Top Banner
Proton Transfer in Nucleobases is Mediated by Water Kirill Khistyaev, Amir Golan, Ksenia B. Bravaya, Natalie Orms, Anna I. Krylov,* ,and Musahid Ahmed Department of Chemistry, University of Southern California, Los Angeles, California 90089-0482, United States Chemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States * S Supporting Information ABSTRACT: Water plays a central role in chemistry and biology by mediating the interactions between molecules, altering energy levels of solvated species, modifying potential energy proles along reaction coordinates, and facilitating ecient proton transport through ion channels and interfaces. This study investigates proton transfer in a model system comprising dry and microhydrated clusters of nucleobases. With mass spectrometry and tunable vacuum ultraviolet synchrotron radiation, we show that water shuts down ionization-induced proton transfer between nucleobases, which is very ecient in dry clusters. Instead, a new pathway opens up in which protonated nucleobases are generated by proton transfer from the ionized water molecule and elimination of a hydroxyl radical. Electronic structure calculations reveal that the shape of the potential energy prole along the proton transfer coordinate depends strongly on the character of the molecular orbital from which the electron is removed; i.e., the proton transfer from water to nucleobases is barrierless when an ionized state localized on water is accessed. The computed energetics of proton transfer is in excellent agreement with the experimental appearance energies. Possible adiabatic passage on the ground electronic state of the ionized system, though energetically accessible at lower energies, is not ecient. Thus, proton transfer is controlled electronically, by the character of the ionized state, rather than statistically, by simple energy considerations. 1. INTRODUCTION Excited-state proton transfer (PT, ESPT) is ubiquitous in chemistry 13 and biology, occurring, for example, in photo- active proteins such as green uorescent 4 and photoactive yellow proteins. 5 In DNA, ESPT between the nucleobases contributes to photoprotection. 6,7 The driving force for ESPT in DNA is the increased acidity of electronically excited nucleobases. Likewise, oxidized nucleobases, in which a valence electron is completely removed, also exhibit enhanced acidity leading to PT between the strands of DNA 7 and competes with electron hole (positive charge) migration along the strands. 8 Studies of isolated model systems, such as clusters of nucleobases, 9,10 reveal that ionization-induced PT is very facile, even in systems with no h-bonds such as the methylated π- stacked uracil dimer. 10 Electronic structure calculations show that ionization-induced PT between nucleobases is endother- mic in the neutral ground state (e.g., 0.8 eV uphill in AT), whereas it is exothermic in ionized species by 0.40.8 eV. Furthermore, ionization-induced PT in h-bonded pairs is barrierless, suggesting a high eciency for this process. Interestingly, even in π-stacked systems that have no h-bonds, PT is only slightly endothermic and involves a moderate barrier (0.2 eV in methylated uracil dimer); consequently, this channel opens up very close to the ionization threshold. A very recent study 11 of one-electron oxidation of DNA in solution has found that the initial steps involve proton transfer from a methyl group of thymine thus providing experimental evidence of the facile PT from non-hydrogen bonded moieties in realistic environments. Water is believed to be instrumental for PT in biological systems, 12 such as through water-lled ion channels, interfaces and membranes, and in aerosols. 13 The ability of water to form so-called water wiresfacilitating a relay-type transport of protons, the Grotthuss mechanism, is essential in all of these processes. 14,15 The proton-coupled electron transfer in DNA also involves water wires. 11 Notwithstanding the importance of water-mediated ground- and excited-state PT, the mechanistic details and dynamics of these processes are not well understood and are being investigated. For example, a recent study of small NO + (H 2 O) n clusters investigated how the shape of the h- bonded network controls proton-coupled water activation in HONO formation in the ionosphere. 16 Sequential PT through water bridges in acidbase reactions has been studied by time- resolved experiments in which the reaction has been initiated by an optical trigger exciting the photoacid. 17 Multiple-step ESPT mediated by solvent has been studied in hydroxyquino- line; 18,19 in particular, triple proton relay through alcohol chains has been characterized using time-resolved uorescence Received: June 18, 2013 Revised: June 26, 2013 Published: June 27, 2013 Article pubs.acs.org/JPCA © 2013 American Chemical Society 6789 dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 67896797
9

Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

Aug 22, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

Proton Transfer in Nucleobases is Mediated by WaterKirill Khistyaev,† Amir Golan,‡ Ksenia B. Bravaya,† Natalie Orms,† Anna I. Krylov,*,†

and Musahid Ahmed‡

†Department of Chemistry, University of Southern California, Los Angeles, California 90089-0482, United States‡Chemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States

*S Supporting Information

ABSTRACT: Water plays a central role in chemistry and biology bymediating the interactions between molecules, altering energy levels ofsolvated species, modifying potential energy profiles along reactioncoordinates, and facilitating efficient proton transport through ion channelsand interfaces. This study investigates proton transfer in a model systemcomprising dry and microhydrated clusters of nucleobases. With massspectrometry and tunable vacuum ultraviolet synchrotron radiation, we showthat water shuts down ionization-induced proton transfer betweennucleobases, which is very efficient in dry clusters. Instead, a new pathwayopens up in which protonated nucleobases are generated by proton transferfrom the ionized water molecule and elimination of a hydroxyl radical.Electronic structure calculations reveal that the shape of the potential energyprofile along the proton transfer coordinate depends strongly on thecharacter of the molecular orbital from which the electron is removed; i.e., theproton transfer from water to nucleobases is barrierless when an ionized state localized on water is accessed. The computedenergetics of proton transfer is in excellent agreement with the experimental appearance energies. Possible adiabatic passage onthe ground electronic state of the ionized system, though energetically accessible at lower energies, is not efficient. Thus, protontransfer is controlled electronically, by the character of the ionized state, rather than statistically, by simple energy considerations.

