COMMUNICATIONS The bifurcation rearrangement in ... - Berkeley

0 downloads 0 Views 73KB Size Report
May 22, 2006 - The bifurcation rearrangement in cyclic water clusters: Breaking ... 2.7 THz established the time scale for bifurcation rearrangements to be approximately 40 ns. ... energy barrier for the bifurcation rearrangement in the pen-.
JOURNAL OF CHEMICAL PHYSICS

VOLUME 109, NUMBER 22

8 DECEMBER 1998

COMMUNICATIONS The bifurcation rearrangement in cyclic water clusters: Breaking and making hydrogen bonds M. G. Brown, F. N. Keutsch, and R. J. Saykallya) Department of Chemistry, University of California Berkeley, Berkeley, California 94720

~Received 31 August 1998; accepted 30 September 1998! Tunneling patterns observed in the vibration–rotation–tunnelling spectrum of ~H2O!5 measured near 2.7 THz established the time scale for bifurcation rearrangements to be approximately 40 ns. This relatively local process is likely to be relevant in the dynamics of liquid water and ice. © 1998 American Institute of Physics. @S0021-9606~98!02146-1#

ing unactivated ~T50 K! process in the condensed phases of water. Walsh and Wales ~WW!7 predicted that bifurcation tunneling would be too small to resolve in ~D2O!5, but that it should indeed be observable in ~H2O!5. Their group theoretical analysis predicted a multiplet pattern of six lines with intensity ratios of 1:3:18:54:81:51 for VRT transitions having angular momentum projection quantum number K equal to a multiple of 5 and 0:3:18:54:81:48 for all other VRT transitions. These intensity ratios were computed from nuclear spin weights of the irreducible representations of the permutation–inversion group G 320 . The magnitude of the tunneling splittings was predicted to be 0.2 GHz from diffusion quantum Monte Carlo calculations by Gregory and Clary,12 using the ASP–NB water potential. We recently observed a VRT spectrum of ~H2O!5 near 90 cm21 @2.670 303 7~6! THz# that revealed the spectral splitting pattern predicted by WW ~Fig. 2!. A total of 200 individual transitions were precisely measured ~63.13 MHz! and fit to a standard symmetric top energy-level expression @Eq. ~1!# including first-order Coriolis perturbations. Separate centrifugal distortion constants were used for the 6 components of the Coriolis perturbations. In recent theoretical studies of the water trimer by van der Avoird14 it was shown that these ‘‘first-order’’ Coriolis effects actually result from second-order coupling between hydrogen torsion and the overall rotation of the water trimer, and a similar mechanism may be operative in the case of the water pentamer.

Rearrangements of the hydrogen-bond network in water clusters are manifested as intricate quantum mechanical tunneling splittings in their vibration–rotation–tunneling ~VRT! spectra. Detailed study of these tunneling dynamics in water clusters can provide new insight into the corresponding thermally activated rearrangement processes that occur in condensed phases. The term ‘‘bifurcation’’ describes a hydrogen-bond network rearrangement ~HBNR! observed in most of the water clusters studied so far by high-resolution VRT spectroscopy viz. the water dimer, trimer, and probably hexamer although explicitly not observed in the water tetramer.1–6 In this process, the hydrogen bonds on a particular water monomer are broken and then reformed by quantum tunneling through a transition state lying ;2 kcal/mol above the potential minimum and having a bifurcated hydrogen-bond arrangement7 ~Fig. 1!. Here, we present the results of a new study of the water pentamer in which the corresponding bifurcation rearrangement has been characterized for the first time. The results have interesting implications for the dynamics of solid and liquid water. While the bifurcation rearrangement has been thoroughly characterized in the H2O, D2O, and mixed isotope forms of the water dimer8–10 and water trimer,4,11 it was not observed in previous studies of the water pentamer. Both calculations13 and experiments6 have recently shown that the properties of the hydrogen-bond networks in small water clusters converge rapidly to condensed phase values, such that the water pentamer already exhibits H-bond lengths, monomer dipole moments, and bond energies comparable to average values found in liquid water and ice. The potentialenergy barrier for the bifurcation rearrangement in the pentamer can be expected to correspond approximately to that of double hydrogen-bonded monomers present in the condensed phases and to be a lower limit for those with more than two hydrogen bonds. The effective mass and path length should remain approximately constant with cluster size, i.e., it is a relatively local process. Characterization of the bifurcation rearrangement in the water pentamer will, therefore, help to constrain the time scale for the correspond-

E ~ J,K ! 5BJ ~ J11 ! 1 ~ C2B ! K 2 2D j @ J ~ J11 !# 2 2D jk @ J ~ J11 ! K 2 # 2DkK 4 62C z K.

Each VRT transition was split by bifurcation tunneling into six lines separated by 4.8 MHz and having the predicted relative intensities ~Fig. 2!. However, the small nuclear spin weights of the weakest two of these tunneling components precluded their observation. The water pentamer is predicted by ab initio theory13 to have a puckered pentagonal global equilibrium structure similar to that of cyclopentane. The results of the spectral analysis show that this asymmetric equilibrium structure

a!

Electronic mail: [email protected]

0021-9606/98/109(22)/9645/3/$15.00

~1!

9645

© 1998 American Institute of Physics

Downloaded 22 May 2006 to 128.32.220.140. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

9646

J. Chem. Phys., Vol. 109, No. 22, 8 December 1998

Communications

FIG. 2. VRT spectrum of (H2O!5 measurement near 89 cm21. A stick spectrum representation of the Q branch. The repelling Q branches clearly show the influence of linear Coriolis mixing of degenerate torsional states, while the inset shows the bifurcation tunneling splittings present in each line in the spectrum. The rotational constants B 8 and B 9 extracted from this spectrum are 1989.45~7! and 1992.46~7! MHz, respectively, and the vibrational band origin is 2.6703037~6! THz. This spectrum was taken with the Berkeley terahertz spectrometer, which can be briefly described as follows. A linetunable CO2 laser pumps a terahertz laser and the resulting fixed frequency light is mixed with tunable microwave radiation on a Schottky barrier diode to produce tunable sidebands of frequency n ~laser! 6n ~microwave!. These tunable sidebands are multipassed in the throat of a pulsed planar supersonic expansion of argon and H2O and detected with a stressed germanium/ gallium photoconductive detector. FIG. 1. The bifurcation rearrangement in the water pentamer. The ab initio barrier height for this rearrangement is ;2 kcal/mol in both the dimer and trimer ~Ref. 23!, while it increases somewhat in larger clusters due to the cooperative interactions ~many-body forces! which produce stronger hydrogen bonds. Note the accompanying torsion of the water molecule adjacent to that undergoing exchange, which allows the oxygen lone pairs to align with the bifurcating hydrogens, an energetically more favorable pathway. The time scale for this process, estimated from the measured 4.6 MHz tunneling splittings is ;431028 s. This splitting is actually the sum of tunneling splitting in the upper and lower stages of the transition.

~which, as for the water trimer, is chiral! vibrationally averages to that of a symmetric top. This vibrational averaging of the oxygen framework was previously observed in the D2O form of the water pentamer,6 and thus was expected to occur in the H2O form as well. It seems likely that the many lowfrequency vibrational modes predicted by ab initio theory13 ~ten normal modes are predicted below 200 cm21) provide a mechanism for averaging the oxygen framework to a quasiplanar structure, while free-hydrogen ‘‘flipping’’ dynamics similar to those characterized in the water trimer4 will average the hydrogen positions. The resulting average structure has C 5h symmetry. Explicit vibrational assignment of the observed transition is difficult at this time since there are so many low-frequency modes. The Coriolis perturbations observed, through comparison with perturbations observed in the water trimer, are consistent with assignment to a transi-

tion between doubly degenerate torsional energy levels.14 The bifurcation tunneling motion described for the water pentamer in Fig. 1 has now been observed in the water dimer, trimer, pentamer, and hexamer, and specifically, is not observed in the water tetramer.1,3–6 In the water dimer the bifurcation tunneling motion has typically been referred to as ‘‘donor tunneling’’ as it exchanges the role of the bound and free protons on the donor water molecule.9 The motion can be described as a geared rotation of both the donor and acceptor water molecules about axes perpendicular to the plane of symmetry of the cluster. The barrier to this motion has been predicted by ab initio theory15 to be 658 cm21 without zero-point correction. The tunneling motion is manifested in spectral shifts of approximately 1 GHz in the relative energies of the various symmetry components of each vibrational band. These shifts are dependent on the approximate quantum number K a of the transition. Shifts, rather than splittings, occur because the other tunneling motions of the dimer ~acceptor switching and donor–acceptor interchange! allow the same rearrangements to occur through different mechanisms with lower barriers, effecting the maximum degree of splitting allowed by symmetry. Bifurcation rearrangement in the water trimer has been studied in detail by van der Avoird et al.16,17 A comparison of their theory and our experimental results revealed that the tunneling motion for this process

Downloaded 22 May 2006 to 128.32.220.140. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

J. Chem. Phys., Vol. 109, No. 22, 8 December 1998

requires a neighboring water molecule to ‘‘flip’’ its free proton from one side of the cluster to the other via torsion about the H bond. This motion can be viewed as an extension of the ‘‘geared’’ motion observed in the water dimer and is similar to the mechanism shown for the pentamer in Fig. 1. The barrier for this process has been calculated by ab initio methods18 to be 823 cm21 and results in a quartet splitting ~expressed as the difference between upper and lower state tunneling splittings in our spectra ! of ;289 MHz. Terahertz spectra of the H2O and the D2O forms of the water tetramer have revealed no tunneling splittings attributed to bifurcation tunneling.3 The cause of this is probably the inability of the highly symmetric tetramer structure to undergo a geared bifurcation pathway. Rather, such tunneling must arise from a pure C2 exchange pathway, a pathway that exhibits an unobservably high barrier to exchange. The global minimum ~after zero-point energy corrections! of the water hexamer is a three-dimensional cage structure with constituent water molecules in several distinct donor–acceptor roles. Four of the molecules are involved in three hydrogen bonds, while two are doubly bonded. The two doubly bonded water molecules undergo a hydrogen-bond rearrangement by an as yet unknown pathway. Recent work by Wales in ambiguous on this topic, bifurcation and nonbifurcation pathways being approximately equally weighted.19 No exchange motions have yet been observed for the triply bonded water molecules. An interesting comparison can be made between bifurcation rearrangements in small water clusters and similar hydrogen-bond rearrangement processes occurring in liquid water and ice. For example, it is known that the protondisordered phases of ice have large dielectric permittivities20 ~e597.5 for ice Ih to e5193 for ice VI!, larger even than for liquid water ( e ;88). These values are 30–50 times larger than the dielectric constants for the proton-ordered phases ~ice II and ice IX!. This suggests that some highly effective relaxation mechanism is available to proton-disordered ice, allowing it to reorient efficiently in an electric field. It has been proposed that defects in the ice lattice are the main contribution to this reorganization.21 In such a model, phonons in the ice lattice cause hydrogen bonds to break or weaken. With a sufficient number of hydrogen bonds broken, the water molecules in the ice lattice can reorient by tunneling, perhaps through a bifurcated transition state. These ideas are supported by temperature-dependent quantummechanical calculations of the dielectric relaxation time of ice Ih by Bruni et al.22 They found that quantum tunneling was the dominant reorganizational mechanism contributing to the dielectric relaxation time at temperatures below 230 K. Significantly, they found that quantum corrections to classical rotations were required to explain the experimental data at all temperatures. The calculated relaxation times ranged from 102 to 104 ms over the temperature range 260–210 K. This time frame is consistent with the bifurcation tunneling time in the water pentamer ~;40 ns!, estimated from our

Communications

9647

tunneling splitting of 4.8 MHz. Although the intermediate states in the reorganization of ice are still unknown, it is possible that bifurcation rearrangement is an essential feature of this process. This certainly deserves further investigation. Other examples wherein bifurcation tunneling may affect the bulk properties of liquid water and ice are known, viz. depolarized Rayleigh Raman scattering23,24 and quasielastic neutron scattering experiments.25 In these experiments on liquid water, classical reorganization is more dominant than in the solid. Reorganization times are ;ps, comparable to hindered rotation in liquid water. Even in these systems, however, quantum reorganization terms ~often referred to as ‘‘jump diffusion’’ terms! are necessary to fit experimental data to experimental precision. This tunneling reorganization process may well involve bifurcation tunneling. In any case, the exact role of the bifurcation rearrangement in the dynamics of water and ice remains an interesting problem. We hope that the results reported here facilitate further efforts. This work was supported by the Experimental Physical Chemistry program of the National Science Foundation. 1

K. L. Busarow, R. C. Cohen, G. A. Blake, K. B. Laughlin, and R. J. Saykally, J. Chem. Phys. 90, 3937 ~1989!. 2 R. J. Saykally and G. A. Blake, Science 259, 1570 ~1993!. 3 J. D. Cruzan, L. B. Braly, K. Liu, M. G. Brown, and R. J. Saykally, Science 271, 59 ~1996!. 4 K. Liu, J. G. Loeser, M. J. Elrod, B. C. Host, and R. J. Saykally, J. Am. Chem. Soc. 116, 3507 ~1994!. 5 K. Liu, M. G. Brown, C. Carter, R. J. Saykally, J. K. Gregory, and D. C. Clary, Nature ~London! 381, 501 ~1996!. 6 K. Liu, M. G. Brown, J. D. Cruzan, and R. J. Saykally, Science 271, 62 ~1996!. 7 D. J. Wales and T. R. Walsh, J. Chem. Phys. 105, 6957 ~1996!. 8 T. R. Dyke, J. Chem. Phys. 66, 492 ~1977!. 9 G. T. Fraser, F. J. Lovas, R. D. Seunram, E. N. Karyakin, A. Grushow, W. A. Burns, and K. R. Leopold, J. Mol. Spectrosc. 181, 229 ~1997!. 10 E. Zwart, J. J. ter Meulen, W. L. Meerts, and L. H. Coudert, J. Mol. Spectrosc. 147, 27 ~1991!. 11 M. R. Viant, J. D. Cruzan, D. D. Lucas, M. G. Brown, and R. J. Saykally, J. Phys. Chem. A 101, 9032 ~1997!. 12 J. K. Gregory and D. C. Clary, J. Chem. Phys. 105, 6626 ~1996!. 13 S. S. Xantheas and T. H. Dunning, Jr., J. Chem. Phys. 99, 8774 ~1993!. 14 M. R. Viant, M. G. Brown, J. D. Cruzan, R. J. Saykally, A. van der Avoird, and Michel Geleijns ~submitted!. 15 B. J. Smith et al., J. Chem. Phys. 92, 1240 ~1990!. 16 A. van der Avoird, E. H. T. Olthof, and P. E. S. Wormer, J. Chem. Phys. 105, 8034 ~1996!. 17 A. van der Avoird, E. H. T. Olthof, and P. E. S. Wormer, J. Chem. Phys. 105, 8051 ~1996!. 18 D. J. Wales, J. Am. Chem. Soc. 115, 11180 ~1993!. 19 D. J. Wales, in Advances in Molecular Vibrations and Collision Dynamics, edited by Z. Bacic and J. Bowman ~JAI, Greenwich, CT, 1997!. 20 G. W. S. Robinson, S. Zhu, S. Singh, and M. W. Evans, Water in Biology, Chemistry and Physics ~World Scientific, Singapore, 1996!. 21 N. Bjerrum, Science 115, 385 ~1952!. 22 F. Bruni, G. Consolini, and G. Careri, J. Chem. Phys. 99, 538 ~1993!. 23 O. Conde and J. Teixeira, Mol. Phys. 53, 951 ~1984!. 24 O. Conde and J. Teixeira, J. Phys. ~France! 44, 525 ~1983!. 25 F. Cavatorta, A. Deriu, D. Di Cola, and H. D. Middendorf, J. Phys. Condens. Matter 6, 113 ~1994!.

Downloaded 22 May 2006 to 128.32.220.140. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp