 and also with the ones of 90 ± 10% 28 or 100 ± 5% 29 for the related 2-thiothymine. With the aim of identifying important geometrical changes along the excited-state dynamics, the time-evolution of selected internal coordinates is displayed in Figure 3. After excitation in the Franck-Condon region, relaxation to two different S2 minima is possible ac-

8

ACS Paragon Plus Environment

Page 9 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

cording to stationary calculations. 32–34 These minima differ in structure and the occupation of the antibonding orbitals, either more localized at the C6 atom 32–34 or C2 atom 34 (recall Figure 1), and therefore labelled as 1 πS π2∗ and 1 πS π6∗ minima, respectively. Which of the minima is actually populated is easily solved by the dynamics simulations. The length of the C5 =C6 bond length shown in Figure 3 (a) undoubtedly reveals that most trajectories exhibit a bond length of 1.43 Å during the first 50 fs, which corresponds to the planar 1 πS π6∗ minimum. In contrast, only two trajectories moved into the pyramidalized minimum (with bond length 1.38Å) of the 1 πS π2∗ state, being trapped there for several hundreds of fs, as there exist sizeable barriers to leave this minimum. Accordingly, the S2 /S1 (1 πS π2∗ /1 nS π2∗ ) conical intersection predicted in Ref. 34 is not operative. In contrast, since the 1 πS π6∗ minimum is located very close to another S2 /S1 conical intersection involving the (1 πS π6∗ /1 nS π2∗ ) states, it effectively funnels the population towards the S1 state, leading to an S2 lifetime of 59 fs. Subsequently, the bond length distribution shifts to smaller values and for the rest of the simulation, the trajectories spent most of the time around 1.38Å, which corresponds to the 1

nS π2∗ pyramidalized minimum in the S1 state (as well as to other minima, see below). At the operative S2 /S1 crossing, both the pyrimidine ring and the thiocarbonyl group

are planar, but at the S1 minimum the S atom is displaced out of the ring plane. This strong pyramidalization at the C2 atom is characteristic of the 1 nS π2∗ minimum and can be quantified by the pyramidalization angles p7213 (defined as the angle between the S7 = C2 bond and the C2 − N1 − N3 plane, see Figure 3 (b), and related to the planarity of the thiocarbonyl group) and p7513 (that describes how much the S atom is displaced from the molecular plane, see Figure 3 (c)). Both pyramidalization angles are initially distributed around zero but within 100 fs an equilibrium value of p7213 = ±40◦ is reached, whereas it takes more than 400 fs for p7513 to stabilize. Inspection of the trajectories (see also Figure S3 in the SI) confirms that the pyramidalization mechanism involves two stages. First, the system relaxes by puckering the C2 atom out of the ring plane, and thereby stabilizing the energy of the 1 nS π2∗ state by about 0.35 eV. Since only the C2 atom needs to be moved, but

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

not the heavy sulfur atom, puckering only takes tens of fs after transition to S1 (Figure 3 (b)). In a second stage, the pyramidalizations at the nitrogen atoms N1 and N3 are removed and the planarity of the ring recovered while the pyramidalization at C2 is conserved. Since this second stage necessitates heavy atom motion, it is far slower, taking several hundred fs (Figure 3 (c)); this stage leads only to a marginal energetic stabilization (0.03 eV). After relaxation to the pyramidalized S1 minimum, ISC proceeds efficiently, promoted by small energy gaps between the three states S1 , T2 and T1 as well as sizeable spin-orbit couplings (up to 150 cm−1 ). 34 Based on the trajectories, the “wagging” mode of the S atom has been identified as an important tuning mode that modulates the gap between 1 nS π2∗ and 3

πS π2∗ states. This mode is quantified with the angle α721 , sketched in Figure 3 (d). This

angle affects the energy of the 3 πS π2∗ state more than that of 1 nS π2∗ and 3 nS π2∗ (Figure S4 in the SI). Thus, at angles of about 102◦ and 128◦ the 3 πS π2∗ state crosses with the 1 nS π2∗ state, at energies only slightly above the minimum energy of the S1 minimum energy. At these two angles one can also see that the two triplet characters are exchanged, suggesting a nearby T2 /T1 conical intersection that favours very efficient internal conversion between both triplet states. These two angles are regularly reached by the trajectories (Figure 3 (d)), providing ample opportunity for ISC and subsequent internal conversion to T1 to take place. After the transition to T1 , the system equilibrates between two distinct local T1 minima, termed “pyramidalized” and “boat” in Ref. 34 In Figure 3 (a), most trajectories exhibit values of the C5 =C6 bond length of around 1.38Å, which is typical for the pyramidalized T1 minimum as well as the S1 minimum. A small fraction of the ensemble shows values around 1.49Å, corresponding to a population of the boat T1 minimum. Pyramidalization angles p7213 (Figure 3 (b)) around ±40◦ indicate the pyramidalized T1 minimum and values close to zero indicate a population of the boat T1 minimum. Detailed examination of the individual trajectories showed that the average population ratio in the two T1 minima is approximately 2:1, in favor of the pyramidalized minimum. Internal conversion to the ground state is minimal, in line with experimental evidence of

10

ACS Paragon Plus Environment

Page 10 of 20

Page 11 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

high ISC yields and phosphorescence. 30,60 In agreement to this, none of the reported S1 /S0 conical intersections 32–34 is observed. Finally, Figure 3 (e) presents the temporal evolution of the permanent dipole moment of the active state. After excitation to S2 , its value is about 6.4 Debye (D), whereas after relaxation to the lower states it quickly changes to 3.8 D and then does not change notably. This implies that the involved excited states (S1 , T1 , T2 ) have very similar dipole moments, which are also very close to the value of the ground state (4.2 D, see also Ref. 34 ). This fact has important consequences for the photochemical behaviour of 2TU in polar solvents, where more polar states tend to be stabilized while unpolar states increase in energy. In the present case, a polar solvent would decrease the energy of S2 (πS π6∗ ), further enhancing S2 → S1 internal conversion, and would leave the remaining states—and hence the remaining dynamics—mostly unchanged. This could offer an explanation for the good agreement of our simulations and time-resolved experiments in solution, 30 even though the simulations neglect solvation effects. This hypothesis is also supported by the very similar absorption maxima reported for different solvents in the literature (see Table S1 in the SI). In this regard, 2TU seems to be a rare exception from the observation that solvation strongly affects the dynamics of nucleobases. 22,23,26 Especially in uracil, the reported ISC yields are strongly solvent-dependent, and vary from a few percent in water to about 50% in aprotic, non-polar solvents. 61,62 The Sharc method, using CASSCF potentials, yields for uracil in the gas phase an ISC yield of 67% (extrapolated from the data reported in Ref. 63 ). In conclusion, nonadiabatic dynamics simulations show that the mechanism of ISC is very efficient in the noncanonical nucleobase 2-thiouracil. The global sketch shown in Figure 4 summarizes the deactivation mechanism. After excitation to the first bright state S2 of 1

πS π ∗ character, the system quickly moves to a mostly planar conical intersection with the

S1 (1 nS π ∗ ) state and decays with a time constant of 59 fs to S1 . Internal conversion from S2 to S1 mainly involves bond length inversion followed by pyramidalization. Almost no

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4: Summary of the excited-state dynamics of 2-thiouracil, showing excitation (hν), internal conversion in the singlet states (54 fs time constant) and intersystem crossing (290 fs). ground state relaxation was observed; instead, extremely efficient ISC from S1 to T2 and T1 occurs, with an average time constant of 400 fs, in agreement with time-resolved experiments in solution. 21,30 The most important ISC channel is the indirect S1 → T2 → T1 one, with S1 → T1 and S2 → T2 playing smaller roles. ISC involves strong pyramidalization of the thiocarbonyl group and is promoted by the wagging mode of the same group, which efficiently modulates the gap between 1 nS π ∗ (S1 ) and 3 πS π ∗ states. Our extrapolated triplet yield is 90 ± 10%, which agrees nicely with the near-unity yields of several experimental studies. 28–30 These results are not only relevant for 2TU but put in evidence how DNA and RNA modifications strongly influence their response to UV light. For example, thionation can abate the photoprotective abilities inherent to the canonical nucleobases, unless relaxation from the reactive triplet states to the ground state takes place sufficiently fast. However, the ability to populate efficiently triplet states can be exploited to generate singlet oxygen with therapeutic aims, as has been experimentally shown in thiouracils. 21,29

Acknowledgement Funding from the Austrian Science Fund (FWF), project P25827, and very generous allocation of computer ressources at the Vienna Scientific Cluster 3 (VSC3) are gratefully ac12

ACS Paragon Plus Environment

Page 12 of 20

Page 13 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

knowledged. We also thank Susanne Ullrich, Marvin Pollum, and Carlos Crespo-Hernández for stimulating discussions.

Supporting Information Available Computational and global fitting details. Simulated absorption spectrum. Potential energy scans. This material is available free of charge via the Internet at http://pubs.acs.org/.

References (1) Carell, T.; Brandmayr, C.; Hienzsch, A.; Müller, M.; Pearson, D.; Reiter, V.; Thoma, I.; Thumbs, P.; Wagner, M. Structure and Function of Noncanonical Nucleobases. Angew. Chem. Int. Ed. 2012, 51, 7110–7131. (2) Carbon, J.; David, H.; Studier, M. H. Thiobases in Escherchia coli Transfer RNA: 2-Thiocytosine and 5-Methylaminomethyl-2-thiouracil. Science 1968, 161, 1146–1147. (3) Ajitkumar, P.; Cherayil, J. D. Thionucleosides in Transfer Ribonucleic Acid: Diversity, Structure, Biosynthesis, and Function. Microbiol. Rev. 1988, 52, 103–113. (4) Zhang, S.; Blain, J. C.; Zielinska, D.; Gryaznov, S. M.; Szostak, J. W. Fast and Accurate Nonenzymatic Copying of an RNA-like Synthetic Genetic Polymer. Proc. Natl. Acad. Sci. 2013, 110, 17732–17737. (5) Motorin, Y.; Helm, M. RNA Nucleotide Methylation. WIREs RNA 2011, 2, 611–631. (6) Wojciechowski, F.; Leumann, C. J. Alternative DNA Base-Pairs: from Efforts to Expand the Genetic Code to Potential Material Applications. Chem. Soc. Rev. 2011, 40, 5669–5679.

13

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(7) Sintim, H. O.; ; Kool, E. T. Enhanced Base Pairing and Replication Efficiency of Thiothymidines, Expanded-size Variants of Thymidine. J. Am. Chem. Soc. 2006, 128, 396–397. (8) Favre, A.; Saintomé, C.; Fourrey, J.-L.; Clivio, P.; Laugâa, P. Thionucleobases as Intrinsic Photoaffinity Probes of Nucleic Acid Structure and Nucleic Acid-Protein Interactions. J. Photoch. Photobio. B 1998, 42, 109 – 124. (9) Wilhelmsson, L. M. Fluorescent Nucleic Acid Base Analogues. Q. Rev. Biophys. 2010, 43, 159–183. (10) Sinkeldam, R. W.; Greco, N. J.; Tor, Y. Fluorescent Analogs of Biomolecular Building Blocks: Design, Properties, and Applications. Chem. Rev. 2010, 110, 2579–2619. (11) Périgaud, C.; Gosselin, G.; Imbach, J. L. Nucleoside Analogues as Chemotherapeutic Agents: A Review. Nucleos. Nucleot. 1992, 11, 903–945. (12) Jordheim, L. P.; Durantel, D.; Zoulim, F.; Dumontet, C. Advances in the Development of Nucleoside and Nucleotide Analogues for Cancer and Viral Diseases. Nat. Rev. Drug Discovery 2013, 12, 447–464. (13) Lennard, L. The Clinical Pharmacology of 6-Mercaptopurine. Eur. J. Clin. Pharmacol. 1992, 43, 329–339. (14) Anderson, G. W.; Halverstadt, I. F.; Miller, W. H.; Jr., R. O. R. Studies in Chemotherapy. X. Antithyroid Compounds. Synthesis of 5- and 6- Substituted 2-Thiouracils from β-Oxoesters and Thiourea. J. Am. Chem. Soc. 1945, 67, 2197–2200. (15) Cooper, D. S. Antithyroid Drugs. N. Engl. J. Med. 2005, 352, 905–917. (16) Euvrard, S.; Kanitakis, J.; Claudy, A. Skin Cancers after Organ Transplantation. N. Engl. J. Med. 2003, 348, 1681–1691.