1. INTRODUCTION

Excited-state proton transfer (PT, ESPT) is ubiquitous inchemistry1−3 and biology, occurring, for example, in photo-active proteins such as green fluorescent4 and photoactiveyellow proteins.5 In DNA, ESPT between the nucleobasescontributes to photoprotection.6,7 The driving force for ESPTin DNA is the increased acidity of electronically excitednucleobases. Likewise, oxidized nucleobases, in which a valenceelectron is completely removed, also exhibit enhanced acidityleading to PT between the strands of DNA7 and competes withelectron hole (positive charge) migration along the strands.8

Studies of isolated model systems, such as clusters ofnucleobases,9,10 reveal that ionization-induced PT is very facile,even in systems with no h-bonds such as the methylated π-stacked uracil dimer.10 Electronic structure calculations showthat ionization-induced PT between nucleobases is endother-mic in the neutral ground state (e.g., 0.8 eV uphill in AT),whereas it is exothermic in ionized species by 0.4−0.8 eV.Furthermore, ionization-induced PT in h-bonded pairs isbarrierless, suggesting a high efficiency for this process.Interestingly, even in π-stacked systems that have no h-bonds,PT is only slightly endothermic and involves a moderate barrier(0.2 eV in methylated uracil dimer); consequently, this channelopens up very close to the ionization threshold. A very recentstudy11 of one-electron oxidation of DNA in solution has foundthat the initial steps involve proton transfer from a methyl

group of thymine thus providing experimental evidence of thefacile PT from non-hydrogen bonded moieties in realisticenvironments.Water is believed to be instrumental for PT in biological

systems,12 such as through water-filled ion channels, interfacesand membranes, and in aerosols.13 The ability of water to formso-called “water wires” facilitating a relay-type transport ofprotons, the Grotthuss mechanism, is essential in all of theseprocesses.14,15 The proton-coupled electron transfer in DNAalso involves water wires.11 Notwithstanding the importance ofwater-mediated ground- and excited-state PT, the mechanisticdetails and dynamics of these processes are not well understoodand are being investigated. For example, a recent study of smallNO+(H2O)n clusters investigated how the shape of the h-bonded network controls proton-coupled water activation inHONO formation in the ionosphere.16 Sequential PT throughwater bridges in acid−base reactions has been studied by time-resolved experiments in which the reaction has been initiatedby an optical trigger exciting the photoacid.17 Multiple-stepESPT mediated by solvent has been studied in hydroxyquino-line;18,19 in particular, triple proton relay through alcohol chainshas been characterized using time-resolved fluorescence

Received: June 18, 2013Revised: June 26, 2013Published: June 27, 2013

Article

pubs.acs.org/JPCA

© 2013 American Chemical Society 6789 dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−6797

Page 2: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

spectroscopy. Resonant ionization spectroscopy of gas-phase h-bonded clusters was employed to investigate proton versushydrogen-transfer pathways.20 Various experimental techniques,most notably ion-based infrared spectroscopy, have been usedto quantify important energetics and dynamics of ionization-induced PT, and the catalytic action of solvating waters ontautomerization equilibria via PT.21−24 However, a detailedunderstanding of the reaction pathways and directionality of PTin model systems, let alone in real biological and chemicalsystems, remains elusive.In this article, we report that water has a dramatic effect on

the PT in ionized species. We focus on methylated uracilclusters, capitalizing on our previous experience with this modelsystem;10,25 however, similar effects were also observed inmicrohydrated thymine species. We consider 1,3-dimethyluracil(mU) and its deuterated analog, 1,3-dimethyluracil-d6 (DmU).Mass spectrometry coupled with tunable vacuum ultraviolet(VUV) radiation molecular beam experiments show thatmicrohydration changes the branching ratio between differentrelaxation channels and entirely shuts down PT between thebases. Instead, a new pathway opens up, where protonatednucleobases are produced via PT from the ionized watermolecule and elimination of the hydroxyl radical. Electronicstructure calculations reveal that the shape of the potentialenergy profile along the PT coordinate depends strongly on thecharacter of the molecular orbital from which the electron isremoved; i.e., the PT from water to nucleobases becomesbarrierless upon access of an ionized state localized on water.The computed energetics of PT is in excellent agreement withthe experimental appearance energies. We also note thatpossible adiabatic processes, which become energeticallyaccessible at lower energies, are not efficient. Thus, PT iscontrolled electronically, by the character of the ionized state,rather than statistically, by simple energy considerations.The structure of the paper is as follows. The next section

describes experimental techniques and theoretical methods.Section III presents results and discussion. Our concludingremarks are given in section IV.

II. EXPERIMENTAL AND COMPUTATIONAL DETAILS

A. Experimental Methods. The experiments wereperformed on a molecular beam apparatus26,27 on the ChemicalDynamics Beamline at the Advanced Light Source (Figure S1,Supporting Information) using protocols developed in ourprevious studies.9,10,26,28−30 Monomers and dimers wereintroduced into a supersonic jet expansion by thermalvaporization. Argon gas was bubbled through water (D2O)and then passed over the sample vapors (mU) beforeexpanding to vacuum to produce a molecular beam at theinteraction region of a reflectron mass spectrometer where it isionized by the VUV light.A judicious combination of experimental source conditions

(backing pressure and reservoir heater temperature) allowed usto vary the population of dimethyluracil monomers, dimers,and their microhydrated clusters with up to seven watermolecules in the molecular beam (Figure 1). Higher temper-ature leads to increased yields of dimers, whereas an increase inbacking pressure facilitates hydration.Figure 2 shows the percentage of different dimer forms of

mU relative to all forms of mU present in the beam. Thepercentage of all (mU)2 is roughly constant beyond 250 Torr,with the value above 25%. However, the percentage of bare mUdimers significantly decreases, while the percentage of thehydrated mU dimers increases. Thus, increased backingpressure does not inhibit the formation of the mU dimer,rather it increases its hydration. Higher mU clusters (trimers,etc.) are not present.Figure 3 shows yield of different deuterated forms of mU

normalized by ion count of all mU forms. We observe that thesignals of deuterated forms of mU and (mU)2 increase withpressure.To determine the origin of transferred proton, we employed

various combinations of deuterated and nondeuterated species,such as DmU-H2O versus mU-D2O. Thus, we repeated theexperiments with H2O vapors and DmU. The latter isunavailable commercially; it was synthesized, as described indetail in our previous paper.10

Figure 1. Mass spectra of hydrated (with H2O) mU and its dimer using 12 eV photons with different backing pressure and nozzle temperature.

The Journal of Physical Chemistry A Article

dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−67976790

Page 3: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

B. Theoretical Methods and Computational Details.Electronic structure calculations were performed following thecomputational protocols developed and validated in ourprevious studies of ionized species.9,10,25,28−31 All calculationswere performed using Q-Chem32,33 and employed methodsranging from equation-of-motion coupled-cluster (EOM-CC)to density functional theory (DFT) with range-separatedfunctionals and dispersion correction (ωB97X-D).34,35

For accurate description of the ionized states, we employed avariant of EOM-CC with single and double substitutions forcalculation of ionization potentials (EOM-IP-CCSD) in whichproblematic target open-shell wave functions are derived byKoopmans-like operators acting on well-behaved closed-shellreference states.36,37 EOM-IP-CCSD simultaneously includesdynamical and nondynamical correlation, describes multipleelectronic states in one calculation, and treats states withdifferent numbers of electrons on the same footing. It is freefrom artificial symmetry breaking and spin-contamination.

All neutral ground-state structures and ionized protonatedstructures were optimized using DFT with the ωB97X-Dfunctional34,35 and the 6-311+G(d,p) basis set. Ionizationenergies were calculated using EOM-IP-CCSD with the 6-311+G(d,p) basis set. Ionized excited-state geometries wereoptimized at the EOM-IP-CCSD/6-31+G(d,p) level of theory.Mono- and dihydrated structures were obtained by placing

the water at positions that are most favorable for h-bonding andoptimizing these structures. Previous studies9,29,38 and chemicalintuition point out that in the most stable microhydrates, waterforms a hydrogen bond with either CO or NH group of anuclear base. mU has two CO groups and there are twopossible water positions for every group. This gives rise to fourmonohydrated structures shown in Figure S2 (SupportingInformation). Dihydrated structures (Figure S3, SupportingInformation) were obtained by adding a water molecule tomonohydrated structures, with a subsequent optimization.

III. RESULTS AND DISCUSSIONThe structures of the representative isomers of mU and itsdimer hydrated with one or two water molecules are shown inFigure 4 (and Figures S2 and S3, Supporting Information).

Because of methylation, only uracil’s oxygens, O(mU) areavailable for h-bonding. Consequently, in monohydratedstructures, water acts as a proton donor. The H(H2O)···O-(mU) bond lengths are 1.87 and 1.80 Å in mU-H2O and(mU)2-H2O, respectively. The second water molecule forms anh-bond with the first water molecule in mU-(H2O)2. Inhydrated (mU)2, the second water forms an h-bond with thesecond uracil ring.

Figure 2. Percentage of different dimer forms relative to the all formsof mU present in the beam. The percentage is calculated as the ratiobetween the signal of the hydrated mU dimer and signal of all forms ofmU present in the beam.

Figure 3. Ion counts of different deuterated forms of mU normalizedby ion count of all mU forms. The signals of deuterated forms of mUand (mU)2 are increasing with pressure.

Figure 4. Structures of 1,3-dimethyluracil (mU) and its dimerhydrated with one or two water molecules. In all structures, wateracts as a proton donor. Hydration of the dimer does not lead toconsiderable changes in the relative position of the two mU moieties;e.g., the distances between CO and C−CH3 groups in dry andhydrated (mU)2 clusters are around 3.3−3.5 Å. Temperature increaseresults in higher concentration of mU clusters, whereas backingpressure controls the degree of hydration.

The Journal of Physical Chemistry A Article

dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−67976791

Page 4: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

Previously,10 we demonstrated that PT between the bases inmU dimers occurs from a methyl group. Thus, the followingPT reactions are possible in ionized (mU)2(D2O) clusters:

→ +

→ + +

+ +

+

(mU) (D O) (mU) D OD

mUD mU OD2 2 2

(1)

→ +

→ + ‐ +

+ +

+

(mU) (D O) (mU) D O

mUH mU H D O2 2 2 2

2 (2)

Likewise, in (mU)n(D2O)m clusters, the appearance ofprotonated species is due to the PT between the bases,whereas the deuteron transfer will signify PT from the solventto uracil. By considering the ratios of the respective m/z peaks,one can quantify the efficiency of these competing PT channels.Figure 5A shows a VUV single photon ionization mass

spectrum of the molecular beam with ion signals corresponding

to the mU monomer (at m/z 140), dimer (at m/z 280), andtheir clusters with D2O. The inset in Figure 5A shows anenlarged portion of the spectrum around m/z 180 where themain feature corresponds to the mU(D2O)2 ion, a cluster ofone mU, and two deuterated waters. The two adjacent smallerpeaks (at m/z N + 1 and N + 2, marked by solid and dashedarrows, respectively, where N = 180) arise either due to thenatural isotope abundance (13C) or from protonated/deuterated species. As discussed below, similar spectra wereobtained for the DmU-H2O mixture. The isotopes account for

7.5% and 1% of the peaks at m/z 181 and m/z 182,respectively; similar values are obtained for the other hydratedspecies. Contrary to the PT yield, the natural isotopecontributions do not depend on the photon energies; thus,the energy dependence of the ratio between N + 1 and N + 2peaks to the parent peak [N = 180 for mU(D2O)2 or N = 182for DmU(H2O)2] allows us to distinguish between proton/deuteron transfer versus natural isotopes. As illustrated inFigure 5B, the (N + 1)/N ratio (solid line) is constant in themU-D2O beam, whereas the (N + 2)/N (dashed line) exhibits aclear onset (followed by a sharp rise) at about 10.8 eV. Thisdemonstrates that there is no PT between the bases; rather,there is a deuteron transfer from the solvating D2O to the mUdimer. Note that in dry mU clusters the N + 1 peak exhibitssuch a sharp rise at 8.9 eV.10

To confirm that the proton/deuteron transfer only occursfrom the solvent, we repeated the experiment with DmU andnondeuterated water (H2O). Figure 5C shows (N + 1)/N and(N + 2)/N ratios (solid and dashed lines, respectively) for N =182, which corresponds to the DmU(H2O)2 cation. Here weobserve that the N + 1 peak (proton transfer) exhibits athreshold behavior (at 10.8 eV, as in the mU-D2O experiments,Figure 5B), while (N + 2)/N remains constant. Thus, there isno deuteron transfer between the DmU species. Note that thekinetic isotope effect on the interbase PT was found to beminor for the stacked mU dimer,10 and therefore, the constantbehavior of the (N + 2)/N peak in the DmU-H2O is not due toH/D exchange in the base. Essentially water shuts down PTbetween the mU bases, which opens up at 8.9 eV in the absenceof water.10

In Figure 5A, there is evidence of mU dimers in themolecular beam. Furthermore, we can control the degree ofdimerization relative to solvation by varying the backingpressure of the carrier gas (Ar), as illustrated in Figure 1.Figure 6 (left panel) shows the effect of backing pressure on

the relative efficiency of PT in the mU-D2O beam. The increase

of backing pressure increases the yield of hydrated species, atthe expense of bare mU dimer and monomer. However, thetotal amount of all forms of the mU dimers (bare dimer plus allhydrated dimers) remains roughly the same (Figure 2). Theyield of interfragment PT is given by the signal of allprotonated species (dominated by mUH+, more than 85%).

Figure 5. Mass spectrum of hydrated mU and the dependence of theyield of protonated and deuterated species on photon energy. (A)Mass spectrum of hydrated (with D2O) mU and its dimer using 12 eVphotons. The inset shows the region at mass to charge (m/z) 180corresponding to [mU(D2O)2]

+. The arrows indicate two additionalpeaks at m/z 181 and 182 arising due to natural isotope abundance(13C) and due to protonated and deuterated species. The intensityratios between the peaks marked by the arrows at different photonenergies for mU(D2O)2 are shown in panel B. The constant behaviorof the m/z 181 (N + 1, N = 180) peak confirms that it arises fromisotopic contributions and is not due to PT. Panel C shows similarratios (for N + 1 and N + 2 m/z peaks) for N = 182 corresponding to[DmU(H2O)2]

+. In this case, the N + 2 peak is constant, revealing thatthere is no deuteron transfer between the bases.

Figure 6. Dependence of the yield of various protonated species onphoton energy and backing pressure. Left panel: effect of backingpressure (Ar gas) on PT. The black curve (mU-D2O PT) characterizesdeuteron transfer from D2O to uracil; the red curve [mU-D2O PT(normalized)] shows deuteron transfer from D2O to uracil divided bythe sum of mU and (mU)2 hydrates, ∑n,m[(mU)n(D2O)mD]

+/∑n,m≠0∑k+l=0,1 [(mU)n(D2O)mHkDl]

+. The blue curve (mU-mU PT)corresponds to PT between the mU molecules, ∑n[mU(D2O)nH]

+/∑m[(mU)2(D2O)m]

+. Right panel: appearance energies of deuteratedspecies [mU(D2O)nD]

+ for different cluster sizes n.

The Journal of Physical Chemistry A Article

dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−67976792

Page 5: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

When normalized to the total dimer population, the yield ofprotonated species decreases with backing pressure (Figure 6,left panel, mU-mU PT blue curve). This suggests that theinterfragment PT is suppressed by hydration of dimers ratherthan reducing the population of dimers via monomer hydration(hence reducing the number of molecules available forclustering). The yield of all deuterated forms of mU increasesupon hydration, as shown by the ratio of all deuterated forms toall forms of mU present in the beam (black line). Finally, uponnormalization to the population of the hydrated species, weobserve a constant ratio of all deuterated forms to all hydratedforms of mU [mU-D2O (normalized) red line around 0.25],hence confirming that the increased yield of PT is proportionalto the degree of hydration. This suggests that the rate of PT inthe hydrated clusters (and, possibly, its mechanism) does notdepend on degree of hydration.To understand the mechanism by which water shuts down

PT between the bases, we turn to electronic structurecalculations. Previous theoretical studies of microhydratednucleobases29 reported a small red shift (∼0.4 eV) in thelowest IE, in excellent agreement with experiments.22,29 Thecalculations revealed that the character of the ionized stateremains the same as in the isolated base (πCC orbital); the redshift was explained by the fact that in the lowest-energymicrohydrated structures, the nucleobase is acting as a protondonor. Using similar computational protocols (see section IIB),we conducted electronic structure calculations of micro-hydrated mU dimers.We observe that the hydration by one or two water

molecules does not change much the relative distance betweenthe two mU moieties (Figure 4); e.g., the distance betweenC(O) and C(CH3) moieties, which are involved ininterbase PT, in the mU dimer is 3.4 Å, whereas in(mU)2(H2O) and (mU)2(H2O)2 it varies between 3.3 and3.5 Å.The effect on the lowest ionized state is small, in terms of

both energy and the character of the state. We observe amoderate blue shift (∼0.1−0.3 eV) in the VIE, which isconsistent with the structures of hydrated species (Figure 4)where uracil acts as a proton acceptor. The character of thelowest ionized state is also unaffected, as evidenced by the wavefunction composition and the shapes of the respectivemolecular orbitals (MOs) shown in Figure 7.

Thus, neither structure nor energetics of the lowest IEexplains the observed behavior. However, we note that waterblocks the proton-accepting sites in mU; it may also addstructural rigidity to the system. Because in the systems with nohydrogen bonds PT requires significant rearrangements of thetwo fragments,10 additional clustering is expected to impedenecessary rearrangements of the two moieties. The analysis of

higher ionized states (Table 1) reveals that, although hydrationhas a relatively small effect on the lowest ionized states of mU,the ionized states localized on water are affected much strongerby the interaction with mU. Specifically, the state correspond-ing to ionization from a lone pair in water appears at 10.9−11.5eV in mU mono- and dihydrates, which is 1−1.5 eV lowercompared to the bare water molecule. The lower bound of theenergy range is remarkably close to the observed onset of PT inmicrohydrated clusters (10.8 eV). These results suggest that thePT channel opens up when the lowest ionized state on solvatedwater that corresponds to an excited ionized state of themU(H2O)n cluster becomes accessible. These results areconsistent with the experimentally observed onsets of PT,which are independent of the cluster size (Figure 6, rightpanel), in stark contrast to the lowest IE of microhydratednucleobases exhibiting notable dependence on the number ofhydrated waters (∼0.1 eV drop in IE per water mole-cule).22,26,29

To gain further insight into the electronic structure ofhydrated species and to validate theory, we focus on thephotoionization efficiency (PIE) curve (obtained by integratingthe area under the respective m/z peak) of the smallesthydrated cation, mU(D2O). The differentiation of the PIEcurve allows identification of multiple ionized states (the peakson the differentiated PIE curve correspond to the VIEs). ThePIE (black) and the differentiated (red) curves are shown inFigure 8, along with the computed VIEs. The curve features theionization onset at ∼8.6 eV and a series of peaks between 8.5and 11.5 eV. The computed AIE is in excellent agreement withthe experimental onset, whereas the computed VIEs match wellthe peaks of the differentiated curve. Thus, the peaks at 8.9, 9.9,10.0, 10.8, and 11.2 eV correspond to vertical ionizations fromthe 1A″, 2A″, 1A′, 2A′, and 3A″ states, respectively. Thecharacter of these states are illustrated by the respective MOs,which are also shown in Figure 8. As Figure 8 clearly illustrates,low-lying electronic ionized states correspond to the ionizationfrom mU, whereas the 3A″ state at 11.2 eV is localized on waterand is similar to the water-localized states observed inmicrohydrated dimers.To understand proton transfer in microhydrated mU, we

consider the lowest energy monohydrated mU (mUW1−1a) asa model structure. The first and fifth ionized states (whichcorrespond to the first ionizations of the mU and watermoieties, respectively) were optimized using EOM-IP-CCSD/6-31+G(d,p). The resulting geometries are presented in FigureS4, Supporting Information. In the first ionized state, in whichthe hole is localized on mU, the electrostatic interactions pushwater away from the CO group (the H−O distance increasesby 1.13 Å); this displacement acts against PT. In contrast, inthe fifth ionized state (the hole is localized on water),electrostatic forces pull water closer to mU, resulting in abarrierless PT and yielding optimized proton-transferredstructure of the cation.To further understand PT in solvated systems, we analyze

the potential energy profiles along the PT coordinate in[mU(H2O)]+. An approximate reaction coordinate wasgenerated by the interpolation between the initial structure ofthe neutral mU·H2O and that of the proton-transferred system,mUH+·OH.37,39

The profiles are shown in Figure 9. The energy of the ionizedstates that are localized on mU (1A″ and 2A″, Figure 8)increases along the PT coordinate (the PT is also endothermicin the neutral state). In contrast, the energy of the fifth state

Figure 7. MOs corresponding to the lowest ionized state in mU·H2O,(mU)2, and (mU)2·H2O.

The Journal of Physical Chemistry A Article

dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−67976793

Page 6: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

(3A″) corresponding to the ionization of water decreases,showing that PT from this state is a barrierless downhillprocess. A similar behavior is observed for other states; that is,energetically water-ionized states go down, whereas uracil-localized states go up. One can also consider a possibility ofadiabatic PT, e.g., on the lowest ionized state (1A″). This willof course involve changes in the electronic-state character fromthe mU-localized one to the water-ionized one and a barrier.The analysis of energy profiles shows that PT is energeticallyaccessible at ∼10.6 eV (upper bound); i.e., at this energy thesystem has enough energy to overcome the barrier on theadiabatic PES corresponding to the lowest ionized state of thesystem. Yet, the onset of PT yield occurs only at 10.8 eV, thussuggesting that such an adiabatic process is inefficient. This canbe readily rationalized by analyzing the respective electronicwave functions. The lowest ionized state corresponds to theionization of mU. In this state, the proton affinity of mU isreduced. Hence the short-time dynamics will involve structuralchanges that are not favorable for PT from water. Indeed, in theFranck−Condon optimized structures of the lowest electronic

state of [mU·H2O]+ (Figure S4, Supporting Information) the

distance between O(mU) and water hydrogen increases from1.87 to 2.90 Å. In contrast, PT is barrierless starting from theFranck−Condon point in the fifth ionized state. Thus, eventhough the system may have enough energy to overcome thebarrier on the lowest ionized-state adiabatic PES, this pathwayis not favored dynamically because the gradients in the Franck−Condon region point away from the PT coordinate. In contrast,when the right electronic state is accessed, the PT may occurballistically on the respective diabatic surface. We observe thatPT in hydrated mU species is controlled electronically, by thecharacter of the state, rather than statistically, by energyconsiderations alone.Figure 10 summarizes relevant energy differences in mUW1-

1a. VIEs for the first and fifth ionized states are 8.93 and 11.21eV, respectively. The energy difference between the final PT

Table 1. Vertical and Adiabatic Ionization Energies (eV) of mU, (mU)2, Water, and Various Hydrated Speciesa

state H2O mU (mU)2 mU·H2O mU·(H2O) (mU)2·H2O

mUW1-1a

mUW1-1b

mUW1-2a

mUW1-2b

mUW2-1a

mUW2-1b

mUW2-2b

mU2W1-1a

mUW1-1b

1 12.22 8.87 8.40 8.93 9.01 9.03 9.08 8.89 9.00 9.18 9.24 8.41 8.722 14.42 9.74 8.81 9.93 9.91 9.95 9.94 9.95 9.87 10.11 10.08 8.85 9.113 18.93 9.77 9.42 10.00 10.02 9.91 9.92 10.03 10.03 10.19 10.20 9.52 9.764 10.66 9.66 10.77 10.77 10.91 10.90 10.74 10.74 11.03 11.05 9.61 9.965 12.16 9.69 11.21 11.14 11.41 11.24 11.24 10.92 11.19 11.14 9.88 9.986 9.85 12.19 12.26 12.16 12.22 11.61 11.63 11.51 11.19 9.96 10.147 10.46 13.23 13.12 13.33 13.23 12.14 12.24 12.31 12.37 10.41 10.398 10.51 13.63 13.67 13.73 13.68 13.01 12.73 13.22 13.10 10.63 10.829 11.67 13.76 13.85 13.87 13.86 13.59 13.65 13.44 13.27 11.45 10.8410 11.88 14.20 14.14 14.25 14.27 13.66 13.65 13.91 13.93 11.73 11.9011 14.27 14.21 14.16 14.13 14.11 13.87 14.03 13.94 11.89 12.1612 14.59 14.6 14.52 14.54 14.17 14.05 14.31 14.25 13.22 12.621stAIE 8.59 8.59 8.56 8.61

aAll energies are computed by EOM-IP-CCSD/6-311+G(d,p) except for (mU)2 and (mU)2·H2O computed with the 6-31+G(d,p) basis set.

Figure 8. Photoionization efficiency curve (black) of [(mU+D2O)]+

and its derivative (red), observed using 8−12 eV photons. Thederivative plot reveals multiple ionized states derived by removing theelectron from different MOs. Black arrows point toward the calculatedionization energies.

Figure 9. Potential energy profiles for low-lying states of [mU·H2O]+

along the PT reaction coordinate. The proton is moved from the watermolecule to the mU oxygen site. The fifth ionized state, 3A″, in whichthe hole is on the water molecule (Figure 8), shows no barrierfacilitating downhill PT. PT from lower ionized states are possible;however, this involves changes in the electronic wave functioncharacter and requires more than 10.6 eV photon energy. The leftpanel shows the experimental ratio between the [mU(D2O)2]

+ (m/z180) and [mU(D2O)2D]

+ (m/z = 182) signals; it shows dramaticenhancement in PT when the 3A″ state is accessed.

The Journal of Physical Chemistry A Article

dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−67976794

Page 7: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

structure (fully relaxed mUH+·OH in the lowest ionized state)and initial ionized state is 2.18 eV (50.4 kcal/mol). This can bedescribed as energy gain due to PT on the diabatic surfacecorresponding to a water ionized state. The PT structure is 0.09eV (2.13 kcal/mol) above the lowest ionized state at theFranck−Condon geometry. Thus, for the lowest ionized state,the PT is an uphill process adiabatically. Finally, energyrequired for PT structure to dissociate producing mUH+ andOH is 0.6 eV (13.76 kcal/mol). Thus, adiabatically, the energydifference for the following process: [mU·H2O]

+ → mUH+ +OH is 0.69 eV or 15.9 kcal/mol (for the lowest ionized state).Combining this value with VIE we arrive at 9.62 eV; this wouldbe the mUH+ appearance energy on the lowest ionized stateassuming no barrier. However, as illustrated by Figure 9, weanticipate a barrier on the lowest adiabatic PES. Theexperimental onset for mUH+ is much higher (∼11 eV) andagrees well with VIE corresponding to the fifth ionized state ofmU·H2O.The appearance energies of mU(H2O)nH

+ does not dependon the cluster size, as illustrated in Figure 6 (right panel), andto understand this, we consider a dihydrated system, mU-(H2O)2 (Figure 4). In this structure, the second (outer) watermolecule acts as a proton donor forming an h-bond with thefirst (inner) water h-bonded to the carbonyl group. One canconsider several possibilities for ionization-induced PT leadingto (A) a structure with H3O

+ resulting from a single PT fromthe outer molecule; (B) a structure with protonated mU boundto the OH radical and solvated by the outer water moleculederived by single PT from the inner water molecule to mU, and(C) a structure with protonated mU solvated by water and anouter OH radical. We prepared such structures by manuallydisplacing the parent structure and optimized them by usingprotocols described above. The initial structures are shown inthe left panel of Figure 11A,C,E, whereas the resultingoptimized structures are shown on the right (Figure 11B,D,F).The lowest ionization energy of water in this dihydrate

correspond to the state with a hole localized on the outer water.Thus, we expect this state to relax via PT to either the first orthe third structures. However, the optimization of the structureshown in Figure 11A converged to the double PT structure(Figure 11B) indicating that there is no local minimumcorresponding to the first structure (with H3O

+). Thus,

ionization of water in a dihydrate leads to the double PTstructure via a Grotthuss-like pathway. The energetics of thisprocess is in agreement with a higher proton affinity of mUcompared with water. However, it is interesting that there isapparently no barrier along the double PT reaction coordinate(the optimization of the state with the hole localized on theouter water converges to a doubly proton-transferred state).We also note that there is no local minimum corresponding tothe structure with H3O

+, which can be considered as anintermediate along the double stepwise PT pathway from theouter water to uracil. Energy differences between the structurescorresponding to the single and double PT (Figure 11B,D,respectively) are 23 kcal/mol (computed by ωB97X-D). This isin agreement with the fact that in clusters, the more stablestructures are the ones with the OH radical on the outside,which can be rationalized by counting the number of h-bondsin the two structures. The energy difference between the fifthionized state at the initial Franck−Condon geometry and thelowest ionized state at the double PT structure is 53.4 kcal/mol.In large water clusters, the lowest ionized states correspond

to the surface states, where there are waters that serve only asproton donors.40 Thus, the IEs corresponding to the surfacestates should be relatively independent of the cluster size (andeven the chemical nature of its core). The experimental onsetsfor protonated mU(H2O)n clusters are remarkably insensitiveto the cluster size (Figure 6, right panel); this suggests that inlarger clusters the surface-ionized states lead to multiplebarrierless PT, yielding solvated protonated uracil with theOH radical on the surface. Thus, such clusters of water withmolecules with relatively high proton affinity could serve as

Figure 10. Energy diagram describing relevant ionized states and theirordering at different geometries along the PT coordinate. All energiesare given in electron volts and are calculated with EOM-IP-CCSD/6-311+G**.

Figure 11. Possible proton-transferred structures in [mU(H2O)2]+.

Left panels show manually distorted structures used as starting pointsfor optimization. Right panels show the final optimized structures ofthe ionized species. Distances are in Ångstroms.

The Journal of Physical Chemistry A Article

dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−67976795

Page 8: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

model systems for studying directionality in Grotthuss-like PTthrough water wires and membrane interfaces.41,42

Although most of the experiments in this work focused onmU, this nucleobase is by no means unique in that water has asignificant effect on PT. Similar experiments performed onthymine show that in the absence of water, PT begins at 9.20eV, with a major rise in signal between 9.7 and 9.9 eV.9

Thymine provides an interesting comparison, because both h-bonded and π-stacked dimers populate the molecular beam,9 incontrast to mU, which forms only π-stacked dimers.10 Thecalculations suggested that it is h-bonded thymine dimers thatgive rise to this signal at 9.7 eV, whereas the lower onset wasexplained by a dimer with π-stacked geometry.9 Uponsolvation, PT switches off at these lower energies, as isevidenced in the signal for TH+ and T(H2O)H

+ shown inFigure S5 (Supporting Information). The onset for PT isaround 10.6 eV, with a major rise at 11.2 eV, which is verysimilar to the onsets observed in mU. The shapes of the curvesfor protonated thymine species are also very similar to those inFigure 5B,C. This suggests that a similar PT mechanism fromthe solvent is occurring.

IV. CONCLUSIONSWe conclude that in both h-bonded and π-stacked nucleobasedimers and larger clusters, solvation shuts down PT betweenthe bases, which is rather efficient in “dry” clusters. It is onlywhen the solvent is ionized that PT begins again. Our findingsillustrate that water has a dramatic effect on PT pathways, notonly by serving as a wire for proton transport but also byshutting down other PT routes. In our model systems, anoutermost ionized water molecule acts as an acid (activated byan ionization event), and the nucleobase acts as a base, whereasother waters may participate in PT either as spectators or asintermediate proton acceptors, as shown recently in photo-induced acid−base reactions.17 We attribute the remarkablesimilarity between the appearance onsets in solvated mU andthymine, as well as insensitivity of the onsets and shapes of theappearance curves on the cluster sizes, to the fact that thelowest ionized states in which the hole is localized on thesolvent correspond to the surface states, i.e., water moleculesacting as proton donors only. Electronic structure calculationsshow that these IEs are rather insensitive to the size and/orchemical identity of the cluster core, mU versus mU dimerversus thymine. Thus, these states become accessible at verysimilar energies, initiating facile PT to the accepting base, eitherby direct means (in monohydrates) or through the mediatingwater molecules.A growing body of studies illustrating the central role of PT

has led to a paradigm shift in the discussion of water and itsactive role in biology and chemistry. Water is no longer seen asjust a solvent but is an active participant in a variety ofprocesses such as enzyme catalysis and membrane transport.Water has also been shown to catalyze reactions43 that areimportant in biology and atmospheric chemistry.16,44 Proton-coupled electron transfer in DNA is mediated by waterchains.11 Autoionization in water also drives a variety ofprocesses that are critical to life and biology,12 whereas PTthrough nanopores, artificial membranes, and structures hasmajor ramifications in energy conversion and storagetechnologies. PT in nanoconfined geometries45 has implica-tions for catalysis and solar energy conversion, whereas ionshave been shown to enhance the transfer of protons throughaqueous interfaces.46 In this work we have shown that PT can

be very effectively controlled by subtly changing how DNAbases hydrogen bond and stack within themselves and uponsolvation and thus can provide a template for novel dynamicalstudies in the temporal, spatial, and spectroscopic domain.

■ ASSOCIATED CONTENT*S Supporting InformationDetails of the experimental setup, relevant structuralinformation on methyluracil clusters, discussion of PTefficiency in thymine clusters. This information is availablefree of charge via the Internet at http://pubs.acs.org.

■ AUTHOR INFORMATIONNotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSA.G., M.A., and the ALS are supported by the Office of Science,Office of Basic Energy Sciences, of the U.S. Department ofEnergy under Contract No. DE-AC02-05CH11231, throughthe Chemical Sciences Division. A.I.K. acknowledges supportfrom the Department of Energy through the DE-FG02-05ER15685 grant. The authors acknowledge the contributionsof Oleg Kostko in providing the thymine/water data and QiaoRuan for mass spectra analysis.

■ REFERENCES(1) Rosspeintner, A.; Lang, B.; Vauthey, E. Ultrafast Photochemistryin Liquids. Annu. Rev. Phys. Chem. 2013, 64, 247.(2) Tolbert, L. M.; Solntsev, K. M. Excited-State Proton Transfer:From Constrained Systems to “Super” Photoacids to Superfast ProtonTransfer. Acc. Chem. Res. 2002, 35, 19−27.(3) Formosinho, S. J.; Arnaut, L. G. Excited-State Proton TransferReactions II. Intramolecular Reactions. J. Photochem. Photobiol. 1992,75, 21−48.(4) van Thor, J. J. Photoreactions and Dynamics of the GreenFluorecent Protein. Chem. Soc. Rev. 2009, 38, 2935−2950.(5) Carroll, E. C.; Song, S.-H.; Kumauchi, M.; vanStokkum, I. H. M.;Jailaubekov, A.; Hoff, W. D.; Larsen, D. S. Subpicosecond Excited-State Proton Transfer Preceding Isomerization During the Photo-recovery of Photoactive Yellow Protein. J. Phys. Chem. Lett. 2010, 1,2793−2799.(6) Schultz, T.; Samoylova, E.; Radloff, W.; Hertel, I. V.; Sobolewski,A. L.; Domcke, W. Efficient Deactivation of a Model Base Pair viaExcited-State Hydrogen Transfer. Science 2004, 306, 1765−1768.(7) Ghosh, A. K.; Schuster, G. B. Role of the Guanine N1 IminoProton in the Migration and Reaction of Radical Cations in DNAOligomers. J. Am. Chem. Soc. 2006, 128, 4172−4173.(8) Charkaborty, T., Ed. Charge Migration in DNA; Springer: BerlinHeidelberg, Germany, 2007.(9) Bravaya, K. B.; Kostko, O.; Ahmed, M.; Krylov, A. I. The Effect ofπ-Stacking, H-Bonding, and Electrostatic Interactions on theIonization Energies of Nucleic Acid Bases: Adenine-Adenine,Thymine-Thymine and Adenine-Thymine Dimers. Phys. Chem.Chem. Phys. 2010, 12, 2292−2307.(10) Golan, A.; Bravaya, K. B.; Kudirka, R.; Leone, S. R.; Krylov, A. I.;Ahmed, M. Ionization of Dimethyluracil Dimers Leads to FacileProton Transfer in the Absence of H-bonds. Nat, Chem. 2012, 4, 323−329.(11) Barnet, R. N.; Joseph, J.; Landman, U.; Schuster, G. B. OxidativeThymine Mutation in DNA: Water-Wire-Mediated Proton-CoupledElectron Transfer. J. Am. Chem. Soc. 2013, 135, 3904−3914.(12) Ball, P. The Importance of Water. In Astrochemistry andAstrobiology: Physical Chemistry in Action; Smith, I.W. M., Cockell, C.S., Leach, S., Eds.; Springer: Berlin, Heidelberg, Germany, 2013; pp169−210.

The Journal of Physical Chemistry A Article

dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−67976796

Page 9: Proton Transfer in Nucleobases is Mediated by Wateriopenshell.usc.edu/pubs/pdf/jpca-117-6789.pdf · A 2013, 117, 6789−6797. spectroscopy. Resonant ionization spectroscopy of gas-phase

(13) Voth, G. A. Computer Simulation of Proton Solvation andTransport in Aqueous and Biomolecular Systems. Acc. Chem. Res.2006, 39, 143−150.(14) Agmon, N. The Grotthuss Mechanism. Chem. Phys. Lett. 1995,244, 456−462.(15) Chandler, D.; Dellago, C.; Geissler, P. Wired-up Water. Nat.Chem. 2012, 4, 245−247.(16) Relph, R. A.; Guasco, T. L.; Elliott, B. M.; Kamrath, M. Z.;McCoy, A. B.; Steele, R. P.; Schofield, D. P.; Jordan, K. D.; Viggiano,A. A.; Ferguson, E. E.; Johnson, M. A. How the Shape of an H-bondedNetwork Controls Proton-Coupled Water Activation in HONOFormation. Science 2010, 327, 308−312.(17) Mohammed, O. F.; Pines, D.; Dreyer, J.; Pines, E.; Nibbering, E.T. J. Sequential Proton Transfer Through Water Bridges in Acid-BaseReactions. Science 2005, 310, 83−86.(18) Kwon, O.-H.; Lee, Y.-S.; Yoo, B. K.; Jang, D.-J. Excited-StateTriple Proton Transfer of 7-hydroxyquinoline along a Hydrogen-Bonded Alcohol Chain: Vibrationally Assisted Proton Tunneling.Angew. Chem., Int. Ed. Engl. 2006, 45, 415−419.(19) Lim, H.; Jeong, H.; Park, S.-Y.; Lee, J. Y.; Jang, D.-J. Excited-State Proton-Relay Dynamics of 7-Hydroxyquinoline Controlled bySolvent Reorganization in Room Temperature Ionic Liquids. Phys.Chem. Chem. Phys. 2012, 14, 218−224.(20) Manca, C.; Tanner, C.; Leutwyler, S. Excited State HydrogenAtom Transfer in Ammonia-Wire and Water-Wire Clusters. Int. Rev.Phys. Chem. 2005, 24, 457−488.(21) Sukhodub, L. F. Interactions and Hydration of Nucleic AcidBases in a Vacuum. Experimental Study. Chem. Rev. 1987, 87, 589−606.(22) Kim, S. K.; Lee, W.; Herschbach, D. R. Cluster BeamChemistry: Hydration of Nucleic Acid Bases; Ionization Potentialsof Hydrated Adenine and Thymine. J. Phys. Chem. 1996, 100, 7933−7937.(23) Kim, N. J.; Kim, H. M.; Kim, S. K. Ionization-Induced ProtonTransfer in Thymine-Ammonia van der Waals Clusters. Int. J. Mass.Spectrom. 2007, 261, 32.(24) Mizuse, K.; Kuo, J.-L.; Fujii, A. Structural Trends of IonizedWater Networks: Infrared Spectroscopy of Water Cluster RadicalCations (H2O)n

+ (n=3−11). Chem. Sci. 2011, 2, 868−876.(25) Zadorozhnaya, A. A.; Krylov, A. I. Zooming into Pi-StackedManifolds of Nucleobases: Ionized States of Dimethylated UracilDimers. J. Phys. Chem. A 2010, 114, 2001−2009.(26) Belau, L.; Wilson, K. R.; Leone, S. R.; Ahmed, M. Vacuum-Ultravioled Photoionization Studies of the Microhydration of DNABases (Guanine, Cytosine, Adenine, and Thymine). J. Phys. Chem. A2007, 111, 7562−7568.(27) Golan, A.; Ahmed, M. Molecular Beam Mass Spectrometry withTunable Vacuum Ultraviolet (VUV) Synchrotron Radiation. J. Vis.Exp. 2012, 68, e50164.(28) Kostko, O.; Bravaya, K. B.; Krylov, A. I.; Ahmed, M. Ionizationof Cytosine Monomer and Dimer Studied by VUV Photoionizationand Electronic Structure Calculations. Phys. Chem. Chem. Phys. 2010,12, 2860−2872.(29) Khistyaev, K.; Bravaya, K. B.; Kamarchik, E.; Kostko, O.;Ahmed, M.; Krylov, A. I. The Effect of Microhydration on IonizationEnergies of Thymine. Faraday Discuss. 2011, 150, 313−330.(30) Bravaya, K. B.; Kostko, O.; Dolgikh, S.; Landau, A.; Ahmed, M.;Krylov, A. I. Electronic Structure and Spectroscopy of Nucleic AcidBases: Ionization Energies, Ionization-Induced Structural Changes,and Photoelectron Spectra. J. Phys. Chem. A 2010, 114, 12305−12317.(31) Golubeva, A. A.; Krylov, A. I. The Effect of π-Stacking and H-Bonding on Ionization Energies of a Nucleobase: Uracil Dimer Cation.Phys. Chem. Chem. Phys. 2009, 11, 1303−1311.(32) Krylov, A. I.; Gill, P. M. W. Q-Chem: An Engine for Innovation.WIREs Comput. Mol. Sci. 2013, 3, 317−326.(33) Shao, Y.; Fusti-Molnar, L.; Jung, Y.; Kussmann, J.; Ochsenfeld,C.; Brown, S.; Gilbert, A. T. B.; Slipchenko, L. V.; Levchenko, S. V.;O’Neill, D. P.; et al. Advances in Methods and Algorithms in a Modern

Quantum Chemistry Program Package. Phys. Chem. Chem. Phys. 2006,8, 3172−3191.(34) Chai, J.-D.; Head-Gordon, M. Systematic Optimization of Long-Range Corrected Hybrid Density Functionals. J. Chem. Phys. 2008,128, 084106.(35) Chai, J.-D.; Head-Gordon, M. Long-Range Corrected HybridDensity Functionals with Damped Atom-Atom Dispersion Inter-actions. Phys. Chem. Chem. Phys. 2008, 10, 6615−6620.(36) Krylov, A. I. Equation-of-Motion Coupled-Cluster Methods forOpen-Shell and Electronically Excited Species: The Hitchhiker’s Guideto Fock Space. Annu. Rev. Phys. Chem. 2008, 59, 433−462.(37) Pieniazek, P. A.; Arnstein, S. A.; Bradforth, S. E.; Krylov, A. I.;Sherrill, C. D. Benchmark Full Configuration Interaction and EOM-IP-CCSD Results for Prototypical Charge Transfer Systems: Non-covalent Ionized Dimers. J. Chem. Phys. 2007, 127, 164110.(38) Ghosh, D.; Isayev, O.; Slipchenko, L. V.; Krylov, A. I. The Effectof Solvation on Vertical Ionization Energy of Thymine: FromMicrohydration to Bulk. J. Phys. Chem. A 2011, 115, 6028−6038.(39) We generated ten structures using the following expression forthe Cartesian coordinates: xn = xinitial + (xfinal − xinitial)·n/10, where xndenotes the respective Cartesian coordinates. The initial and finalstructures are optimized structures of the neutral mU·H2O andmUH+·OH; n can be interpreted as a step along the PT coordinate.(40) Pieniazek, P. A.; Sundstrom, E. J.; Bradforth, S. E.; Krylov, A. I.The Degree of Initial Hole Localization/Delocalization in IonizedWater Clusters. J. Phys. Chem. A 2009, 113, 4423−4429.(41) Springer, A.; Hagen, V.; Cherepanov, D. A.; Antonenko, Y. N.;Pohl, P. Protons Migrate along Interfacial Water without SignificantContributions from Jumps Between Ionizable Groups on theMembrane Surface. Proc. Natl. Acad. Sci. 2011, 108, 14461−14466.(42) Kwon, O.-H.; Mohammed, O. F. Water-Wire Catalysis inPhotoinduced Acid-Base Reactions. Phys. Chem. Chem. Phys. 2012, 14,8974−8980.(43) Matsuda, Y.; Yamada, A.; i.Hanaue, K.; Mikami, N.; Fujii, A.Catalytic Action of a Single Water Molecule in a Proton-MigrationReaction. Angew. Chem., Int. Ed. Engl. 2010, 49, 4898−4901.(44) Buszek, R. J.; Barker, J. R.; Francisco, J. S. Water Effect on theOH + HCl Reaction. J. Phys. Chem. A 2012, 116, 4712−4719.(45) Kofinger, J.; Hummer, G.; Dellago, C. Single-File Water inNanopores. Phys. Chem. Chem. Phys. 2011, 13, 15403−15417.(46) Mishra, H.; Enami, S.; Nielsen, R. J.; Hoffmann, M. R.;Goddard, W. A.; Colussi, A. J. Anions Dramatically Enhance ProtonTransfer through Aqueous Interfaces. Proc. Natl. Acad. Sci. 2012, 109,10228−10232.

The Journal of Physical Chemistry A Article

dx.doi.org/10.1021/jp406029p | J. Phys. Chem. A 2013, 117, 6789−67976797