O Overtone Bending Vibration in the Water Dimer - American

May 30, 2018 - other molecules1−11 is important in studying solvent−solute .... taken between 0 and π, whereas those in eqs 3.5 and 3.6 are ...
1 downloads 0 Views 3MB Size
Subscriber access provided by University of Massachusetts Amherst Libraries

A: Kinetics, Dynamics, Photochemistry, and Excited States 2

Relaxation of the HO Overtone Bending Vibration in the Water Dimer Hydroxyl Radical Complex ...

Hyung Kyu Shin J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b03674 • Publication Date (Web): 30 May 2018 Downloaded from http://pubs.acs.org on May 30, 2018

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 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 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.

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 28 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

Relaxation of the H2O Overtone Bending Vibration in the Water Dimer...Hydroxyl Radical Complex H. K. Shin* Department of Chemistry, University of Nevada, Reno, Nevada 89557, United States

ABSTRACT: The relaxation mechanism of the overtone bending vibration in the collision of the water dimer with the vibrationally excited hydroxyl radical is studied by use of trajectory procedures. The transfer of the OH(v=1) energy to the dimer stretches is followed by a near-resonant first overtone transition to the donor monomer. Nearly a quarter of the trajectories undergo a complex-mode collision forming the (H2O)2...OH complex bound by a hydrogen bond with the lifetime ranging from a subpicosecond scale to >100 ps. The overtone vibration relaxes to the ground state, transferring approximately half of its energy to the dimer hydrogen-bonding (H2O...H2O) and the remaining half to the complex hydrogen-bonding (H2O)2...OH, via near-resonant pathways, each consisting of a series of intermolecular low-frequency vibrations.

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

1. INTRODUCTION Understanding the nature of a water dimer interacting with other molecules1-11 is important in studying solvent-solute interactions, which are intrinsic occurrences in all chemical reactions taking place in an aqueous environment. The important processes in these interactions are intermolecular energy transfer between the collision partners and intramolecular energy redistribution in each molecule. Such interactions involving the collision partner ranging from a diatomic molecule7,9 or another water molecule(s)11-26 to a large organic molecule3 accompany the flow of energy among the interacting molecules, which leads to opening reaction pathways or stabilizing reactive intermediates or products. A significant process occurring in the water + molecule interactions is inter- and intramolecular energy transfer through the hydrogen bond.11,16,17 A particularly interesting feature of the interaction system is the formation of a new hydrogen bonding, when the dimer interacts with another water molecule(s) or the molecule consists of an electronegative atom. Recent study11 of the direct-mode interaction of the dimer with the vibrationally excited water molecule shows the decay of the energy collisionally transferred to the OH stretch generates a near-resonant first overtone transition of the donor bend mode, which then efficiently redistributes among various lowfrequency intermolecular vibrations associated with the motion of the H-bonded OH via a series of small energy steps. A noteworthy result in the study is that relaxation of the first overtone bend proceeds through a series of low-frequency intermolecular vibrations in the dimer, but leaves approximately one quantum of energy in the bend. Relaxation of the bending mode to near v = 1 state rather than to the ground state differs from that found from experimental and theoretical studies on the relaxation of liquid water, where the overtone is known to relax to the ground state.16,17 The two-body collision of (H2O)2 + H2O proceeds primarily through a direct-mode mechanism on a subpicosecond time scale and the relaxation pathways consisting of low-frequency intermolecular 2 ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28 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

vibrations are shown to be efficient for transporting one quantum of the bending vibration, but not two.11 However, even in a simple dimer + molecule interaction, relaxation of the overtone bend to the ground state can occur, if additional intermolecular energy decay pathways are present. Such intermolecular pathways can exist, if the molecule exerts attractive forces strong enough to form a collision complex. One such system is (H2O)2 + OH, where the energies of electrostatic and induction are strong, so that the collision-induced complex can form. The complex (H2O)2...OH has been observed in solid Ne by Tsuji and Shibuya27 and in liquid He by Hernandez et al.28 in their IR spectroscopic studies. Several research groups have reported formation of the complex in their theoretical studies.2,4,5,10 Hernandez et al. also reports theoretical studies.28 In some of these studies, the newly formed complex bond is shown to be stronger than the hydrogen-bonding in the water dimer.5,10,27 In this paper we study inter- and intramolecular energy redistribution in the complex (H2O)2...OH in the collision between the water dimer and the hydroxyl radical, where the vibrationally excited OH(v=1) transfers its energy to the stretches of the dimer, followed by a near-resonant first overtone transition from the stretches to the bending mode. We follow transfer of the overtone bending energy to various low-frequency intermolecular vibrations in the complex, which consists of two hydrogen bonds, one in the dimer and another between the dimer and the radical. We use the approach developed in ref. 11 to explore the pathways of relaxation leading the overtone bend to the ground state in the presence of two hydrogen bonds and the effectiveness of each pathway. 2. INTERACTION MODEL AND NUMERICAL METHODS The cylindrical coordinate system (ρ, Z, Φ) for the (H2O) 2 + OH collision is shown in Fig. 1, which is related to the Cartesian coordinates as (ρ, Z, Φ) = [(X2 + Y2)1/2, Z, tan-1(Y/X)], chosen such that 3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 4 of 28

the direction of the initial relative velocity vector coincides with the Z-axis.11 The distance ρ is the impact parameter b and R is the (H2O)2-to-OH relative separation. We denote the atoms of the donor and acceptor monomers in the dimer by (Hd2OdHd1) and (Ha2OaHa1), respectively, where the subscripts “d” and “a” refer to the donor and acceptor, respectively. These notations distinguish the dimer atoms from those of OH, written without a subscript. At a large distance the molecules approach one another with interaction energy primarily determined by the electrostatic and induction effects: Uµ-µ = - (µH2OµOH)[2cosθ1cosθ2 - sinθ1sinθ2cos(φ1 - φ2) ]/R3 ,

(1.1)

Uq-indµ = - (ce)2αH2O/2R4,

(1.2)

Uµ-indµ = - ( µH2O αOH)(3cosθ2 + 1)/2R6- ( µOH αH2O)(3cosθ2 + 1)/2R6, 2

2

(1.3)

where Uq-indµ is the interaction of dipole with the partial charge on the radical determined by the ratio c = µOH/ere,OH. Here re,OH is the equilibrium OH bond distance. The long-range interaction VLR is the sum of these and the Lennard-Jones energy UL-J = 4ε[(σ/R)12 - (σ/R)6]. We use the values of the dipole moments, polarizabilities and LJ constants for the dimer given in ref. 11. For OH, they are

µ = 1.660 Debye,29 α = 1.165 Å3,30 σ = 3.111 Å and ε/k = 281.3 K.31 When we average Uµ-µ and Uµindµ

over the orientation angles and add it to ULJ, the resulting energy can be fit closely to the Morse

form V(R) = D{exp[(Re - R)/a] – 2 exp[(Re – R)/2a]}, D = 3.64 kcal mol-1, Re = 3.13 Å and a = 0.232 Å. We use this energy to initialize the (H2O)2 + OH interaction in the trajectory calculation. When the molecules approach in close range of one another, strong short-range forces take over the (H2O)2 + OH motion and guide it through the repulsive interaction region, where energy transfer takes place. We use the pairwise interaction between the atoms of the dimer and OH at a 4 ACS Paragon Plus Environment

Page 5 of 28 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

configuration of inter- and intramolecular coordinates. For ij intermolecular atom-atom pairs, the short-range interactions then can be expressed as Vinter = ∑ Dij {exp[(re,ij - rij)/aij] - 2exp[(re,ij - rij)/2aij]},

(2)

ij

where the subscript “e” refers to the equilibrium value such that rij = re,ij + xij, defining the instantaneous displacement xij. For the O-O and H-H interactions, we use the B3LYP/6-311+G(d,p)level values DO-O = 5.188 and DH-H = 4.769 eV, along with the frequencies 1633 and 4420 cm-1, respectively. We derive the complete form of Vinter using the ij distances rOdH/OaH = [R2+ (½rOO)2 + (γOrOH)2 ± R(rOOcosθ2 m 2γOrOHcosθ1) m rOOγOrOHF] ½,

(3.1)

rOOd/OOa = [R2+ (½rOO)2 + (γHrOH)2 ± R(rOOcosθ2 ± 2γHrOHcosθ1) ± rOOγHrOHF] ½,

(3.2)

rHHd1/HHd2 = [rOdH2 + rOdHd1/OdHd22 - 2rOdHrOdHd1/OdHd2cosψHOdHd1/HOdHd2]½ ,

(3.3)

rOHd1/OHd2 = [rOOd2 + rOdHd1/OdHd22 - 2rOOdrOdHd1/OdHd2cosψOOdHd1/OOdHd2]½ ,

(3.4)

rOHa1/OHa2 = [rOOa2 + rOaHa1/OaHa22 - 2rOOarOaHa1/OaHa2cosψOOaHa1/OOaHa2]½ ,

(3.5)

rHHa1/HHa2 = [rOaH2 + rOaHa1/OaHa22 - 2rOaHrOaHa1/OaHa2cosψHOaHa1/HOaHa2]½ ,

(3.6)

where γO = [mO/(mH + mO)], γH = [mH/(mH + mO)] and F = cosθ1cosθ2 + sinθ1sinθ2cos(φ1 - φ2). The subscript of ψ signifies the associated angle; e.g., ψHOdHd1 is for ∠HOdHd1. The angles in Eqs. (3.3) and (3.4) are replaced by their averages taken between 0 and π, whereas those in Eqs. (3.5) and (3.6) are replaced by the averages between 0 and π - 60°, where 60° is the angle between the Od-Oa direction and the three-atom acceptor plane. The intramolecular potentials for the ij bond are Vintra =

∑D

ij

{1 – exp[(re,ij – rij)/2aij]}2; ij = OH, OdHd1, OdHd2, OaHa1, OaHa2,

ij

5 ACS Paragon Plus Environment

(4)

The Journal of Physical Chemistry 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

where aij = (Dij/2µOH)1/2/ωij, ωij = 2πν ij and µOH = mOmH/mOH. The exponential parameters aij in Eq. (2) are defined similarly. The sum of these two potentials will be referred to as the two-body interaction V(2) = Vinter + Vintra. For the radical the molecular constants are DOH = 4.634 eV and re,OH = 0.9758 Å, the B3LYP-level values as in the atom-atom interactions shown above along with νOH = 3568 cm-1 from the data listed in ref. 32; 3737.76 – 2(84.881) = 3568 cm-1.33 For the water dimer, the bending frequencies of the water dimer donor unit are ν2 = 1616 and 2ν2 = 3194 cm-1 from the infrared spectrum of the water trapped in neon at 3 K.33 The observed frequencies of the bondedOdHd2 and non-bonded OdHd1 of water dimer are 3601 and 3735 cm-1, respectively.34 We use these observed values in sampling the initial conditions of bonded and non-bonded donor vibrations. We note that the latter two stretching frequencies are close to the normal mode values ν1 = 3638 and ν3 = 3733 cm-1, respectively, of the Rayleigh-Schrӧdinger vibrational second-order perturbation theory (VPT2) reported by Kjaergaard et al.35 The ν3 value is nearly identical to the observed non-bonded donor frequency.34 The VPT2 method is known to work well for bending motions. The bending frequencies obtained by VPT2 calculations are ν2 = 1596 and 2ν2 = 3159 cm-l, which are only slightly lower (∼1%) than the experimental data shown above.33 For energy transfer occurring at close range, all atoms of the dimer and OH interact. Thus in (2) addition to two-body terms of V , which is determined by only one atom-atom distance, we consider

three-body V(3) and four-body V(4) contributions, using a procedure similar to the formulation of cluster diagrams described elsewhere.36 For example, one of the two nearest atom-atom interactions belongs to V(3) is the Hd1-Od-O interaction between the donor-monomer atoms and O of OH, V(rHd1Od,rOdO); i.e., a V(rij,rjk) type. Here no interaction between Hd1 and O is included. When the latter interaction is included, the representation is V(rHd1Od,rOdO,rOHd1), a ring pattern of three atom-atom 6 ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28 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

distances, V(rij,rjk,rki), also belongs to the three-body energy V(3). The V(3) is a sum of 25 two atomatom distance-containing terms; i.e., 11 for Hd1, 11 for Hd2 and three not involving Hd1 or Hd2. The latter three are V(rOH,rOdO), V(rOH,rOdH) and V(rOdH,rOdO). The three-body contribution V(3) also contains seven terms of the V(rij,rjk,rki) type mentioned above. The four-body contribution V(4) contains four types. The representative terms of each of the four types with Hd1 are V(rHd1Od,rOdH,rHd1O), V(rHd1Od,rOdH,rHd1O,rHd1H), V(rHd1Od,rOdH,rHd1O,rHd1H,rOdO) and V(rHd1Od,rOdH,rHd1O,rHd1H,rOdO,rOH). Here the first term is dependent on three atom-atom distances as in the terms of V(3), but now contains four atoms. Note that the last representative term is for the case wherein each atom interacts with every one of the three neighbors. When all these terms are counted for both Hd1 and Hd2, the number is 48, but they contribute only slightly to the total potential energy VT = V(2 ) + V(3) + V(4) + VLR. With the coordinates and angles given above, the functional dependence of the total energy is VT = VT(Z,ρ,{r},θ1,φ1,θ2,φ2). The solution of the equations of motion for the inter- and intramolecular coordinates using VT has been described in ref. 11. Brief review of the procedure is that the solution is carried out using Adams-Moulton’s method34 for a 10 000 set of the randomly sampled initial collision energies, relative distances, impact parameters, bond distances and rotational angles at 300K. The integration step is 1/125th of the OH vibrational period, or 0.074 fs. We start the integration from the initial separation of R = 15 Å and follow each trajectory for 100 ps. 3. RESULTS AND DISCUSSION The ensemble-averaged amount of intermolecular transfer of vibrational energy from OH in the first excited state to the dimer stretch is 3132 cm-1, which is 87% of the available energy from OH(v=1) and is very close to the 2ν2 energy 3194 cm-1 of the donor bending mode. Thus the energy transfer out of the stretch to two quanta of intramolecular bend efficiently proceeds, as there is a near7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

resonant pathway.11,24 The probability of the bending vibration gaining two quanta from the stretch is 0.616 at 300 K, the value nearly identical to that reported in ref. 11 for (H2O)2 + H2O. The pathway then is followed by the overtone relaxation to the ground state. When the relaxation completes, the OdHd1 and OdHd2 stretches remain in their ground state. Thus the relaxation process raises the question as to where such a large amount of energy is transferred or stored in the (H2O)2 + OH system. To answer this question, we note that presence of strong attractive interactions between the radical and the dimer lead to a complex-mode collision, forming collision complexes of significant lifetime between Hd1 of the donor monomer and OH. For example, at the most probable distance of energy transfer R = 2.69 Å, the upper limit of the dipole-dipole interaction, the leading part of VLR, is 6.5 kcal mol-1, which is significantly higher than the H-bonding strength between two water molecules. This and other terms of VLR can make important contributions to the interaction of the colliding molecules, forming a complex (H2O)2...OH with hydrogen bonding between the nonbonding Hd1 and the oxygen atom of the radical. When the H-bonding forms, the intermolecular region of the complex now generates low-frequency vibrational modes which store energy and facilitate the relaxation of the overtone bend, the function similar to the dimer H-bonding studied in ref. 11. We find more than 20% of trajectories proceed through complex mode with a lifetime longer than 1 ps. Even some trajectories with a collision time shorter than 1 ps are found to proceed through complex-mode, thus raising the fraction of complex-mode collisions to nearly one-quarter. The complex redissociates to (H2O)2 + OH or fragments to H2O + H2O + OH, where energy transfer ruptures the dimer H-bond as well as the newly formed complex bond; i.e., the bonds redissociate on energy transfer. About 2.5% of the complexes are found to survive for a lifetime longer than 100 ps. A small fraction (100 ps). A small fraction leads to the formation of H2O + H2O…OH. Due to the large disparity in frequencies (∼1600 cm-1 vs ∼200 cm-1), the large amplitude intermolecular vibrations as a single mode is not an efficient energy carrier, but collectively they form a series of small steps for immediate dissipation of the bending energy. During the lifetime of the complex (H2O…HOH…OH), the overtone bending vibration relaxes to the ground state by transferring approximately half of its energy to the low-frequency vibrations of the dimer (H2O…HOH) and the remaining amount to the similar motions in the intermolecular region of the hydrogen-bonded complex (HOH…OH). Therefore, the participation of at least two hydrogen bonds generating low-frequency vibrations is the key to the relaxation of the first overtone bending vibration. Appendix We use the symmetry coordinates for small vibrations of the donor monomer44 17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

S1 = xOd - ½(xHd1 + xHd2), S2 = yOd - ½(yHd1 + yHd2), S3 = (xHd1 - xHd2) and the relations45 mOxOd + mH(xHd1 + xHd2) = 0, mOyOd + mH(yHd1 + yHd2) = 0, mO(xe,OdyOd – ye,OdxOd) + mH(xe,Hd1yHd1 – ye,Hd1xHd1 + x e,Hd2yHd2 – ye,Hd2xHd2) = 0, to express the displacements xOd = ( m H2O /mO)S1, xHd1= - ( m H2O /2mH)S1 + ½S3, xHd2 = -( m H2O /2mH)S1 - ½S3, yOd = ( m H2O /mO)S2, yHd1 = -( m H2O /2mH)[S1cot(½Θe) + S2], yHd2 = ( m H2O /2mH)[S1cot(½Θe) - S2]. In the above expressions, the equilibrium coordinates are xe,Od = 0, xe,Hd1 = - xe,Hd2 = rOdHd1sin(½Θe), ye,Od = ( m H2O /mO)rOdHd1sin(½Θe)cot(½Θe), ye,Hd1 = ye,Hd2 = -( m H2O /2mH)rOdHd1sin(½Θe)cot(½Θe), ze,Od = ze,Hd1 = ze,Hd2 = 0. Shaffer and Newton43 have shown that the symmetry coordinates are related to the normal coordinates in the forms S1 = (1/ m†H2O )1/2Q3, S2 = (1/ m H2O )1/2(c2Q1 + c1Q2), S3 = (2/mH)1/2(c1Q1 - c2Q2), where c1 = 0.810, c2 = 0.586 and m†H2O = m H2O [1 + ( m H2O /2mH)cot2(½Θe)]. For small oscillations, the x-component displacement can be expressed as (xHd1 - xHd2)sin(½Θe) or S3sin(½Θe), whereas the y-component contribution is 2S2cos(½Θe), so that the bond displacement can be expressed as

18 ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28 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

∆rOdHd2 + ∆rOdHd1 = S3sin(½Θe) + 2S2cos(½Θe) and the difference ∆rOdHd2 - ∆rOdHd1 is 2S1( m†H2O /mH2O)sin(½Θe) and the bond angle displacement ∆Θ = [(S3cos(½Θe) - 2S2sin(½Θe)]/rOdHd1. By substituting the symmetry coordinates given above in the latter three expressions, we derive rOdHd2 = a1Q1 + a2Q2 + a3Q3, rOdHd1 = b1Q1 + b2Q2 + b3Q3 and rOdHd2 = c1Q1 +c2Q2, where a1 = b1, a2 = b2 and a3 = -b3. For the relaxation of the overtone bending, it is only necessary to consider the Q2 terms of the non-bonding bond and the bond angle. That is, the coefficients b2 and c2, which take the explicit forms: -1/2

b2 = -0.586(2mH) sin(½Θe) + 0.810 μ -1/2 H 2 O cos(½Θe), 1/2

c2 = -0.586(2/mH) cos(½Θe)/re,OdHd1 - 1.620 μ -1/2 H 2 O sin(½Θe)/re,OdHd1. AUTHOR INFORMATION Corresponding Author *E-mail: [email protected] ORCID Hyung Kyu Shin: 0000-0003-2482-3194 Notes The author declares no competing financial interest. ACKNOWLEDGMENT I thank Professor Matthew Tucker for stimulating discussion on the spectra of water dimers.

19 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 20 of 28

REFERENCES (1) Seta,T.; Yamamoto,M.; Nishioka, M.; Sadakata, M. Structures of Hydrated Oxygen Anion Clusters: DFT Calculations for O-(H2O)n, O2-(H2O)n, and O3-(H2O)n (n = 0 – 4) J. Phys. Chem. A 2003, 107, 962-967. (2) Allodi, M. A.; Dunn, M. E.; Livada, J.; Kirschner, K. N.; Shields, G. S. Do Hydroxyl RadicalWater Clusters, OH(H2O)n, n = 2 - 5, Exist in the Atmosphere? J. Phys. Chem. A 2006, 110, 1328313289. (3) Cringus, D.; Yeremenko, S.; Pshenichnikov, M. S.; Wiersma, D. A. Hydrogen Bonding and Vibrational Energy Relaxation in Water-Acetonitrile Mixtures. J. Phys. Chem. B 2004, 108, 1037610387. (4) Du, S.; Francisco, J. S. The OH Radical - H2O Molecular Interaction Potential. J. Chem. Phys. 2006, 124, 224318. (5) Du, S.; Francisco, J. S. Interaction between OH Radical and the Water Interface. J. Phys. Chem. A. 2008, 112, 4826-4835. (6) Rocher-Casterline, B. E.; Molliner, A. K.; Lee, C. C.; Reisler, H. Imaging H2O Photofragments in the Predissociation of the HCl-H2O Hydrogen-Bonded Dimer. J. Phys. Chem. A 2011, 115, 69036909. (7) Samanta, A. K.; Czako, G.; Wang, Y.; Mancini, J. S.; Bowman, J. M.; Reisler, H. Experimental and Theoretical Investigations of Energy Transfer and Hydrogen-Bond Breaking in small Water and HCl Clusters. Accounts Chem. Res. 2014, 47, 2700-2709. (8) Yuan, R.; Yan, C.; Tamimi, A.; Fayer, M. D. Molecular Anion Hydrogen Bonding Dynamics in Aqueous Solution. J. Phys. Chem. B 2015, 119, 13407-13415. (9) Ree, J.; Kim, Y. H.; Shin, H. K. Dynamics of the Water Dimer + Nitric Oxide Collision. Bull. Korean Chem. Soc. 2017, 38, 196-204. (10) Gao, A.; Li, G.; Peng, B.; Xie, Y.; Schaefer III, H. F. The Water Dimer Reaction OH + (H2O)2 (H2O)-OH + H2O. Phys. Chem. Chem. Phys. 2017, 19, 18279-18287. (11) Shin, H. K. Energy Transfer to the Hydrogen Bond in the (H2O)2 + H2O Collision. J. Phys. Chem. B 2018, 122, 3307-3317. (12) Zittel, P. F.; Masturzo, D. E. Vibrational Relaxation of H2O from 295 to 1020 K. J. Chem. Phys. 1989, 90, 977-987. (13) Shin, H. K. Self-Relaxation of Vibrationally Excited H2O Molecules, J. Chem. Phys. 1993, 98, 1964-1978. (14) Rey, R.; Hynes, J. T. Vibrational Energy Relaxation of HOD in Liquid D2O. J. Chem. Phys. 1996, 104, 2356-2367.

20 ACS Paragon Plus Environment

Page 21 of 28 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

(15) Ingrosso, F.; Rey, R.; Elsaesser, T.; Hynes, J. T. Ultrafast Energy Transfer from the Intramolecular Bending Vibration to Librations in Liquid Water. J. Phys. Chem. A. 2009, 113, 66576665. (16) Rey, R.; Ingrosso, F.; Elsaesser, T.; Hynes, J. T. Pathways for H2O Bend Vibrational Relaxation in Liquid Water. J. Phys. Chem. A. 2009, 113, 8949-8962. (17) Deàk, J. C.; Rhea, S. T.; Iwaki, L. K.; Dlott, D. D. Vibrational Energy Relaxation and Spectral Diffusion in Water and Deuterated Water. J. Phys. Chem. A 2000, 104, 4866-4875. (18) Dlott, D. D. Vibrational Energy Redistribution in Polyatomic Liquids: 3D Infrared-Raman Spectroscopy. Chem. Phys. 2001, 266, 149-166. (19) Larsen, O. F. A.; Woutersen, S. Vibrational Relaxation of the H2O Bending Mode in Liquid Water. J. Chem. Phys. 2004, 121, 12143-12146. (20) Woutersen, S.; Bakker, H. J. Resonant Intermolecular Transfer of Vibrational Energy in Liquid Water. Nature. 1999, 402, 507-509. (21) Piatkowskim L.; Eisenthal, K. B.; Bakker, H. J. Ultrafast Intermolecular Energy Transfer in Heavy Water. Phys. Chem. Chem. Phys. 2009, 11, 9033-9038. (22) Seifert, G.; Patzlaff, T.; Graener, H. Pure Intermolecular Vibrational Relaxation of the OH Bending Mode of Water Molecules. J. Chem. Phys. 2004, 120, 8866-8867. (23) Seifert, G.; Graener, H. Solvent Dependence of OH Bend Vibrational Relaxation of Monomeric Water Molecules in Liquids. J. Chem. Phys. 2007, 127, 224505. (24) Ch′ng, L. C.; Samanta, A. K.; Czako, G.; Bowman, J. M.; Reisler, H. Experimental and Theoretical Investigations of Energy Transfer and Hydrogen-Bond Breaking in the Water Dimer. J. Am. Chem. Soc. 2012, 134, 15430-15435. (25) Ramasesha, K.; De Marco, L.; Mandal, A.; Tokmakoff, A. Water Vibrations Have Strongly Mixed Intra- and Intermolecular Character. Nature Chem. 2013, 5, 935-940. (26) Braunstein, M.; Conforti, P. F. Classical Dynamics of H2O Vibrational Self-Relaxation. J. Phys. Chem. A 2015, 119, 3311-3322. (27) Tsuji, K.; Shibuya, K. Infrared Spectroscopy and Quantum Chemical Calculations of OH(H2O)2 Complexes. J. Phys. Chem. A 2009, 113, 9945-9951. (28) Hernandez, F. J.; Brice, J. T.; Leavitt, C. M.; Liang, T.; Raston, P. L.; Pino, G. A.; Douberly, G E. Mid-infrared Signatures of Hydroxyl Containing Water Cluster: Infrared Laser Start Spectroscopy of OH-H2O and OH(H2O)n (n = 1-3). J. Chem. Phys. 2015, 143, 164304. (29) Computational Chemistry Comparison and Benchmark Database Release 18 (October 2016), Standard Reference Database 101, National Institute of Standards and Technology. Available at http://cccbdb.nist.gov/diplistx.asp#NSRDS-NBS10. Accessed May 14, 2017. Also see, Powell, F. X.; Lide, Jr., D. R. Improved Measurement of the Electric-Dipole Moment of the Hydroxyl Radical. J. Chem. Phys. 1965, 42, 4201-4202. 21 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 22 of 28

(30) Adamowicz, L. Numerical Multiconfigration Self-Consistent Field Study of the Total (Electric and Nuclear) Parallel Polarizability and Hyperpolarizability for the OH, OH+, and OH- J. Chem. Phys. 1988, 89, 6305-6309. (31) Bailey, A. E.; Heard, D. E.; Henderson, D. A.; Paul, P. H. Collisional Quenching of OH(A 2Σ+, v′ = 0) by H2O between 211 and 294 K and the Development of a Unified Model for Quenching. Chem. Phys. Lett. 1999, 302, 132-138. (32) Huber, K. P.; Herzberg, G. Constants of Diatomic Molecules; van Nostrand Reinhold; New York, 1979, p. 508. (33) Bouteiller, Y.; Tremblay, B.; Perchard, J. P. The Vibrational Spectrum of the Water Dimer: Comparison between Anharmonic ab initio calculations and Neon Matric Infrared Data between 14,000 and 90 cm-1. Chem. Phys. 2011, 386, 29-40. (34) Huisken, F.; Kaloudis, M.; Kulcke, A. Infrared Spectroscopy of Small Size-Selected Water Clusters. J. Chem. Phys. 1996, 104, 17-25. (35) Kjaergaard, H. G.; Garden, A. I.; Chaban, G. M.; Matthews, D. A.; Stanton, J. F. Calculation of Vibrational Transition Frequencies and Intensities in Water Dimer: Comparison of Different Vibrational Approaches. J. Phys. Chem. A. 2008, 112, 4324-4335. (36) Ree, J.; Kim, Y. H.; Shin, H. K. Dependence of the Four-Atom Reaction HBr + OH → Br + H2O on Temperature between 20 and 2000 K. J. Phys. Chem. A 2015, 119, 3147-3160. (37) IMSL MATH/LIBRARY, Fortran Subroutines for Mathematical Applications, Version 2.0; IMS: Houston, TX, 1991; DIVPAG, pp. 755-771 to solve an initial-value problem for ordinary differential equations using Adams-Moulton’s method and DRUN, 1319-1320 to generate random numbers from a uniform (0,1) distribution. (38) Mannfors, B.; Palmo, K.; Krimm, S. Spectroscopically Determined Force Field for Water Dimer: Physically Enhanced Treatment of Hydrogen Bonding in Molecular Mechanics Energy Functions. J. Phys. Chem. A 2008, 112, 12667-12678. (39) Ceponkus, J.; Uvdal, P.; Nelander, B. Intermolecular Vibrations of Different Isotopologs of the Water Dimer: Experiments and Density Functional Theory Calculations. J. Chem. Phys. 2008, 129, 194306. (40) Mackeprang, K.; Kjaergarrd, H. G.; Salmi, T.; Hänninen, V.; Halonen, L. The Effect of Large Amplitude Motions on the Transition Frequency Redshift in Hydrogen Bonded Complexes: A Physical Picture. J. Chem. Phys. 2014, 140, 184309. (41) Ayers, G. P.; Pullin, A. D. E. The Infrared Spectra of Matrix Isolated Water Species – I. Assingment of Bands to (H2O)2, (D2O)2 and HDO Dimer Species in Argon Matrices. Spectrochimica Acta. 1976, 32A, 1629-1639.

22 ACS Paragon Plus Environment

Page 23 of 28 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

(42) Bentwood, R. M.; Barnes, A. J.; Orville-Thomas, W. J. Studies of Intermolecular Interactions by Matrix Isolation Vibrational Spectroscopy. Self-Association of Water. J. Mol. Spectrosc. 1980, 84, 391-404. (43) Langford, V. S.; McKinley, A. J.; Quickenden, T. I. Identification of H2O.HO in Argon Matrices. J. Am. Chem. Soc. 2000, 122, 12859-12863. (44) Herzberg, Z. Molecular Spectra and Molecular Structure. II Infrared and Raman Spectra of Polyatomic Molecular; D. van Nostrand; Princeton, NJ, 1968, p. 146. (45) Shaffer, W. H.; Newton, R. R. Valence and Central Forces in Bent Symmetrical XY2 Molecules. J. Chem. Phys. 1941, 10, 405-409.

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 CAPTIONS Fig. 1. Collision model displayed in the cylindrical coordinates (ρ, Z, Φ). The inset shows the relative separation R and the rotational angels (θ1,θ2,φ1,φ2). Fig. 2. Time evolution of the distance between the non-bonding hydrogen and the oxygen of the radical, rOHd1 for selected trajectories. (a) The relative separation R, rOdH and rOH for the directmode collision, (b) R and rOH for the complex-mode collision and (c) long-lived complex. The distance rOHd1 is shown in all three panels, Fig. 3. Power spectrum for the long-lived complex (H2O)2…OH shown in Fig. 2(c).

24 ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28 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

Z

θ1

ρ

φ1, φ2

H

O

OH

R θ2

Hd1

OaOd

Θ Donor monomer

Hd2

Od Y

Φ X

Oa

Ha2

Acceptor monomer

Ha1

Figure 1

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry

10

(a)

8 R

rOdH

6 4

rOHd1

2

rOH

0 -0.2

0.0

0.2

0.4

(b)

Distance (Å)

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 26 of 28

8 6 rOHd1

R 4 2

rOH

0 0.0

10

0.2

0.4

0.6

0.8

8

10

(c) 8 6 4

rOHd1

2 0 0

2

4

6

Time (ps) 26 ACS Paragon Plus Environment

Figure 2

Page 27 of 28

266 226

Intensity

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

3596 856

0

1000

2000

3000

Wavenumber (cm-1)

27 ACS Paragon Plus Environment

4000 Figure 3

The Journal of Physical Chemistry 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 28 of 28

TOC Graphic

8

v2=2

rOHd1

6 v2=1

4 Radical

v2=0

2

H-bond Acceptor

R rOH

Å H-bond

Donor

0

0.0ps

28 ACS Paragon Plus Environment

0.2

0.4

0.6