multiphoton ionization studies of c6h6--(ch3oh)n clusters ... · photophysics of various...

14
Multiphoton ionization studies of C, H,-(CH, OH),, clusters. I. Comparisons with C, H,-(H, 0), clusters Aaron W. Garrett, Daniel L. Severance,‘ ) and Timothy S. Zwierb)vC) Department of Chemistry, Purdue University, WestLafayette, Indiana 47907-1393 (Received 7 November 199 1; accepted7 February 1992) Resonant two-photon ionization (R2PI) scansof the S,-S, spectra of C, H6-( CH, OH) n clusters with n = l-5 have been recorded. These scansprovide an interesting comparison with earlier spectra from our laboratory on C, H, -( H, 0) n clusters. A variety of vibronic level arguments are used to constrain the geometries of the C!,H, -( CH, OH) n clusters. The 1:1 and 1:2 clusters possess vibronic level features which are very similar to their aqueous counterparts. The 1:1 cluster places the methanol molecule in a P hydrogen-bonded configuration on or near the sixfold axis of benzene. The spectral characteristics of the 1:2 cluster are consistent with both methanol moleculesresiding on the sameside of the benzene ring as a methanol dimer. Higher C, H, -( CH, OH) n clusters show distinct differences from the corresponding C!,H, -( H, 0) n clusters. Vibronic level arguments lead to the following conclusions: the methanol molecules in the 1:3 cluster show the strongest hydrogen bonding to the ?rcloud of any of the clusters and attach to benzene in such a way as to strongly break the sixfold symmetry of its r cloud. The 1:4 clusters are at most only very weakly hydrogen bonded to the P cloud, break benzene’s sixfold symmetry moderately well, and possess strong activity in a very low frequency intermolecular mode. The methanol moleculesin the 1:5 cluster show no hydrogen-bondinginteraction with benzene’s r cloud, induce remarkably little asymmetry in the P electron density, and produce very little van der Waals’ activity. Monte Carlo simulations using intermolecular potentials developed for liquid simulations serve as a guide to the possibleminimum-energy structures for the clusters. The experimental results are used to distinguish betweenthe possiblestructures. In all cases, the lowest energy structures produced by the calculations satisfactorily fit the vibronic level constraints placed on the structures by our data. I. INTRODUCTION Over the past severalyears, a number of groups,l-i9 in- cluding our own, 5020-22 have studied the spectroscopy and photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’ -8 since they serve as useful model systems for the study of weak intermolecular forces. Recently, attention has increas- ingly turned to more chemically relevant solvents,‘ -** but the increasing complexity of thesesolvents adds a new level of challengeto the detailed structural characterization of the clusters. We have recently taken a somewhat different ap- proach to the study of larger clusters containing a single benzenemolecule. In this case,the high symmetry of ben- zene and the forbidden nature of its S’ ,,-S’ , origin produce vibronic level probes of the structures of the clusters.2G22 Admittedly these constraints are loose ones. Nevertheless, they provide a global view of the structures of theseclusters for which the acquisition and analysis of fully rotationally resolvedspectra are exceedinglyformidable tasks. In a recent seriesof papers,** we used such a combina- tion of rotational band contour analysis and vibronic level “Present address: Department ofchemistry, Yale University, New Haven, CT 06511. ) Alfred P. Sloan Research Fellow. ) Author to whom correspondence should be addressed. argumentsto study C!,H, -( H, 0) n clusters containing from one to five water molecules.Thesespectra point consistently to the clusters being composed of hydrogen-bondedwater networks situated on one face of the benzene ring. The pic- ture which is emerging is thus one in which the strong bind- ing between water molecules is the major determinant of cluster structure, with the benzene molecule relegated (for clusters containing up to five water moleculesat least) to the outskirts of the cluster due to its weaker rr hydrogen bonding to the water cluster. In this paper, we report similar spectroscopic measure- ments on C,H,-(CH, OH) n clusters with an eye toward comparison with those on C, H, -( H, 0) n. The substitution of CH, OH for H, 0 has several interesting consequences. First, despite the strong hydrogen bonding betweenmetha- nol molecules,benzene is very soluble in methanol, in nota- ble contrast to the immiscibility of benzene and water. Sec- ond, each CH,OH molecule is capable of acting only as a single hydrogen donor to its neighbors, whereas H,O can donate two hydrogens to hydrogen bonds. This is borne out by calculations on pure methanol clusters which predict the lowest energy structure for clusters containing up to 11 methanols to be a closed ring in which each methanol is involved in hydrogen bonding as single donor and accep- tor.23 By contrast, (H,O), clusters are calculated to form many similar-energy, compact, ice-like networks for cluster sizes containing six or more water molecules.24 Third, the J. Chem. Phys. 96 (lo), 15 May 1992 0021-9606/92/l 07245-I 4$06.00 0 1992 American Institute of Physics 7245 Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Upload: others

Post on 20-Jun-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

Multiphoton ionization studies of C, H,-(CH, OH),, clusters. I. Comparisons with C, H,-(H, 0), clusters

Aaron W. Garrett, Daniel L. Severance,‘) and Timothy S. Zwierb)vC) Department of Chemistry, Purdue University, West Lafayette, Indiana 47907-1393

(Received 7 November 199 1; accepted 7 February 1992)

Resonant two-photon ionization (R2PI) scans of the S,-S, spectra of C, H6-( CH, OH) n clusters with n = l-5 have been recorded. These scans provide an interesting comparison with earlier spectra from our laboratory on C, H, -( H, 0) n clusters. A variety of vibronic level arguments are used to constrain the geometries of the C!, H, -( CH, OH) n clusters. The 1: 1 and 1:2 clusters possess vibronic level features which are very similar to their aqueous counterparts. The 1: 1 cluster places the methanol molecule in a P hydrogen-bonded configuration on or near the sixfold axis of benzene. The spectral characteristics of the 1:2 cluster are consistent with both methanol molecules residing on the same side of the benzene ring as a methanol dimer. Higher C, H, -( CH, OH) n clusters show distinct differences from the corresponding C!, H, -( H, 0) n clusters. Vibronic level arguments lead to the following conclusions: the methanol molecules in the 1:3 cluster show the strongest hydrogen bonding to the ?r cloud of any of the clusters and attach to benzene in such a way as to strongly break the sixfold symmetry of its r cloud. The 1:4 clusters are at most only very weakly hydrogen bonded to the P cloud, break benzene’s sixfold symmetry moderately well, and possess strong activity in a very low frequency intermolecular mode. The methanol molecules in the 1:5 cluster show no hydrogen-bonding interaction with benzene’s r cloud, induce remarkably little asymmetry in the P electron density, and produce very little van der Waals’ activity. Monte Carlo simulations using intermolecular potentials developed for liquid simulations serve as a guide to the possible minimum-energy structures for the clusters. The experimental results are used to distinguish between the possible structures. In all cases, the lowest energy structures produced by the calculations satisfactorily fit the vibronic level constraints placed on the structures by our data.

I. INTRODUCTION

Over the past several years, a number of groups,l-i9 in- cluding our own, 5020-22 have studied the spectroscopy and photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’-8 since they serve as useful model systems for the study of weak intermolecular forces. Recently, attention has increas- ingly turned to more chemically relevant solvents,‘-** but the increasing complexity of these solvents adds a new level of challenge to the detailed structural characterization of the clusters. We have recently taken a somewhat different ap- proach to the study of larger clusters containing a single benzene molecule. In this case, the high symmetry of ben- zene and the forbidden nature of its S’,,-S’, origin produce vibronic level probes of the structures of the clusters.2G22 Admittedly these constraints are loose ones. Nevertheless, they provide a global view of the structures of these clusters for which the acquisition and analysis of fully rotationally resolved spectra are exceedingly formidable tasks.

In a recent series of papers,** we used such a combina- tion of rotational band contour analysis and vibronic level

“Present address: Department ofchemistry, Yale University, New Haven, CT 06511.

‘) Alfred P. Sloan Research Fellow. ‘) Author to whom correspondence should be addressed.

arguments to study C!, H, -( H, 0) n clusters containing from one to five water molecules. These spectra point consistently to the clusters being composed of hydrogen-bonded water networks situated on one face of the benzene ring. The pic- ture which is emerging is thus one in which the strong bind- ing between water molecules is the major determinant of cluster structure, with the benzene molecule relegated (for clusters containing up to five water molecules at least) to the outskirts of the cluster due to its weaker rr hydrogen bonding to the water cluster.

In this paper, we report similar spectroscopic measure- ments on C,H,-(CH, OH) n clusters with an eye toward comparison with those on C, H, -( H, 0) n. The substitution of CH, OH for H, 0 has several interesting consequences. First, despite the strong hydrogen bonding between metha- nol molecules, benzene is very soluble in methanol, in nota- ble contrast to the immiscibility of benzene and water. Sec- ond, each CH,OH molecule is capable of acting only as a single hydrogen donor to its neighbors, whereas H,O can donate two hydrogens to hydrogen bonds. This is borne out by calculations on pure methanol clusters which predict the lowest energy structure for clusters containing up to 11 methanols to be a closed ring in which each methanol is involved in hydrogen bonding as single donor and accep- tor.23 By contrast, (H,O), clusters are calculated to form many similar-energy, compact, ice-like networks for cluster sizes containing six or more water molecules.24 Third, the

J. Chem. Phys. 96 (lo), 15 May 1992 0021-9606/92/l 07245-I 4$06.00 0 1992 American Institute of Physics 7245 Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 2: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

methyl groups are more bulky and polarizable than the hy- drogen atoms of H, 0, adding steric constraints and disper- sive forces to the interaction with other methanol molecules and with benzene which are not present in H,O. Finally, many of the tunneling mechanisms present in water clusters will be turned off in (CH, OH) R clusters, while others, in- volving methyl internal rotation, will be present.

0:/6; intensity ratios must be taken as upper bounds on the true intensity ratios due to the potential for saturation ef- fects.

Ill. RESULTS AND DISCUSSION

We will see that the smallest CgH6-(CH30H),, clus- ters have a vibronic spectroscopy quite similar to that of the corresponding C!, H, -( H, 0) n clusters. Larger C!, H,- (CH, OH) n clusters, on the other hand, show striking dif- ferences with their aqueous counterparts. Monte Carlo sim- ulations using previously optimized intermolecular poten- tials25v26 predict minimum energy structures falling into one of two broad classes involving either chain or cyclic hydro- gen-bonded CH,OH clusters attached to benzene. The vi- bronic level features of the clusters distinguish between the structural possibilities presented by the calculations, provid- ing insight to the very different ways in which methanol and water molecules interact with the benzene ring. In the pro- cess of the experimental work, we have observed a rich, size- dependent intracluster ion-molecule chemistry following photoionization of the neutral clusters. This ion-molecule chemistry is the subject of the adjoining paper (Paper II).

A. C, H,-CH, OH and C, H,-(CH, OH), 1. Assignment of spectral features to a given cluster size

Figure 1 (a) presents one-color R2PI scans of the re- gions just to the blue of the 6: transition of C6 H, monitoring the (C, H,-CH, OH) + mass in the TOF mass spectrum. The assignment of the observed transitions to a given cluster size is complicated by the efficient fragmentation these rr hydrogen-bonded complexes undergo following photoioni-

II. EXPERIMENT The molecular beam reflectron time-of-flight (TOF) I

mass spectrometer used in these studies has been described previously.20S27 C, H, -( CH, OH) n clusters are formed by expanding a mixture containing C6 H, and CH, OH in heli- I:1 um from a pulsed valve of 0.8 mm diam operating at 20 Hz. The concentrations of these vapors are controlled by meter- ing flows of helium over the room temperature liquids using needle valves and mixing these flows with the main flow of

hl&/

helium. Typical expansion conditions employ - 0.5% C, H, I:2 and 0.1%-0.4% CH, OH at a total pressure of 2-4 bar. The - -I

clusters are resonantly ionized by the unfocused output of an excimer-pumped dye laser operating on Coumarin 503, dou- bled in a /?-barium borate crystal. Typical per pulse energies of 0.1-1.0 mJ/pulse are used. Mass-selected resonant two- photon ionization (R2PI) scans are recorded in a linear TOF mass spectrometer using a 100 MHz digital oscillo- scope.

The ratio of intensities of the origin and 6; transitions are reported in this work for the C, H6-( CH, OH) ,, clusters. These ratios are recorded by integrating the intensity of the origin and 6: transitions of a given cluster in order to avoid nonresonant contributions to the ion intensity. Contribu- tions from more than one mass channel are summed in cases where the resonant features are present in more than one mass channel by virtue of fragmentation. A single determin- ation involves recording the spectra of all cluster sizes simul- taneously in a single scan with origin and 6; regions scanned immediately after one another to minimize the effects of changes in expansion or laser conditions. In order to mini- mize saturation effects, all spectra are recorded under unfo- cused laser conditions at peak powers of about 3 X 10’ W/cm2. Despite these modest laser powers, the reported

5 I, I I I I, I I I I, I I, I,,

50 100 150 200 Relative Frequency (cm-‘)

FIG. 1. Resonant two-photon ionization (R2PI) scans in the 6; region monitoring the (a) (C,H,-CH30H) + and (b) (C,H,-H,O)+ mass channels. The 1:l and 19 designations denote the numberofbenzene:meth- an01 or benzene:water molecules in the cluster responsible for the given ab- sorption feature. The zero of the relative frequency scale is the 6: transition of free C, H, .

7246 Garrett, Severance, and Zwier: CBH,-(CH,OH), clusters. I

J. Chem. Phys., Vol. 96, No. IO,15 May 1992

I:2

(b)

Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 3: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

Garrett, Severance, and Zwier: C,HB(CH,OH), clusters. 1 7247

zation. High fragmentation efficiencies have been observed, even under extremely mild laser conditions, for C,H,-X complexes with other hydrogen-bonding solvents such as HCI,” CHC13 ,27 and ( H20),.22 Furthermore, this frag- mentation is difficult to remove in two-color scans due to poor Franck-Condon factors to regions of the ionic poten- tial energy surface below the dissociation threshold. We would expect similar fragmentation to plague the present studies as well, necessitating careful attention in making as- signments of features to a given neutral cluster size. As stressed in the Introduction, our recent studies of C, H,-( H, 0) n clusters are especially helpful as a backdrop to these studies.” They will serve here as an aid in the assign- ments and later as an aid in the interpretation of the spectra.

The R2PI scan in the 6: region monitoring the 1: 1 (C, H,-CH, OH) + channel [Fig. 1 (a) ] shows strong simi- larities to the spectrum of the (C,H,-H,O) + channel shown in Fig. 1 (b). Both spectra exhibit similar frequency shifts of the main features and similar low frequency struc- ture built on the blue-shifted peaks.

The analogous R2PI scans in the origin region monitor- ing the 1:l and 1:2 mass channels are shown in Figs. 2(a) and 2 (b), respectively. These spectra are also very reminis- cent of corresponding spectra in C, H6-( H, 0) n. First, the transition + 44 cm - ’ (relative to C,H, 6; ) in Fig. 1 (a) is not observable at the origin [Fig. 2(a) 1, while those begin- ning at + 80 cm - ’ are readily observed. Second, the set of transitions near + 80 cm - ’ is observed very weakly in the 1:2 mass channel [Fig. 2(b) 1, indicating that these transi- tions are actually due to the 1:2 cluster fragmenting with high efficiency ( - 80%) into the I:1 mass channel upon photoionization. This is precisely what is observed in C,H,-(H,O),, as shown in Fig. 2(c), where the spectrum in the origin region monitoring the 1:l H,O mass channel has been proven to be due entirely to the fragmenting 1:2 H, 0 cluster.22

The transition at + 44 cm - ’ in Fig. 1 (a) is very differ- ent in its behavior than the peaks beginning at + 80 cm- ‘, not only in its weaker origin intensity, but also in its much more efficient fragmentation into the (C, H, > + mass chan- nel. Unfortunately, it was not possible to find expansion con- ditions where the 1:2 cluster was small by comparison to this peak. Nevertheless, we assign the + 44 cm-’ transition to the C, H,-CH, OH complex, in direct analogy to the assign- ment of the + 50 cm - r transition of Fig. 1 (b) to the C, H,-H, 0 complex.