14

ACS Paragon Plus Environment

Page 14 of 20

Page 15 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(17) Massey, A.; Xu, Y.-Z.; Karran, P. Photoactivation of DNA Thiobases as a Potential Novel Therapeutic Option. Curr. Biol. 2001, 11, 1142 – 1146. (18) Karran, P.; Attard, N. Thiopurines in Current Medical Practice: Molecular Mechanisms and Contributions to Therapy-Related Cancer. Nat. Rev. Cancer 2008, 8, 24. (19) Reelfs, O.; Karran, P.; Young, A. R. 4-Thiothymidine Sensitization of DNA to UVA Offers Potential for a Novel Photochemotherapy. Photochem. Photobiol. Sci. 2012, 11, 148–154. (20) Pridgeon, S. W.; Heer, R.; Taylor, G. A.; Newell, D. R.; O’Toole, K.; Robinson, M.; Xu, Y.-Z.; Karran, P.; Boddy, A. V. Thiothymidine Combined with UVA as a Potential Novel Therapy for Bladder Cancer. Br. J. Cancer 2011, 104, 1869. (21) Pollum, M.; Jokusch, S.; Crespo-Hernández, C. Increase in the Photoreactivity of Uracil Derivatives by Doubling Thionation. Phys. Chem. Chem. Phys. 2015, 17, 27851–27861. (22) Crespo-Hernández, C. E.; Cohen, B.; Hare, P. M.; Kohler, B. Ultrafast Excited-State Dynamics in Nucleic Acids. Chem. Rev. 2004, 104, 1977–2020. (23) Middleton, C. T.; de La Harpe, K.; Su, C.; Law, Y. K.; Crespo-Hernández, C. E.; Kohler, B. DNA Excited-State Dynamics: From Single Bases to the Double Helix. Ann. Rev. Phys. Chem. 2009, 60, 217–239. (24) Pollum, M.; Martínez-Fernández, L.; Crespo-Hernández, C. E. In Photoinduced Phenomena in Nucleic Acids I ; Barbatti, M., Borin, A. C., Ullrich, S., Eds.; Topics in Current Chemistry; Springer Berlin Heidelberg, 2015; Vol. 356; pp 245–327. (25) Sobolewski, A.; Domcke, W. On the Mechanism of Nonradiative Decay of DNA Bases: Ab Initio and TDDFT Results for the Excited States of 9H-Adenine. Euro. Phys. J. D 2002, 20, 369–374.

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(26) Kleinermanns, K.; Nachtigallová, D.; de Vries, M. S. Excited State Dynamics of DNA Bases. Int. Rev. Phys. Chem. 2013, 32, 308–342. (27) Vendrell-Criado, V.; Saez, J. A.; Lhiaubet-Vallet, V.; Cuquerella, M. C.; Miranda, M. A. Photophysical Properties of 5-Substituted 2-Thiopyrimidines. Photochem. Photobiol. Sci. 2013, 12, 1460–1465. (28) Taras-Goślińska, K.; Burdziński, G.; Wenska, G. Relaxation of the T1 Excited State of 2-Thiothymine, its Riboside and Deoxyriboside-Enhanced Nonradiative Decay Rate Induced by Sugar Substituent. J. Photoch. Photobio. A 2014, 275, 89–95. (29) Kuramochi, H.; Kobayashi, T.; Suzuki, T.; Ichimura, T. Excited-State Dynamics of 6-Aza-2-thiothymine and 2-Thiothymine: Highly Efficient Intersystem Crossing and Singlet Oxygen Photosensitization. J. Phys. Chem. B 2010, 114, 8782–8789. (30) Pollum, M.; Crespo-Hernández, C. E. Communication: The Dark Singlet State as a Doorway State in the Ultrafast and Efficient Intersystem Crossing Dynamics in 2Thiothymine and 2-Thiouracil. J. Chem. Phys. 2014, 140, 071101. (31) Pollum, M.; Jockusch, S.; Crespo-Hernández, C. E. 2,4-Dithiothymine as a Potent UVA Chemotherapeutic Agent. J. Am. Chem. Soc. 2014, 136, 17930–17933. (32) Cui, G.; Fang, W.-h. State-Specific Heavy-Atom Effect on Intersystem Crossing Processes in 2-Thiothymine: A Potential Photodynamic Therapy Photosensitizer. J. Chem. Phys. 2013, 138, 044315. (33) Gobbo, J. P.; Borin, A. C. 2-Thiouracil Deactivation Pathways and Triplet States Population. Comput. Theor. Chem. 2014, 1040-1041, 195–201. (34) Mai, S.; Marquetand, P.; González, L. A Static Picture of the Relaxation and Intersystem Crossing Mechanisms of Photoexcited 2-Thiouracil. J. Phys. Chem. A 2015, 119, 9524–9533. 16

ACS Paragon Plus Environment

Page 16 of 20

Page 17 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

(35) Matsika, S. Photoinduced Phenomena in Nucleic Acids I ; Topics in Current Chemistry; Springer Berlin Heidelberg, 2014; Vol. 355; pp 209–243. (36) Gustavsson, T.; Bányász, A.; Lazzarotto, E.; Markovitsi, D.; Scalmani, G.; Frisch, M. J.; Barone, V.; Improta, R. Singlet Excited-State Behavior of Uracil and Thymine in Aqueous Solution: A Combined Experimental and Computational Study of 11 Uracil Derivatives. J. Am. Chem. Soc. 2006, 128, 607–619, PMID: 16402849. (37) Gengeliczki, Z.; Callahan, M. P.; Svadlenak, N.; Pongor, C. I.; Sztaray, B.; Meerts, L.; Nachtigallova, D.; Hobza, P.; Barbatti, M.; Lischka, H. et al. Effect of Substituents on the Excited-State Dynamics of the Modified DNA Bases 2,4-Diaminopyrimidine and 2,6-Diaminopurine. Phys. Chem. Chem. Phys. 2010, 12, 5375–5388. (38) Serrano-Andrés, L.; Merchán, M. Are the Five Natural DNA/RNA Base Monomers a Good Choice from Natural Selection? A Photochemical Perspective. J. Photochem. Photobiol. C: Photochem. Rev. 2009, 10, 21 – 32. (39) Martínez-Fernández, L.; González, L.; Corral, I. An Ab Initio Mechanism for Efficient Population of Triplet States in Cytotoxic Sulfur Substituted DNA Bases: The Case of 6-Thioguanine. Chem. Commun. 2012, 48, 2134–2136. (40) Martínez-Fernández, L.; Corral, I.; Granucci, G.; Persico, M. Competing Ultrafast Intersystem Crossing and Internal Conversion: a Time Resolved Picture for the Deactivation of 6-Thioguanine. Chem. Sci. 2014, 5, 1336–1347. (41) Cui, G.; Thiel, W. Intersystem Crossing Enables 4-Thiothymidine to Act as a Photosensitizer in Photodynamic Therapy: An Ab Initio QM/MM Study. J. Phys. Chem. Lett. 2014, 5, 2682–2687. (42) Yu, H.; Sánchez-Rodríguez, J. A.; Pollum, M.; Crespo-Hernández, C. E.; Mai, S.; Marquetand, P.;

González, L.;

Ullrich, S. Internal Conversion And Intersys-

tem Crossing Pathways In UV Excited, Isolated Uracils And Their Implications 17

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

In Prebiotic Chemistry. Phys. Chem. Chem. Phys. 2016, accepted manuscript. DOI:10.1039/C6CP01790H. (43) Richter, M.; Marquetand, P.; González-Vázquez, J.; Sola, I.; González, L. SHARC: ab Initio Molecular Dynamics with Surface Hopping in the Adiabatic Representation Including Arbitrary Couplings. J. Chem. Theory Comput. 2011, 7, 1253–1258. (44) Mai, S.; Marquetand, P.; González, L. A General Method to Describe Intersystem Crossing Dynamics in Trajectory Surface Hopping. Int. J. Quantum Chem. 2015, 115, 1215. (45) Mai, S.; Richter, M.; Ruckenbauer, M.; Oppel, M.; Marquetand, P.; González, L. SHARC: Surface Hopping Including Arbitrary Couplings – Program Package for NonAdiabatic Dynamics. sharc-md.org, 2014. (46) Serrano-Andrés, L.; Merchán, M. Quantum Chemistry of the Excited State: 2005 Overview. J. Mol. Struc. - THEOCHEM 2005, 729, 99 – 108. (47) Yamazaki, S.; Taketsugu, T. Nonradiative Deactivation Mechanisms of Uracil, Thymine, and 5-Fluorouracil: A Comparative ab Initio Study. J. Phys. Chem. A 2012, 116, 491–503. (48) Nakayama, A.; Harabuchi, Y.; Yamazaki, S.; Taketsugu, T. Photophysics of Cytosine Tautomers: New Insights into the Nonradiative Decay Mechanisms from MS-CASPT2 Potential Energy Calculations and Excited-State Molecular Dynamics Simulations. Phys. Chem. Chem. Phys. 2013, 15, 12322–12339. (49) Barbatti, M.; Borin, A. C.; Ullrich, S. Photoinduced Phenomena in Nucleic Acids I ; Topics in Current Chemistry; Springer Berlin Heidelberg, 2014; Vol. 355; pp 1–32. (50) Mai, S.; Richter, M.; Marquetand, P.; González, L. In Photoinduced Phenomena in Nu-

18

ACS Paragon Plus Environment

Page 18 of 20

Page 19 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

cleic Acids I ; Barbatti, M., Borin, A. C., Ullrich, S., Eds.; Topics in Current Chemistry; Springer Berlin Heidelberg, 2014; Vol. 355; pp 99–153. (51) Tao, H.; Levine, B. G.; Martínez, T. J. Ab Initio Multiple Spawning Dynamics Using Multi-State Second-Order Perturbation Theory. J. Phys. Chem. A 2009, 113, 13656– 13662. (52) Finley, J.; Malmqvist, P.-Å.; Roos, B. O.; Serrano-Andrés, L. The Multi-State CASPT2 Method. Chem. Phys. Lett. 1998, 288, 299. (53) Gozem, S.; Huntress, M.; Schapiro, I.; Lindh, R.; Granovsky, A. A.; Angeli, C.; Olivucci, M. Dynamic Electron Correlation Effects on the Ground State Potential Energy Surface of a Retinal Chromophore Model. J. Chem. Theory Comput. 2012, 8, 4069–4080. (54) Granovsky, A. A. Extended Multi-Configuration Quasi-Degenerate Perturbation Theory: The New Approach to Multi-State Multi-Reference Perturbation Theory. J. Chem. Phys. 2011, 134, 214113. (55) Granucci, G.; Persico, M.; Toniolo, A. Direct Semiclassical Simulation of Photochemical Processes with Semiempirical Wave Functions. J. Chem. Phys. 2001, 114, 10608–10615. (56) Plasser, F.; Granucci, G.; Pittner, J.; Barbatti, M.; Persico, M.; Lischka, H. Surface Hopping Dynamics using a Locally Diabatic Formalism: Charge Transfer in The Ethylene Dimer Cation and Excited State Dynamics in the 2-Pyridone Dimer. J. Chem. Phys. 2012, 137, 22A514. (57) Feinberg, M. In Chemical Reactor Theory: A Review ; Amundson, N., Lapidus, L., Eds.; Prentice Hall: Englewood Cliffs, New Jersey, 1977; pp 1–78. (58) Nangia, S.; Jasper, A. W.; Miller, T. F.; Truhlar, D. G. Army Ants Algorithm for Rare Event Sampling of Delocalized Nonadiabatic Transitions by Trajectory Surface 19

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Hopping and the Estimation of Sampling Errors by the Bootstrap Method. J. Chem. Phys. 2004, 120, 3586–3597. (59) Tao, H.; Allison, T. K.; Wright, T. W.; Stooke, A. M.; Khurmi, C.; van Tilborg, J.; Liu, Y.; Falcone, R. W.; Belkacem, A.; Martínez, T. J. Ultrafast internal conversion in ethylene. I. The excited state lifetime. J. Chem. Phys. 2011, 134, 244306. (60) Taherian, M.-R.; Maki, A. Optically Detected Magnetic Resonance Study of the Phosphorescent States of Thiouracils. Chem. Phys. 1981, 55, 85 – 96. (61) Hare, P. M.; Crespo-Hernández, C. E.; Kohler, B. Solvent-Dependent Photophysics of 1Cyclohexyluracil: Ultrafast Branching in the Initial Bright State Leads Nonradiatively to the Electronic Ground State and a Long-Lived 1 nπ ∗ State. J. Phys. Chem. B 2006, 110, 18641–18650. (62) Brister, M. M.; Crespo-Hernández, C. E. Direct Observation of Triplet-State Population Dynamics in the RNA Uracil Derivative 1-Cyclohexyluracil. J. Phys. Chem. Lett. 2015, 6, 4404–4409, PMID: 26538051. (63) Richter, M.; Mai, S.; Marquetand, P.; González, L. Ultrafast Intersystem Crossing Dynamics in Uracil Unravelled by Ab Initio Molecular Dynamics. Phys. Chem. Chem. Phys. 2014, 16, 24423.

20

ACS Paragon Plus Environment

Page 20 of 20