2. Wibronic levelprobes of the structures of the L-1 and i:2 clusters

The similar spectral characteristics of C, H,-CH, OH and C,H,-(CH,OH), clusters to those of C,H,-H,O and C, H6-( H, O), lead us to expect similar geometries for the corresponding clusters. In our recent work on the C, H,-H, 0 complex,22 we used a combination of rotational band contour analysis and vibronic level arguments to show that the water molecule is near the sixfold axis of C,H,, internally rotating about this axis. The hydrogen(s) of the water molecule are pointed in toward the benzene r cloud,

1:2 CH,CH (a)

I I , , I t I I 8 I , I I , I I:3 CH,OH

(c) l-l

I

1:2 H,O

I ” ” I ” “I”“1 50 100 150 200

Relative Frequency (cm”)

FIG. 2. RZPI scans in the S,,-S, origin region monitoring (a) the (C,H,-CH,OH) + mass; (b) the [C,H,-(CH,OH), ] + mass; and (c) the (C,H,-H,O) + mass for comparison. Note that for both C,H,-(CH,OH), and C,H,-(H,O),, all the features present at the ori- gin in the 1 :l mass channel are due to the 1:2 cluster fragmenting with high efficiency into the 1 :l mass channel. The zero of the relative frequency scale marks the position of the forbidden origin transition of free C, H,.

giving rise to the efficient fragmentation and blue-shifted absorption observed.22V27 Due to the lack of signal present in the C&He-CH,OH cluster, its rotational band contour was not recorded. At the present experimental resolution, the rotational band contours of larger clusters also provided lim- ited structural information. As a result, we must seek to con- strain the geometries of the clusters using vibronic level ar- guments. Such arguments have been outlined in some detail in our study of C,H,-(H20),22 and C,Hg-CC142’ clus- ters. We summarize them for the C, H,-( CH, OH) n clus- ters with n = 1,2 below.

a. Frequency shifts. The accumulated experience of workers investigating a large number of aromatic-X com- plexes has led to the generalization that complexes which are hydrogen bonded to the benzene rr cloud produce absorp- tions which are blue shifted from those of the parent aroma-

J. Chem. Phys., Vol. 96, No. lo,15 May 1992 Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 4: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

tic.‘8-22*27 For example, C,H,-HCl, -CHCl,, and -H,O possess frequency shifts of + 125, + 179, and + 50 cm - ‘, respectively.27 Purely dispersive interactions, on the other hand, typically give rise to red-shifted transitions (e.g., C, H,-Ar and C, H,-Ccl, have transitions shifted by - 8 and - 68 cm - ’ from that of C, H, ) . Thus the blue-shifted absorptions of the 1:l and 1:2 clusters ( + 44 and + 80 cm - ’ from the transitions of C,H, ) suggest a hydrogen- bonding interaction of the methanol(s) with the benzene rr cloud.

b. Ejicientfragmentation upon photoionization. Consis- tent with the frequency shift, the efficient fragmentation fol- lowing photoionization also supports a rr hydrogen-bonding interaction in the clusters. In a manner analogous to C 6 H,-HC1,20*27 the fragmentation of the ionized complex results from vertical ionization to a repulsive part of the ionic potential energy surface in which the positive end of the methanol dipole is initially oriented toward the newly creat- ed positive charge on the benzene ring.

c. van der Waals’ structure. The positions and intensities of van der Waals structure built on a given transition reflect changes in the geometry of the cluster which accompany electronic excitation on benzene. The spectrum of Fig. 1 shows little van der Waals structure in the 1:l cluster, though interference from the 1:2 features inhibits clear iden- tification of such structure. By contrast, the 1:2 cluster pos- sesses a number of very strong van der Waals transitions, indicating that the 1:2 cluster is more sensitive to changes in electronic excitation in benzene than is the 1: 1 complex. Ta- ble I lists the relative frequencies and intensities of the van der Waals structure of the C,H,-( CH, OH), cluster at the origin.

Again, the 1: 1 and 1:2 C, H,-H, 0 clusters exhibit simi- lar trends. While it is difficult to observe any van der Waals transitions in the spectra of the C!, H, -H, 0 complex (with a strongest peak - 10% of the AvVdw = 0 transition), the cor- responding structure in C,H,-(H, O), is easily observed, with transitions 5.5, 17, and 22.5 cm-’ above the origin [Fig. 2(c)].

d. The 0 g/6 6 intensity ratio. In several previous studies, we have made use of the forbidden nature of benzene’s

TABLE I. Relative frequencies and intensities of van der Waals’ structure built on the origin transition of C, H,-( CH, OH), .

7248 Garrett, Severance, and Zwier: C,H,-(CH,OH), clusters. I

S, ( B2U ) -So (A Ig ) transition as a vibronic level indicator of the binding sites taken up by the complexing mole- cule (s) .20-22 While theSo-+!?, transition is electric dipole for- bidden in benzene, it can be vibronically induced by vibra- tions of ezg symmetry.28 v~, an ezg in-plane ring elongation mode, is first-order allowed, and the 6; transition is one of the most intense vibronic transitions in the spectrum of free benzene. In benzene-containing clusters, however, intensity can be induced in the SOS, origin of C,H, by the com- plexed molecule (s ) . So 4, origin intensity will be induced if (i) the vibrationally averaged positions of the complexing molecules relative to benzene reduce the effective symmetry of benzene to lower than threefold symmetric and (ii) the interactions of these molecules with benzene’s v cloud are strong enough to distort the r electron density in this asym- metric fashion.

The quantitative measure of this symmetry breaking is the 0:/6: intensity ratio. Table II lists these ratios for the C, H6-( CH, OH) n clusters together with the corresponding measurements in C, H, -( H, 0) n for comparison. As noted earlier, systematic errors introduced by saturation effects re- quire the reported intensity ratios to be treated as upper bounds to the true intensity ratios. Nevertheless, it is the relative magnitude of this ratio in successive cluster sizes which is important for the arguments presented. Further- more, since the observed trends in 0:/6: intensity ratio with changing cluster size will only be heightened in the fully unsaturated limit, the reported intensity ratios can reliably be used as a general structural guide in the way indicated.

The origin transition of the C!, H,-CH, OH complex is not observable in our spectra, with an upper bound on its intensity of 2% of that at 6:. Difficulties in making large quantities of the 1:l complex have inhibited placing better

TABLED. Cluster-inducedspectral featuresoftheC,H,-(CH,OH), and C,H,-(H,O), clusters.

Relative frequency (cm-‘)” Relative intensity

0 82 17 50 28 100 41 97 43 42 46 28 52 41

‘The origin of C,H,(CH,OH), is 77 cm-’ blue shifted from free ben- zene’s origin.

CH, OH

Hz0

Cluster (13)

1:l 1:2 1:3 1:4 1:5 1:l 1:2 1:3 1:4 1:5

Frequency 0: /6:, shift” Intensity

(cm-‘) ratiod ( % )

+44 <2 + 80 7 + 147 14 + 19 5 - 14 1 + 49 <O.l + 75 14 + 98 0.6 + 100 1.0 + 97 1.6

6; splitting (cm-‘)

. . .E 2.0 1.8 -2

. . .c

. ..b 5.4 . . .= . . .= . . .c

*Frequency shift to the blue of the 6; transition of C,H,. b 1.6 cm-’ splitting due to internal rotation ofthe water molecule. See Ref. 1

for details. ‘No 6; splitting is observable at the present resolution. dThe intensity ratios presented must be considered as upper bounds due to

the possible contribution from saturation effects. Despite this, the relative numbers from one cluster size to the next are reliable. See the text for further discussion.

J. Chem. Phys., Vol. 96, No. lo,15 May 1992

Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 5: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

bounds on the origin intensity. Nevertheless, the weakness of the origin is entirely analogous to the forbidden origin of the C, H,-H, 0 complex, indicating that the methanol molecule induces at most minor asymmetry in the benzene r cloud upon complexation. By contrast, the origin transition of the 1:2 cluster is readily observed with an 0:/6; intensity ratio of - 7%. Thus, addition of a second methanol induces signifi- cant asymmetry in the benzene r cloud.

e. The 6 i splitting. The same symmetry reduction which results in induced So-& origin intensity can also break the degeneracy of the 6’ vibrational level in the excited state. This splitting thus serves as a second measure of the degree of asymmetry imposed on benzene by the methanol mole- cule(s). In the case of the 6: splitting, the magnitude of the splitting may have both electronic (i.e., force constant) and kinematic (i.e., mass-related) contributions which are diffi- cult to assess.2’*22 Nevertheless, the qualitative correlation is a good one-a large 6: splitting indicates that the methanol molecule(s) have significant interaction with the benzene ring from vibrationally averaged position(s) off the sixfold axis of benzene. As indicated in Table I, the 1: 1 complex has no observable 6; splitting, while the 1:2 complex has a split- ting of 2.0 cm - ‘, somewhat smaller than its C, H6-( H, O), counterpart.

Taken as a whole, these vibronic level arguments show strong similarities between the C,H,-CH30H and C!, H,-H, 0 complexes. This close correspondence points to similar binding of methanol and water to the benzene ring. The methanol molecule induces only a very weak origin transition and does not observably split the degeneracy of ye, consistent with CH,OH taking up position on or near the sixfold axis. The efficient fragmentation and blue shift of the absorption of the complex suggest hydrogen-bonding inter- actions with the benzene ring, probably primarily through the O-H hydrogen. The determination of the exact orienta- tion of the methanol molecule and its degree of nonrigidity will require rotationally resolved studies. However, as we will see in Sec. III C, the vibronic level constraints on the geometry are confirmed by Monte Carlo simulations based on intermolecular potentials developed by Jorgenson and co-workers.25.26

The C,H,-( CH,OH), cluster also shows striking similarities with C, H6-( H, 0), in the general features of its spectroscopy: (i) a very similar frequency shift; (ii) the analogous presence of significant low-frequency van der Waals structure; (iii) a significant increase in 0:/6; intensi- ty ratio relative to the 1: 1 complex; and (iv) an easily observ- able 6: splitting. We surmise from this that the two metha- nol molecules are interacting with the benzene ring in much the same way as two water molecules are in C, H,-( H, 0) 2. In C, H6-( H, O), , rotational band contour fitting provided center-of-mass positions of the two water molecules.” No- tably, the two water molecules are determined to be on the same side of the benzene ring at a water-water separation similar to that in the free water dimer. It is likely, then, that the C,H,-( CH,OH), complex is also characterized as a methanol dimer29 bound to one side of the ring with primary interaction with the benzene ring via the O-H’s of the two methanols, much as in CbH6-(H20)2,

Garrett, Severance, and Zwier: C,H6(CH,0H), clusters. I 7249

The rotational band contour of the 1:2 cluster at the origin is shown in Fig. 3(a). No observable splittings are present in this band contour due, e.g., to methyl internal rotation. At the present resolution, only the shape of the band contour is resolved, making a structure determination untenable. In spite of this, the shape of the band contour does hold some confirming evidence to support the “same side of the ring” geometry for the two methanol molecules in the 1:2 cluster. Figure 3 (b) presents an asymmetric top band con- tour with methanol dimer attached to one side of the benzene ring at a geometry close to that provided by the calculations of Sec. III D. This contour is in qualitative accord with ex- periment. By contrast, the band contour of Fig. 3 (c), result- ing from the two methanol molecules being placed on the

-2 -1 0 1 Relative Frequency (cm”)

FIG. 3. (a) Experimental rotational bandcontourofthe.S,,-S, origin ofthe C,H,-(CH,OH), cluster. (b) Calculated rotational band contour for the 1:2 cluster using a geometry close to the lowest energy structure predicted by the calculations of Sec. III D in which the two methanols are hydrogen bonded to one another on the same side of the benzene ring. (c) Calculated contour for the 1:2 cluster placing the two methanols near the sixfold axis on opposite sides of the ring. Note that the experimental band contour is much better reproduced by (b) than (c).

J. Chem. Phys., Vol. 96, No. lo,15 May 1992 Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 6: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

7250 Garrett, Severance, and Zwier: C,H&H,OH), clusters. I

sixfold axis on opposite sides of the benzene ring, is not con- sistent with experiment. Clearly, higher resolution spectra are needed before the detailed structure and degree of nonri- gidity of this cluster can be determined.

B. C,H,-(CH,OH),, n=3-5 1. Assignment of spectral features to a given cluster size

Figures 4( a)-4( c) present R2PI scans in the 6: region monitoring the 1:2-1:4 mass channels, respectively. The broad background present in the spectra of the larger clus- ters is especially sensitive to the amount of benzene present in the expansion, indicating that the background arises from fragmentation of higher clusters containing two or more benzene molecules. The scans in the figure were taken at concentrations of C, H, of less than 0.2%. By extrapolation from what we have learned already about the efficient frag-

Unassigned

Higher Clusters

-100 -50 0 50 100 150 200 250 Relative Frequency (cm-‘)

FIG. 4. R2PI scans ofthe 6; region monitoring the (a) 1:2; (b) 1:3; and (c) 1:4 mass channels. The assignments for the major features are given in the figure and discussed in the text. The zero of the relative frequency scale is the 6h transition of free C, H, .

mentation of the 1: 1 and 1:2 clusters following photoioniza- tion, one would expect that the major features in mass chan- nel 1:2-1:4 can be assigned to the 1:3-15 clusters, respectively. This is precisely the assignment that was made in the R2PI spectra 22 of C, H6-( H, 0) n. It is remarkable, however, that no remnant of the absorption features is ob- servable in the next higher mass channel, even though little interference exists in that channel at the resonant wave- lengths of interest. One must postulate either extremely effi- cient fragmentation ( >99%) or the presence of reactive pro- cesses which occur alongside fragmentation to further reduce the ion intensity in the parent channel. Both these mechanisms appear to be in operation in the C, H6-( CH, OH) n clusters.

The identification of product ions resulting from intra- cluster ion-molecule chemistry following photoionization leads to an unambiguous assignment of the neutral cluster responsible for a given absorption feature. Paper II focuses on this intracluster ion chemistry. Here, the assignment af- forded by the ion-molecule chemistry is illustrated using the 1:3 cluster as example. The spectrum of Fig. 5 (a) is recorded

(a)

(b)

150 200 Relative Frequency (cm-‘)

FIG. 5. R2PI scans of the 6; region monitoring the (a) 1:2 and (b) [ (CH,OW, ] + mass channels. The fact that the same resonant features are present in both these channels confirms our assignment of these features to the I:3 cluster. The zero of the relative frequency scale is the 6; transition of free C, H, .

J. Chem. Phys., Vol. 96, No. IO,15 May 1992

Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 7: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

Garrett, Severance, and Zwier: C,H,-(CH,OH), clusters. I 7251

monitoring the 1:2 mass channel. The spectrum is dominat- ed by a strong, sharp absorption feature shifted 147 cm- * blue of the C!, H, 6: transition. However, the mass spectrum recorded with the laser tuned to this absorption feature also shows significant ion intensity at mass 96, corresponding to M,+ (M = methanol) arising from an intracluster dissocia- tive electron transfer reaction. The presence of this ion in resonance with the absorption feature in the I:2 channel [Fig. 5(b) ] ensures that the neutral precursor contains three methanol molecules, strongly supporting the assign- ment of the peak at + 147 cm-’ as being due to the C, H6-( CH, OH) 3 cluster. Similar processes are observed in higher clusters. In each case, the major feature(s) present in the 1:n channel also appear in the M,++ , channel follow- ing dissociative electron transfer and we assign these absorp- tions to the 1: (n + 1) neutral cluster.

There are several other features present in the R2PI scans of the 1:3 and 1:4 mass channels [Figs. 4(b) and 4(c) ] besides the main peaks we have described. The band at + 155 cm - ’ in Fig. 4(b) appears to be due to the 2:3 cluster,

while the red-shifted doublet at - 35 cm - ’ in Fig. 4(b) is not assigned, but is most likely due to a larger cluster as well. Based on the present data, we cannot exclude the possibility that the - 35 cm - ’ peak arises from a second 1:4 conform- er. However, differences in the intracluster ion chemistry of these peaks (Paper II) confirm their carrier as a chemically distinct species from the main peaks at + 19 cm- I.

2. Spectral characteristics of the i:3- 1.5 clusters

As with the 1: 1 and 1:2 clusters, the major point of com- parison of the C, H6-( CH, OH) n clusters is with the corre- sponding C, He-( H, O), spectra which are reproduced in Fig. 6. The important vibronic level characteristics of both types of clusters are summarized in Table II.

a. The I:3 cluster. The overall appearance of the 6; spec- tral region of the C,H,-(CH,OH), [Fig. 4(a)] and C,H6-(H,O), [Fig. 6(a)] clusters is similar. Both spectra are dominated by a single, sharp absorption well blue shifted from that of smaller clusters. In C, H6-( CH, OH) 3, a short, weak progression in a single van der Waals mode of 47 cm - ’ is built on the 6: transition. Closer scrutiny reveals signifi- cant differences between the spectra of the two types of 1:3 clusters, however. First, the frequency shift of C, H6-( CH, OH) 3 ( + 147 cm - ’ ) is 50% larger than that in C,H,-(H,O), ( +98 cm-‘). Second, whereas the C,H,-(H,O), cluster has a very weak origin, the C, H6-( CH, OH) 3 cluster has a strong origin about 14% as intense as the vibronically induced 6: transition. Third, the 6: splitting is readily observed in C,H,-(CH,0H)3, but unresolved in C,H,-(H,O), .

Thus the vibronic level constraints on the structure of the C,H,-( CH, OH) 3 cluster are these:

(i) the cluster exhibits by far the strongest hydrogen bonding to the Pcloud of any of the C, H, -( CH, OH) n clus- ters and significantly greater than that in C,H,-(H,0)3;

(ii) the three methanols bind to benzene in such a way as to strongly break the sixfold symmetry of benzene’s rr cloud, qualitatively different than C, H6-( H, 0) 3 ;

la)

50 100 150 Relative Frequency (cm-‘)

200

FIG. 6. RZPI scans of the 6: region monitoring the (a) 1:2; (b) 1:3; and (c) 1:4 mass channels of C, H6-( H, 0) n clusters for comparison with Fig. 4. The zero of the relative frequency scale is the 6; transition of free C, H, .

(iii) the small amount of van der Waals structure sug- gests a fairly rigid structure for those methanols which are most strongly interacting with benzene.

b. The I:4 cluster. The C,H,-( CH,OH), spectrum [Fig. 4(b) ] shows even less similarity to that of C6 H, -( H, 0 ) 4. The 1:4 methanol cluster absorption shifts dramatically to the red by 128 cm - ’ relative to the 1:3 ab- sorption, while C,H,-( H,O), exhibits virtually no shift from C,H,-(H,O),. The C,H6-(CH,OH), 0:/6: inten- sity ratio is seven times larger than that in C,H,-(H,O), . C, H6-( CH, OH), also exhibits a highly congested spec- trum composed of seven or eight transitions spaced from one another by no more than 2-3 cm - ‘, indicating significant Franck-Condon activity along a very low frequency inter- molecular mode. A close-up scan of this congested spectrum at the origin of the 1:4 cluster is shown in Fig. 7.

The structural consequences of these features are the following:

J. Chem. Phys., Vol. 96, No. lo,15 May 1992 Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 8: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

7252 Garrett, Severance, and Zwier: CBH,(CH,OH), clusters. I

2’0 . I * I * - .

Relative Frequenyy (cm-‘) 6’0

FIG. 7. R2PI scan of the origin region of the 1:4 cluster (monitoring the 1:3 mass channel). Note the closely spaced set of transitions present in this cluster. The zero of the relative frequency scale marks the forbidden origin transition of free C,H,.

(i) the large red shift in the absorption of the 1:4 cluster relative to 1:3 indicates a significant reduction in the strength of rr hydrogen bonding of the methanols to benzene;

(ii) the sixfold symmetry of benzene’s r cloud is broken moderately strongly, comparable to that in the 1:2 cluster, but less than that in the 1:3 cluster;

(iii) electronic excitation of benzene produces a defor- mation of the methanol cluster along one or more very low frequency vibrations.

c. The I:5 Cluster. The C, H, -( CH, OH) 3 R2PI scan of Fig. 4(c) (monitoring the 1:4 mass channel) is once again dominated by a single transition (i.e., Au = 0 Franck-Con- don factors), but is shifted even further red to a position - 13 cm- ’ from C, H, 6;. The C, H6-( H, 0) 5 absorption,

by comparison, remains at + 97 cm- ‘. Despite this differ- ence, the 0:/6: intensity ratio is comparable to that in C,H,-(H,O), and is very weak (- 1%).

The structural constraints placed on the 1:5 cluster are as follows:

(i) the further red shift to a position red of that of free C,H, indicates that the 7r hydrogen-bonding interaction of the methanol molecules has been replaced almost entirely by dispersive interactions;

(ii) intriguingly, the methanol molecules now produce a very small asymmetry in the benzene 7r cloud;

(iii) the lack of van der Waals structure suggests a fairly rigid structure for the methanol molecules which is relative- ly insensitive to benzene electronic excitation.

C. A qualitative comparison with C,H,-(H,O), These striking differences between the C,H,-

(CH,OH), and C,H,-(H,O), spectra with n>3 must re- flect different interactions of the C,H, molecule with methanol than with water clusters. In C,H,-( H,O),, n = 3-5, the nearly unchanged spectral characteristics of the successively larger clusters argue for cluster structures in

which the fourth and fifth water molecules are added to the C,H,-( H,O), cluster in positions far from the benzene ring, thereby disturbing neither the positions nor strengths of the interactions of the other water molecules with the benzene ring.22 Guided by calculations on water clusters which predict cyclic, hydrogen-bonded structures for the trimer, tetramer, and pentamer, our data pointed consistent- ly to C,H,-( H,O), structures which are composed of cy- clic hydrogen-bonded networks of water molecules oriented above the plane of the benzene ring (Fig. 8) .24 The primary interaction of the water cluster with the benzene ring seems to occur via a single water molecule hydrogen bonded to the ring near benzene’s sixfold axis. As we will see, these general features of the C, H6-( H, 0) n clusters are confirmed by cal- culations on the clusters discussed in Sec. III D.

It seems likely that the C, Hs-( CH, OH) n clusters will retain the general features of the corresponding C, H6-( H, 0) n cluster geometries in being composed of hy- drogen-bonded methanol clusters bound to one face of the benzene ring. Structures such as this are consistent with the presence of (CH, OH) ,’ product ions following photoioni- zation (Paper II). This dissociative electron transfer reac- tion would seem unlikely in a neutral cluster structure in which the methanol molecules were distributed on both sides of the benzene ring.

(al (b)

.

P

(d)

FIG. 8. Lowest energy structurescalculated for theC,H,-(H,O), clusters containing from n = 2-5 water molecules using the methods of Sec. III D. The calculated structures are in excellent accord with the experimental data on these clusters from Refs. 1 and 2.

J. Chem. Phys., Vol. 96, No. lo,15 May 1992

Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 9: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

The structure of pure methanol clusters larger than the dimer is not known experimentally. Electric deflection mea- surements”’ on (CH,OH), with n = 3-5 show them all to have negligible dipole moments, consistent with cyclic struc- tures in which all methanol molecules are equivalent or near- ly equivalent. This general picture is confirmed by the se- miempirical calculations (Sec. III D and Ref. 23) which predict minimum-energy structures which are cyclic, hydro- gen-bonded networks in which each methanol acts singly as donor and acceptor with neighbors on either side.

In C,H,-(CH,OH),, clusters, the possibility thus ex- ists that, beginning with the 1:3 cluster, the OH hydrogens may be bound up in hydrogen bonding with other methanol molecules and would thereby not be free to hydrogen bond with the C, H, P cloud. The lack of “dangling” OH groups would mean that the methanol cluster must interact with C,H, via the methyl groups and/or oxygen atoms of the methanol cluster. Thus, while the 1: 1 and 1:2 GH,-(CH,OW,, clusters are quite analogous to C, H,-( H, O), , it is not surprising that the larger clusters show substantial differences.

D. C,H,-(CH,OH), and C&H,-(H,O), structure calculations

The indirect, vibronic level arguments given in the pre- vious section are capable of providing only qualitative tests of the intermolecular potentials governing the interactions of C, H, and methanol molecules. However, the large size of the clusters of interest here require calculations using sim- ple, computationally efficient intermolecular potentials, so that the match between the present levels of experiment and theory is not a bad one. In this section, we present results of Monte Carlo molecular dynamics simulations using a set of intermolecular potentials developed by Jorgenson and co- workers.25*26 Given the very different task for which these potentials have been optimized (namely, bulk solution phase behavior), the potentials are not necessarily an accurate pre- dictor ofevery detail of the positions and orientations of each of the molecules in the cluster, nor will our results require them to be such. The calculations will be used to focus our attention on the broad classes of stable cluster structures which may be present. The experimental data can then be used to select between these possible structure types.

The optimized intermolecular potential functions for both benzene and methanol have been developed previous- 1y.25*26 The functional form of the intermolecular potential incorporates both Lennard-Jones and electrostatic contribu- tions. Both molecules are represented as collections of inter- action sites centered on the atoms. Charges on the atoms and Lennard-Jones parameters were optimized for (i) reason- able geometric and energetic results for the gas phase ben- zene and methanol dimers and (ii) good agreement with the properties of the bulk liquids. The lack of explicit account of dispersive forces and three-body effects is crudely corrected for by a choice of charge distributions which compensate for their absence.26 While more sophisticated intermolecular potentials may need to be developed eventually to account for the detailed structures and intermolecular vibrations of

these clusters, we will see that the present intermolecular potentials possess lowest energy structures very much in keeping with experiment.

Table III collects the results of these calculations for the C,H,-(CH,OH). and C,H,-(H,O), clusters for n = l- 5. In the table, the entries in bold type are those which are consistent with the experimental constraints on their struc- tures. The lowest energy structures calculated for the C, H6-( H, 0) n clusters are presented in Fig. 8 for compari- son with those for the C, H6-( CH, OH) n clusters.

1. The 1:l and 1.2 clusters

Figure 9 presents the lowest energy structures for the 1: 1 and 1:2 clusters. These structures confirm the qualitative interpretation given to the vibronic level characteristics of the clusters. The methanol molecule in the 1:l cluster has a center-of-mass position near the sixfold axis, consistent with the low 0:/6: intensity ratio observed and the lack of an observable 6; splitting. The methanol molecule interacts with benzene’s r cloud via a hydrogen bond involving O-H, in keeping with the blue shift of its absorption and the clus- ter’s efficient fragmentation following photoionization. The similar frequency shift to that in C, H,-H, 0 is also consis- tent with similar positions and strengths of interaction in the C,H,-H,O and C,H,-CH30H. The binding energy pre- dicted for the C,H,-CH,OH cluster (ignoring zero point corrections) is 4.8 kcal/mol (Table III).

The Monte Carlo simulations likewise predict a struc- ture for the 1:2 cluster consistent with the qualitative struc- ture suggested by experiment (Sec. IV A 2). The calculated structure is very analogous to the C, He--( H, O), structure [Figs. 9(b) and 9(c) (cf. Fig. 8) ] in incorporating two methanol molecules on the same side of the benzene ring as a slightly distorted methanol dimer. The structure of Fig. 10 is 3.9 kcal/mol more stable than the most likely alternative in which the second methanol attaches to the other side of the benzene ring. The positions of the methyl groups in these structures are not well tested by the present experiments.

2. The 1~3 cluster Figures IO(a) and IO(b) present the lowest energy

structure calculated for the 1:3 cluster. Interestingly, this structure does not involve a cyclic methanol cluster as we might expect based on the calculated structures for the pure methanol clusters, but rather a hydrogen-bonded chain of methanols. A secondary minimum involving a cyclic metha- nol trimer is shown in Fig. lO( c). This structure is computed to be 1.1 kcal/mol less stable than that in Figs. 10(a) and 10(b).

The I:3 clusters formed in the supersonic expansion are clearly more consistent with the chain structure than the cyclic structure. The chain structure has a dangling O-H free to hydrogen bond to the rr cloud, while the cyclic trimer interacts with the Z- cloud primarily via a methyl group. Thus the former is capable of producing a large So-S, blue shift while the latter is not. Furthermore, the strength of r hydrogen bonding present in the 1:3 chain structure would be predicted to be significantly greater than that for the 1:2

Garrett, Severance, and Zwier: C,H,(CH,OH), clusters. I 7253

J. Chem. Phys., Vol. 96, No. lo,15 May 1992 Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 10: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

7254 Garrett, Severance, and Zwier: C,H,(CH,OH), clusters. I

TABLE III. Calculated structures and binding energies for C,H,-(CHsOH). and C,H,-(H,O), clusters.

Cluster AE(C,H,..*M,)” (M = CH,OH, AE’ AEW, or W,P (Cd-h ..*W,) W = H,O) (kcal/mol) (kcal/mol) (kcal/mol) Structure’

C,H,-M 4.8 . . . 4.8 On the sirfold axis

GH,-M, 13.5 6.9 (7.6) 6.6 The same side, methanol dimer 9.6 0.0 9.6 Opposite sides of the benzene ring

G&.-M, 23.8 159 7.9 (I) Methanol chain 22.7 18.8 (19.0) 3.9 (II) Methanol cycle

‘3%-W 34.7 26.2 8.5 (I) Methanol chain 37.5 32.7 (32.9) 4.8 (II) Methanol cycle

C&-M, 41.1 44.3 2.8 (I) T-shaped, M, cycle as donor 49.6 43.9 (44.3) 5.7 (II) T-shaped, C, H, as donor

C,H,-W 3.8 .” 3.8 On the sixfold axis

Cd-b-W, 11.7 6.5 (6.8) 5.2 The same side, water dimer 7.6 0.0 7.6 Opposite sides of the benzene ring

GH,-Wa 22.1 14.9 7.2 (I) Water chain 22.4 18.2 (18.3) 4.2 (II) Water cycle

G He-W, -3od . . . . . . (I) Water chain” 35.1 31.1(31.2) 4.0 (II) Water cycle

C&.-W, 45.0 41.3 (41.5) 3.7 (I) T-shaped, W, cycle as donor -43-44 -41.5 1.5-2.5 (II) T-shaped, C, H, as donor

“Calculated total binding energy for the cluster. Zero point energy corrections are not included. %alculated binding energy of the methanol or water cluster in the C,H,-W, and C,H,-M, clusters. The difference of columns 1 and 3. The numbers in parentheses are the calculated binding energies of the pure methanol and water clusters after the clusters are allowed to relax to their minimum energy configuration.

Calculated binding energy of C,H, to the M, or W, cluster in the specified configuration. In determining these energies, the M, and W, clusters are held llxed in the structures, while C,H, is pulled away from the cluster.

d Not an energy minimum. eStructures in bold face type are those which are most consistent with the present experimental data.

cluster by virtue of the third methanol molecule further po- larizing the hydrogens on the other methanols which inter- act directly with the benzene rr cloud. This is also in accord with the additional blue shift of the transitions of the 1:3 cluster from those of the 1:2 cluster. Finally, the chain struc- ture is also consistent with the strong breaking of the sixfold symmetry of benzene’s r cloud observed for the 1:3 cluster. The strong polarization of the two methanol molecules by hydrogen bonding to the third methanol should distort the benzene QT cloud along a single direction, inducing a strong origin transition.

The comparison with C, H, -( H, 0) 3 is striking. As Ta- ble III indicates, the C, H,-( H, 0) 3 cluster is also calculat- ed to have two nearly equally stable conformers composed of either a cyclic or a chain-like hydrogen-bonded water trimer attached to benzene’s 7r cloud. The experimental results sug- gest that, of these two structures, the cyclic (H, 0) 3 struc- ture [Fig. 8(b) ] is the one preferentially formed by C,H,-( H,O), clusters in the expansion, even though the chain structure is more consistent with the experimental datainC,H,-(CH,OH),.

3. The i-4 cluster

The lowest energy structure calculated for the I:4 clus- ter is shown in Figs. 11 (a) and 11 (b) (0, = 37.5 kcal/

mol), while a secondary minimum structure (D, = 34.8 kcal/mol) is shown in Fig. 11 (c) . Interestingly, by contrast to the 1:3 cluster, the calculations predict that in the 1:4 cluster, the methanols prefer a cyclic structure over a hydro- gen-bonded chain structure. The present experiments sup- port this predicted preference for the cyclic structure. The most striking feature of the 1:4 cluster is the dramatic shift to the red of its absorption relative to the 1:3 cluster. The inter- pretation we have given this shift is that it represents a large reduction in the strength of the hydrogen-bonding interac- tion with the 7r cloud. This is precisely what one would ex- pect for the cyclic C, H, -( CH, OH) 4 cluster of Figs. 11 (a) and 11 (b) in which the primary interaction with the IT cloud is now via the methyl group near the sixfold axis. The O-H hydrogen bonding in the chain structure of Fig. 11 (c) would likely produce a strong blue shift comparable to that in the 1:3 cluster.

The modest 0:/6: intensity ratio (5% ) and 6: splitting (2 cm - ‘) of the cluster is also consistent with the cyclic structure in that the calculated strength of interaction of the cyclic methanol cluster with the benzene P cloud is reduced significantly. Table III lists the binding energy of the ben- zene molecule to the methanol cluster [ AE(C, H, * * *M, ) ] in each of the cluster structures of interest. In the 1:3 cluster calculation, the chain type methanol trimer, which is hydro- gen bonded to the P cloud, has a binding energy of some 7.9 kcal/mol to benzene. By contrast, in the 1:4 cluster, the cy-

J. Chem. Phys., Vol. 96, No. lo,15 May 1992

Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 11: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

Garrett, Severance, and Zwier: C,H,(CH,OH), clusters. I 7255

(81 ,a (a)

(b)

FIG. 9. Lowest energy structurescalculated for (a) 1:l and (b) and (c) I:2 C,H,-(CH,OH), clusters. Note the similarity of these structures to the experimentally determined structures for C,H,-H, 0 and C,H,-(H,O), clusters.

clic methanol tetramer, which lacks this hydrogen bonding, is calculated to bind to benzene by only 4.8 kcal/mol. The possibility is thus raised that even though the location of the methanols is still highly asymmetric in the 1:4 cluster, the induced origin intensity may be weakened (relative to the 1:3 cluster) by the weaker interaction of the benzene P cloud with the cyclic methanol tetramer. The drop in C,H,**- (CH, OH) n binding energy is more than compen- sated for in the cyclic 1:4 cluster by the extra methanol- methanol hydrogen bond so produced [ AE(M, > 1. That just the reverse is true in the 1:3 cluster is a testimony to the subtle balance of intermolecular interactions which is at play in determining the global m inimum energy structure in these clusters.

Thus, based on the present data, the C,H,-( CH, OH)., [Figs. 11(a) and 11(b)] and C,H,-(H,O), [Fig. 8(c)] clusters formed in the expansion are both consistent with structures composed of cyclic, hydrogen-bonded solvent clusters bound to benzene. The calculated structures point to

lb)

FIG. 10. (a) and (b) Lowest energy structure calculated for the 1:3 cluster, involving a chain-type hydrogen-bonded methanol trimer. (c) A secondary minimum structure for the 1:3 cluster involving a cyclic hydrogen bonded methanol cluster. This structure is calculated to be 1.1 kcal/mol less stable than that in (a) and (b).

the Z- hydrogen-bonding interaction present in C,H,- (H,O), orienting the water cluster off away from the ben- zene plane. However, the calculations suggest that the much less directional bonding of the methyl group in C, H, -(CH, OH), allows the methanol cluster to lie down on the benzene ring and even to begin to wrap around the ring.

4. The 1:5cluster

Figure 12 presents the lowest energy structure calculat- ed for the 15 cluster. It is not surprising that a cyclic metha- nol cluster, is predicted to be the most stable configuration once again. What is surprising at first is the orientation of the cyclic methanol pentamer with respect to the benzene ring. The methanol pentamer takes up a T-shaped configuration in which the primary interaction is with the hydrogen(s) on the benzene ring rather than with the benzene r cloud. The structure is similar in type to that of the benzene dimer in which the cyclic methanol cluster plays the role of one of the

J. Chem. Phys., Vol. 96, No. IO, 15 May 1992 Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 12: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

7256 Garrett, Severance, and Zwier: C,H&CH,OH), clusters. I

FIG. 11. (a) and (b) Lowest energy structure calculated for the 1:4 cluster involving a cyclic hydrogen-bonded methanol cluster. (c) A secondary minimum structure for the 1:4 cluster involving a chain-type hydrogen- bonded methanol tetramer. This structure is calculated to be 2.8 kcal/mol less stable than that in (a) and (b) .

benzene molecules. The beginnings of this tipping of the methanol cluster onto the side of the benzene ring are al- ready evident in the calculated structure for the 1:4 cluster [Figs. 11 (a) and 11 (b) 1, but here the reorientation is com- plete.

As with the smaller clusters, the intermolecular poten- tial used in the Monte Carlo calculations reproduces the ma- jor characteristics of the 15 clusters present in the molecular beam. For instance, in the calculated structure, the interac- tion of the methanol cycle with the benzene ring is complete- ly lacking in 7r hydrogen-bonding character, supporting the further shift of its absorption to a position red of that of free benzene ( - 14 cm - ’ ). Second, the very small 0:/6; inten- sity ratio and negligible 6: splitting of the cluster requires that the methanol molecules induce only small asymmetry in the benzene rr cloud. If the 15 clusters in the expansion take on astructure similar to that of Fig. 12, they m ight be expect- ed to induce little asymmetry as a simple consequence of the very weak interaction of the methanol pentamer with the benzene P cloud. The asymmetric orientation predicted for

(a) *1-

+v c ‘jg, c

FIG. 12. Lowest energy structure calculated for the I:5 cluster.

the methanol pentamer relative to benzene’s sixfold axis may not be enough to induce the origin transition since it is the rr cloud which is the sensitive probe of benzene’s surroundings in the ~-+r* electronic transition. The calculated structure is even consistent with the sharp reduction observed in the number and strength of the van der Waals transitions in the spectrum (by comparison to 1:4) by the weak interaction of the methanol cluster with the v cloud involved in the elec- tronic excitation.

The comparison of the calculated structures of C!, H,- (CH, OH), and C,H,-( H,O) 5 is once again revealing. Qualitatively speaking, both calculated lowest-energy struc- tures are T-shaped structures composed of two rings. How- ever, in C, H, -( H, 0) 5, the dangling hydrogens present ori- ent the water cluster as a proton donor into the benzene R cloud. The cyclic methanol cluster, which possesses no such dangling O-H groups, shows a calculated preference toward orienting itself as proton acceptor for the hydrogen of ben- zene. In this case, the electron-rich region presented by the five oxygen atoms of the methanol cluster serves as the bind- ing site for one of benzene’s hydrogen atoms. Additional experimental constraints on the structures of these clusters will need to be devised to test these general conclusions further.

IV. CONCLUSIONS In this paper, we have used a series of vibronic level

arguments to constrain the geometries of several small

J. Chem. Phys., Vol. 96, No. lo,15 May 1992

Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 13: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

C,H,-( CH,OH), clusters and to provide points of com- parison with earlier work on C, H6-( H, O), clusters.22 The vibronic level characteristics of the C,H,-CH30H and C,H,-( CH,OH), clusters are very similar to those of C, H,-H2 0 and C, Hb-( H, 0) *. These similarities point to- ward the first methanol molecule, like the first water mole- cule, adding to benzene on or near the sixfold axis of benzene in a hydrogen-bonding configuration. The spectra of C,H,-(CH,OH), suggest that the second methanol binds to the first methanol on the same side of the benzene ring. Our proposed structure is composed of a methanol dimer bound to benzene with one methanol near the benzene six- fold axis and one arranged well off that axis. The primary interaction of the methanol dimer with the ring is still via the O-H hydrogens on the methanol molecules, as they are with the hydrogens in the water molecules of C,H,-(H,O),.

The spectral characteristics of the C, H,-( CH, OH) n and C, H,-( H, O), clusters begin to diverge in the 1:3 clus- ter and become very dissimilar for n>4. Intermolecular po- tentials optimized for solution-phase properties are used to focus attention on a few of the lowest energy structures for each cluster size. The experimental data are then used to select between these structures. Whether fortuitously or not, in each case, the lowest energy calculated structure for each cluster size is entirely consistent with our experimental data. In particular, our past work on C, Hb-( H, 0) n clusters with n = 3-5 are most consistent with calculated structures com- posed of cyclic hydrogen-bonded water clusters oriented nearly perpendicular to the benzene ring.22 This maximizes the strength of the hydrogen bonding of a dangling O-H on one of the waters with benzene’s a- cloud. By contrast, the analogous cyclic hydrogen-bonded methanol clusters pro- vide no dangling O-H’s to interact favorably with the ben- zene YT cloud. In the 1:3 cluster, the Monte Carlo simulations predict a lowest energy structure in which the cyclic metha- nol cluster is broken to a hydrogen-bonded chain lying down on the ring in such a way as to allow favorable r hydrogen bonding by one of the methanols. The spectral characteris- tics of the 1:3 cluster support this hydrogen-bonded chain structure. By contrast, the calculated structures for the 1:4 and 15 clusters are composed of cyclic methanol clusters. This occurs at the expense of an almost complete loss of the hydrogen-bonding interaction with benzene, as surmised from the experimental data. In these larger clusters, the ex- tra binding to benzene accorded a rr hydrogen-bonded chain-type structure is more than compensated for by the extra methanol-methanol hydrogen bond formed in the cy- clic methanol cluster.

This restructuring in which the methanol molecules take up a less preferred benzene-methanol orientation in or- der to facilitate hydrogen bonding with neighboring metha- nol molecules may contribute to the large solubility of ben- zene in methanol in bulk so1ution.3’ Benzene, as a prototypical nonpolar aromatic hydrocarbon, is often cate- gorized as having two dissimilar regions for interaction with polar solvents-a hydrophilic out-of-plane region generated by its electron-rich r cloud and a hydrophobic in-plane re- gion.32 It is significant that the lowest energy 1:4 and 1:5 structures predicted by calculations, and supported by the

present experiments, involve substantial in-plane interac- tions with C, H, . In fact, the methanol pentamer in the cal- culated 1:5 structure interacts with the benzene ring almost solely via the in-plane hydrogen(s) of benzene. It is an in- triguing thought that methanol’s inability to simultaneously hydrogen bond both to neighboring methanols and to the benzene rr cloud (as H, 0 can) may contribute to the ability of methanol to solvate C,H, in bulk solution. Clearly, further experimental and theoretical work is needed to more fully characterize the structures and properties of these in- teresting clusters, to test and modify the intermolecular po- tentials used to predict their behavior, and to clarify the rela- tionship of these clusters’ properties to bulk solution phase behavior.

ACKNOWLEDGMENTS

Acknowledgment is made to the Donors of the Petrole- um Research Fund, administered by the American Chemi- cal Society, and to the National Science Foundation (Grant No. CHE-9108376) for their support ofthis research. T.S.Z. also thanks the Alfred P. Sloan Foundation for a Research Fellowship.

‘J. C. Alfano, S. J. Martinez, and D. H. Levy, J. Chem. Phys. 91, 7302 ( 1989); C. A. Haynam, D. V. Brumbaugh, and D. H. Levy, ibid. 80,2256 (1984).

*T. Weber, A. VonBarger, E. Riedle, and H. J. Neusser, J. Chem. Phys. 92, 90 (1990); Th. Weber and H. J. Neusser, ibid. 94,706l (1991).

‘E. J. Bieske, M. W. Rainbird, I. M. Atkinson, and A. E. W. Knight, J. Chem. Phys. 91,752 ( 1989); E. J. Bieske, M. W. Rainbird, and A. E. W. Knight, ibid. 94,7019 (1991); E. J. Bieske, A. S. Uichanco, M. W. Rain- bird, and A. E. W. Knight, ibid. 94, 7029 ( 1991).

4P. M. Weber, J. T. Buontempo, F. Novak, and S. A. Rice, J. Chem. Phys. 88,6082 (1988).

‘D. 0. DeHaan, A. L. Holton, and T. S. Zwier, J. Chem. Phys. Xl,3952 (1989).

6T. Troxler and S. Leutwyler, J. Chem. Phys. 95, 4010 (1991); S. Leutwyler and J. Bosiger, Faraday Discuss. Chem. Sot. 86,225 ( 1988).

‘M. Y. Hahn and R. L. Whetten, Phys. Rev. Lett. 61, 1190 (1988). *K. W. Butz, D. L. Catlett, G. E. Ewing, D. Krajnovich, and C. S. Par-

menter, J. Phys. Chem. 90,3533 (1986). 9 J. Wanna, J. A. Menapace, and E. R. Bernstein, J. Chem. Phys. 85,1795

( 1986); J. A. Menapace and E. R. Bernstein, J. Phys. Chem. 91, 2533 (1987); J. Chem. Phys. 86,6707 (1987); P. J. Breen, E. R. Bernstein, J. I. Seeman, and H. V. Secor, J. Phys. Chem. 93,673l (1989); H. S. Im, V. H. Grassian, and E. R. Bernstein, ibid. 94,222 ( 1990); S. K. Kim and E. R. Bernstein, ibid. 94, 3531 (1990).

“0. Cheshnovsky and S. Leutwyler, J. Chem. Phys. 88,4127 (1988); T. Droz, R. Knochenmuss, and S. Leutwyler, ibid. 93, 4520 ( 1990); J. Bo- siger, R. Bombach, and S. Leutwyler, ibid. 94, 5098 ( 199 1) .

” C. L. Morter, A. Koskelo, Y. R. Wu, and D. H. Levy, J. Chem. Phys. 89, 1867 (1988); C. A. Haynam, C. Morter, L. Young, and D. H. Levy, J. Phys. Chem. 91,2519 (1987); 91,2526 (1987); L. A. Pataneauand D. H. Levy, ibid. 92, 6554 (1988).

‘*J. A. Syage and J. E. Wessel, J. Chem. Phys. 89, 5962 ( 1988). I3 R. J. Lipert and S. D. Colson, J. Chem. Phys. 89,4579 (1988). 14J J Breen, K. Kilgore, W.-B. Tzeng, S. Wei, R. G. Keesee, and A. W. . .

Castleman, J. Chem. Phys. 90, 11 ( 1989); R. J. Stanley and A. W. Castle- man, ibid. 94, 7744 ( 199 1) .

“L L Connell, S. M. Ohline, P. W. Joireman, T. C. Corcoran, and P. M. . . Felker, J. Chem. Phys. 94,466s (1991).

16E. Shalev, N. Ben-Horin, and J. Jortner, J. Chem. Phys. 94,7757 ( 1991); E. Shalev, N. Ben-Hot-in, U. Even, and J. Jortner, ibid. 95,3147 ( 1991).

“B. Brutschy, J. Phys. Chem. 94.8637 (1990); B. Brutschy, J. Eggert, C. Janes, and H. Baumgartel, ibid. 95, 5041 ( 1991).

‘*M. Ito J Mol. Struct. 177, 173 ( 1988), and references therein; T. Ebata, , . M. Furukawa, T. Suzuki, and M. Ito, J. Opt. Sot. Am. B 7, 1890 ( 1990).

Garrett, Severance, and Zwier: C,H6(CH30H), clusters. I 7257

J. Chem. Phys., Vol. 96, No. IO,15 May 1992 Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

Page 14: Multiphoton ionization studies of C6H6--(CH3OH)n clusters ... · photophysics of various aromatic-X,, clusters. Many of these studies have focused on rare gas atoms as solvents’

7258 Garrett, Severance, and Zwier: C,H&CH,OH), clusters. I

l9 K. Fuke and K. Kaya, Chem. Phys. Lett. 94,97 (1983). *‘A. J. Gotch and T. S. Zwier, J. Chem. Phys. 93,6977 ( 1990). *‘A. J. Gotch, A. W. Garrett, and T. S. Zwier, J. Phys. Chem. 95, 9699

(1991). “A. J. Gotch, A. W. Garrett, D. L. Severance, and T. S. Zwier, Chem.

Phys. Lett. 178,121 (1991); A. J. Gotch and T. S. Zwier, J. Chem. Phys. %,3388 (1992); A. W. Garrett and T. S. Zwier, ibid. 96, 3402 (1992).

“G. Brink and L. Glasser, J. Comput. Chem. 3,219 (1982); T. P. Martin, T. Bergmann, and B. Wassermann, in Lcrrge Finite Syszems, edited by J. Jortner (Reidel, Dordrecht, 1987), p. 121.

*‘B. J. Mhin, H. S. Kim, C. W. Yooni and K. S. Kim, Chem. Phys. Lett. 176.41 11991): C. E. Dvhkstra. J. Chem. Phvs. 91.6472 (1989); K. S. - -, . , , Kim, M. Dupuis, G. C. Lie, andE. Clementi, Chem: Phys. Lett. lk, 45 1 (1986); D. Belfordand E. S. Campbell, J. Chem. Phys. 86,7013 (1987).

‘s W. L. Jorgenson, J. Phys. Chem. 90, 1276 (1986). 26W. L. Jorgenson and D. L. Severance, J. Am. Chem. Sot. 112, 4768

( 1990). The best balance of parameters typically leads to gas phase dipole moments which are too large by 25%, dimer bond lengths which are slightly too short (-0.2 A), and dimer interaction energies which are

slightly stronger than the best ab initio calculations would predict. 27J. R. Gord, A. W. Garrett, R. E. Bandy, and T. S. Zwier, Chem. Phys.

Lett.171,443 (1990). 28A. E. W. Knight, C. S. Parmenter, and M. W. Schuyler, J. Am. Chem.

sot. 97,1993 (1975). 29 J. P. LaCosse and J. M. Lisy, J. Phys. Chem. 94,4398 ( 1990); A. Bizzarri,

S. Stolte, J. Reuss, J. G. C. M. van Duijneveldt-Vande Rijt, and F. B. van Duijneveldt, Chem. Phys. 143,423 (1990); U. Buck, X. Gu, C. Lauen- stein, and A. Rudolph, J. Phys. Chem. 92, 5561 ( 1988); F. Huisken and M. Stemmler, Chem. Phys. Lett. 144,391 ( 1988).

MB. D. Kay and A. W. Castleman, J. Phys. Chem. 89,4867 ( 1985); J. A. Gdutola, R. Viswanathan, and T. R. Dyke, J. Am. Chem. Sot. lo&4787 (1979).

“CRC Handbook of Chembtry and Physics, 65th ed., edited by R. C. Weast, (CRC, Boca Raton, FL. 1984). D. C-128.

32P. Linse; J. Am. Chem. S&112; 1744 (1990); G. Karlstrom, P. Linse, A. Wallqvist and B. Jonsson, ibid. 105, 3777 (1983); B. V. Cheney and M. W. Schulz, J. Phys. Chem. 946268 (1990).

J. Chem. Phys., Vol. 96, No. lo,15 May 1992

Downloaded 21 Aug 2003 to 128.210.142.204. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp