SPECTROSCOPY OF METAL ION COMPLEXES: Gas-Phase Models for Solvation

Similar documents
Infrared spectroscopy to probe structure and dynamics in metal ion molecule complexes

Laser Dissociation of Protonated PAHs

Wolfgang Demtroder. Molecular Physics. Theoretical Principles and Experimental Methods WILEY- VCH. WILEY-VCH Verlag GmbH & Co.

Headspace Raman Spectroscopy

CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions

PHOTOELECTRON SPECTROSCOPY OF RADICALS

ATOMIC STRUCTURE, ELECTRONS, AND PERIODICITY

5.80 Small-Molecule Spectroscopy and Dynamics

Vibrational Autoionization in Polyatomic molecules

Atoms, Molecules and Solids (selected topics)

ATOMIC STRUCTURE, ELECTRONS, AND PERIODICITY

V( x) = V( 0) + dv. V( x) = 1 2

( ) x10 8 m. The energy in a mole of 400 nm photons is calculated by: ' & sec( ) ( & % ) 6.022x10 23 photons' E = h! = hc & 6.

NPTEL/IITM. Molecular Spectroscopy Lectures 1 & 2. Prof.K. Mangala Sunder Page 1 of 15. Topics. Part I : Introductory concepts Topics

CHEM1901/ J-5 June 2013

Collisionally Excited Spectral Lines (Cont d) Diffuse Universe -- C. L. Martin

infrared photodissociation spectroscopy.

Atom-molecule molecule collisions in spin-polarized polarized alkalis: potential energy surfaces and quantum dynamics

Lecture 14 Organic Chemistry 1

Spectroscopy Of Hydrogen-bonded Formanilide Clusters In A Supersonic Jet: Solvation Of A Model Trans Amide

Infrared photodissociation spectra of CH 3 Ar n complexes n 1 8

Application of IR Raman Spectroscopy

Photoelectron Spectroscopy using High Order Harmonic Generation

Chapter IV: Electronic Spectroscopy of diatomic molecules

IR Spectrography - Absorption. Raman Spectrography - Scattering. n 0 n M - Raman n 0 - Rayleigh

Problems with the Wave Theory of Light (Photoelectric Effect)

Spectroscopy in Inorganic Chemistry. Vibration and Rotation Spectroscopy

Solvation effects on the molecular 3s Rydberg state: AZAB/CYCLO octanes clustered with argon

LECTURE NOTES. Ay/Ge 132 ATOMIC AND MOLECULAR PROCESSES IN ASTRONOMY AND PLANETARY SCIENCE. Geoffrey A. Blake. Fall term 2016 Caltech

What dictates the rate of radiative or nonradiative excited state decay?

8. Which of the following could be an isotope of chlorine? (A) 37 Cl 17 (B) 17 Cl 17 (C) 37 Cl 17 (D) 17 Cl 37.5 (E) 17 Cl 37

Secondary Ion Mass Spectrometry (SIMS)

DEVELOPMENT OF THE PERIODIC TABLE

The broad topic of physical metallurgy provides a basis that links the structure of materials with their properties, focusing primarily on metals.

Chapter 8 Covalent Boding

Rethinking Hybridization

Photoelectron spectroscopy via the 1 1 u state of diacetylene

Lecture Presentation. Chapter 8. Periodic Properties of the Element. Sherril Soman Grand Valley State University Pearson Education, Inc.

Spectroscopic Investigation of Polycyclic Aromatic Hydrocarbons Trapped in Liquid Helium Clusters

s or Hz J atom J mol or -274 kj mol CHAPTER 4. Practice Exercises ΔE atom = ΔE mol =

VIBRATION-ROTATION SPECTRUM OF CO

Chemistry 543--Final Exam--Keiderling May 5, pm SES

CHM Physical Chemistry II Chapter 12 - Supplementary Material. 1. Einstein A and B coefficients

Chapter 10: Modern Atomic Theory and the Periodic Table. How does atomic structure relate to the periodic table? 10.1 Electromagnetic Radiation

Ultraviolet-Visible and Infrared Spectrophotometry

IB Chemistry. Chapter 4.1

Abstract... I. Acknowledgements... III. Table of Content... V. List of Tables... VIII. List of Figures... IX

Solutions and Non-Covalent Binding Forces

Fundamentals of Spectroscopy for Optical Remote Sensing. Course Outline 2009

Photon Interaction. Spectroscopy

MULTIPLE CHOICE. Choose the one alternative that best completes the statement or answers the question.

Theoretical Chemistry - Level II - Practical Class Molecular Orbitals in Diatomics

Electron Configuration and Chemical Periodicity

Molecular spectroscopy Multispectral imaging (FAFF 020, FYST29) fall 2017

Saturation Absorption Spectroscopy of Rubidium Atom

Elements react to attain stable (doublet or octet) electronic configurations of the noble gases.

8.1 Early Periodic Tables CHAPTER 8. Modern Periodic Table. Mendeleev s 1871 Table

Chem 442 Review of Spectroscopy

Spectroscopy and chemical dynamics of group II metal ion-formaldehyde complexes

PAPER No. : 8 (PHYSICAL SPECTROSCOPY) MODULE NO. : 23 (NORMAL MODES AND IRREDUCIBLE REPRESENTATIONS FOR POLYATOMIC MOLECULES)

Lecture 9 Electronic Spectroscopy

CHEM J-5 June 2014

Rb, which had been compressed to a density of 1013

Chemistry 101 Chapter 9 CHEMICAL BONDING. Chemical bonds are strong attractive force that exists between the atoms of a substance

Laser Induced Fluorescence of Iodine

high temp ( K) Chapter 20: Atomic Spectroscopy

5.80 Small-Molecule Spectroscopy and Dynamics

Chapter 8. Periodic Properties of the Element

Ch. 7- Periodic Properties of the Elements

ATOMIC STRUCTURE. Content

Microsolvation of HN 2 + in Argon: Infrared Spectra and ab Initio Calculations of Ar n -HN 2 + (n ) 1-13)

"What Do I Remember From Introductory Chemistry?" - A Problem Set

Atomic Spectroscopy II

CHEM6416 Theory of Molecular Spectroscopy 2013Jan Spectroscopy frequency dependence of the interaction of light with matter

Quantum chemistry and vibrational spectra

A study of the predissociation of NaK molecules in the 6 1 state by optical optical double resonance spectroscopy

Experiment 6: Vibronic Absorption Spectrum of Molecular Iodine

Quantum Chemistry. NC State University. Lecture 5. The electronic structure of molecules Absorption spectroscopy Fluorescence spectroscopy

GCE Chemistry Eduqas AS Component 1

Chapter 02 The Chemical Basis of Life I: Atoms, Molecules, and Water

CDO AP Chemistry Unit 5

Ch. 8 Introduction to Optical Atomic Spectroscopy

MS/MS .LQGVRI0606([SHULPHQWV

Atomic structure & interatomic bonding. Chapter two

Electronic Structure and Bonding Review

Chapter 7. Ionic & Covalent Bonds

Electronic Spectra of Complexes

Physical Chemistry Laboratory II (CHEM 337) EXPT 9 3: Vibronic Spectrum of Iodine (I2)

two slits and 5 slits

Section 11: Electron Configuration and Periodic Trends

Chapter 9: Electrons and the Periodic Table

Welcome to Organic Chemistry II

Assumed knowledge. Chemistry 2. Learning outcomes. Electronic spectroscopy of polyatomic molecules. Franck-Condon Principle (reprise)

INFRARED SPECTROSCOPY OF CATION-WATER COMPLEXES BISWAJIT BANDYOPADHYAY. (Under the Direction of Michael A. Duncan) ABSTRACT

UV-vis (Electronic) Spectra Ch.13 Atkins, Ch.19 Engel

A. 24 B. 27 C. 30 D. 32 E. 33. A. It is impossible to tell from the information given. B. 294 mm C. 122 mm D. 10 mm E. 60 mm A. 1 H B. C. D. 19 F " E.

Bonding forces and energies Primary interatomic bonds Secondary bonding Molecules

25 Mn Ni Co Rh Fe Ru Os Uns (262) Une (266) 195.

Electron Configuration in Ionic Bonding Ionic Bonds Bonding in Metals

( ) electron gives S = 1/2 and L = l 1

Transcription:

Annu. Rev. Phys. Chem. 1997. 48:69 93 Copyright c 1997 by Annual Reviews Inc. All rights reserved SPECTROSCOPY OF METAL ION COMPLEXES: Gas-Phase Models for Solvation Michael A. Duncan Department of Chemistry, University of Georgia, Athens, Georgia 30602-2556; e-mail: maduncan@uga.cc.uga.edu KEY WORDS: metal ions, metal complexes, electronic spectroscopy, clusters ABSTRACT Weakly bound metal ion complexes are produced in molecular beams and studied with mass-selected laser photodissociation spectroscopy. The metal ions Mg + and Ca + are the focus of these studies because they have a single valence electron and strong atomic resonance lines in convenient wavelength regions. Weakly bound complexes of these ions with rare-gas atoms and small molecules are prepared with laser vaporization in a pulsed nozzle cluster source. The vibrationally and rotationally resolved electronic spectra obtained for these complexes help to determine the complexes structures and bonding energetics. Observations from these studies have provided many new insights into the fundamental interactions in electrostatic bonding. Introduction Metal ions and their electrostatic interactions are ubiquitous in chemistry. Solvated metal ions in various oxidation states are responsible for much of condensed-phase inorganic chemistry. Many enzymes have metal ions at their active sites, and electrostatic interactions influence binding at these sites. Specific electrostatic interactions have been invoked to explain selective metal ion transport through cell membranes (1). Ionic metal interactions influence the binding on many catalytic surfaces (e.g. metal oxides) (2). Gaseous metal ions (e.g. Mg +,Ca +,Al + ) are present in surprising abundance in the earth s atmosphere, deposited there by meteor ablation (3). Similarly, metal ions (e.g. Mg + ) are observed in planetary atmospheres (4 6), e.g. in the impact of the comet 0066-426X/97/1001-0069$08.00 69

70 DUNCAN Shoemaker-Levy 9 at Jupiter (6). Metal-containing plasmas are vital ingredients in thin-film deposition for microelectronic materials (7). In these and other varied chemical environments, the electrostatic interactions of metal ions determine to a large degree the outcome of significant chemistry. Unfortunately, the ability of theory to describe these phenomena is limited. It is therefore important to measure the details of these interactions. This can be accomplished through the study of weakly bound complexes containing metal ions with small molecules, which can be produced under controlled conditions in the gas phase. Such complexes provide simple model systems, tractable for experiment and theory, that can probe bonding interactions in metal-solvent or metal-ligand systems. This article describes our recent progress in the spectroscopy of sizeselected magnesium and calcium ion complexes, which has provided many new insights into the structure and bonding in electrostatic systems. There have been previous studies of metal complexes with equilibrium mass spectrometry (8 13), collision-induced dissociation (14a d), ligand switching or radiative association (15a,b), and fixed-wavelength laser photodissociation (16a 19b). However, the spectroscopic data on these systems is limited. Several research groups have begun to apply tunable laser photodissociation spectroscopy to the study of metal complexes over the past few years [e.g. Farrar (20a f), Brucat (21a h, 22a,b), Stwalley & Kleiber (23a,b), Fuke (24a e), Miller (25)]. However, many of these studies use low-resolution measurements to probe the size-dependent chemistry and solvation. There have been promising new rotationally resolved spectroscopic studies by Brucat and coworkers (21a h, 22a,b) and Miller and coworkers (25). However, there are only a handful of polyatomic complexes for which structures have been determined. Theoretical investigations are also beginning to focus on metal complexes (26 41c). Bauschlicher and coworkers (26 34) have performed extensive calculations on magnesium, aluminum, and calcium ions in complexes with rare gases and small molecules. Metals such as these have few enough electrons that high-level calculations are possible for both ground and excited electronic states. These investigations have examined structures, energetics, and electronic spectra for mono-ligand complexes as well as multiple-ligand species. The interaction of theory and experiment has been extremely productive for these complexes. Many mass spectroscopy and theoretical studies of metal ion-molecular complexes focus on the singly charged alkali cations or the doubly charged alkaline earth cations because these are the familiar charge states in solution. However, both of these kinds of metal ions are closed-shell species with no electronic spectra at low energy. Spectroscopic studies are facilitated for singly charged alkaline earth metal ions because they are open-shelled with intense atomic resonances ( 2 P 2 S) at low energy (42). In weakly bound complexes of these ions,

METAL ION COMPLEXES 71 molecular transitions from the ground to the low-lying electronic states most often correspond to excitations from the ns to the np atomic orbitals centered on the metal. These transitions are analogous to the neutral sodium D line and possess similar spectral intensity. Studies on such systems in other laboratories include survey photodissociation spectra of Sr + -(NH 3 ) N and Sr + -(H 2 O) N (20a f); vibrationally resolved photodissociation spectra of Mg + -D 2 (23a,b), Ca + -CH 4 (23a,b), and Ca + -Ar (22a,b); and rotationally resolved spectra of Ba + -Ar (25). Our work, as described here, has examined several magnesium and calcium complexes with rare gases (RGs) and with small molecules (43a 51). These include Mg + -RG (45a,b, 46a,b), Ca + -RG (48), Mg + -CO 2 (43a c), Ca + -CO 2 (50), Mg + -N 2 (51), Mg + -H 2 O (44a,b), Ca + -H 2 O (49), and Ca + -N 2 (52). Production and Detection of Metal Cation Complexes Metal cation complexes with rare-gas atoms or small molecules are prepared using a rotating rod laser vaporization cluster source (46a,b). A solid sample of the appropriate metal is mounted in a homemade holder attached to a pulsed molecular beam valve (Newport Corporation BV-100, with homemade modifications) (53). An excimer laser operating at the XeCl wavelength (308 nm) is focused with a lens onto the surface of the sample, and the timing of this laser pulse is adjusted with respect to the expansion-gas pulse to optimize ion production, as monitored with the downstream mass spectrometer. The holder is designed to allow rapid expansion of the pulsed-beam gas at exactly the vaporization point. This so-called cutaway design provides a vigorous supersonic expansion leading to efficient metal atom complex formation while limiting the formation of larger metal-metal clusters (46a,b). The complexes produced from this source pass through a skimmer into a differentially pumped mass spectrometer chamber where detection and spectroscopy measurements are made. The mass spectrometer is a specially designed reflectron time-of-flight instrument, which makes it possible to measure the distribution of complexes produced by the source and to size select certain complexes for photodissociation spectroscopy (54). The configuration of the reflectron instrument that allows mass selection and photodissociation has been described in detail previously (54). We use the flight time along an initial drift tube section together with a pulsed deflection plate to transmit only selected ion masses to the reflectron grid region. A tunable dye laser intersects the selected ion trajectory in the turning region of the reflectron field, and photodissociation may take place if the laser is tuned to a vibronic resonance of the complex. In many complexes, absorption to a bound excited state followed by absorption of a second photon leads to efficient resonance-enhanced photodissociation (REPD). If dissociation occurs, the flight time down a second drift tube section provides mass analysis of the

72 DUNCAN fragments. The wavelength dependence of the fragment ion channel(s) as the dye laser is scanned provides a photofragment excitation spectrum, which is equivalent to an absorption spectrum in its line positions. This technique relies on the availability of an efficient dissociation channel, which may not always be found. However, when the method works, it provides electronic spectroscopy with unequivocal identification of the cluster size. M + -Ar Complexes The simplest electrostatic complexes are those formed with rare-gas atoms. Diatomic molecules formed from a single rare-gas atom bound to a metal cation have easily described electronic states, and there is no uncertainty about the qualitative structure. These are therefore ideal systems with which to investigate fundamental electrostatic interactions. Figure 1 shows the electronic Figure 1 The atomic versus molecular orbital interactions in the Mg + -Ar system and the resulting electronic states.

METAL ION COMPLEXES 73 correlation between the atomic states on magnesium cation and those of the Mg + -Ar complex. Ca + -Ar has essentially the same level structure. In order to understand the electronic spectroscopy we measure, it is important to compare the electrostatic interactions in the ground and excited electronic states. The interactions in both states are expected to be charge-induced dipole in nature, and the resulting bonds are likely to be very weak, but there may be significant relative differences in binding strength in different electronic states as a result of the specific orbital interactions. The ground state of the complex correlates to the 2 S atomic ground state, which has a 3s 1 valence electronic configuration on the metal. When this ion binds to a rare-gas atom in its ground state, a 2 molecular state is formed. The binding is relatively weak in this state because the charge-dipole attraction is screened by the repulsion between the valence electrons on the metal ion and the rare-gas atom. The pattern of orbitals is such that two excited molecular states, 2 and 2, correlate to the excited atomic 2 P asymptote. The same kind of screening effect makes the binding in the excited 2 state, which has one electron in a metal p orbital oriented along the molecular axis, even weaker than it is in the ground state. However, the 2 excited state has the metal p orbital oriented perpendicular to the molecular axis. This configuration removes the repulsion in the valence shell and enhances the attractive interactions because the metal core, which resembles a doubly charged atom, is exposed more fully to the rare gas. These general features of orbital interactions suggest that the 2 excited state will be more strongly bound than the 2 excited state. The molecular transitions to be expected are therefore 2 2 and 2 2, and these should be found at lower and higher energies, respectively, than the atomic 2 P 2 S transition. This general pattern should be found for complexes of many metals that are isoelectronic to Mg + in their valence shell (e.g. Na, Ca +,Sr +,Ba +, Ag, Cu). Studies in our laboratory and others have verified this prediction for both cation and neutral metal rare-gas complexes (55). However, the specific energetics in each system reveal many interesting features of weak bonding interactions. Figure 2 shows the low-resolution photodissociation spectrum measured for Mg + -Ar when this ion is produced by laser vaporization, mass-selected, and photodissociated on resonance (45a,b). The spectrum is recorded as the intensity of the Mg + photofragment ion versus the excitation energy. It is measured in each of the isotopic channels ( 24 Mg, 25 Mg, and 26 Mg). The spectrum consists of a series of doublet lines that appear near 319 nm. This is about 4300 cm 1 to the red of the Mg + 2 P 2 S atomic resonance line. The position (to the red of the atomic transition) and the doubled vibronic lines lead to the assignment of the band system as 2 1/2,3/2 X 2 +. The doubled lines arise from the 2 1/2,3/2 spin-orbit splitting. The spin-orbit splitting in the Mg + atom is 91.55 cm 1, but this value should be reduced in the complex by a factor of two thirds to a value near 61.0 cm 1, because angular momentum is recoupled in the molecular

74 DUNCAN Figure 2 The low-resolution photodissociation spectrum of Mg + -Ar, showing the vibrational structure and spin-orbit splittings in the 2 1/2,3/2 2 + band system. Isotopic shifts establish that the first band in the spectrum is the 0-0 band. framework (56). The splittings observed here are about 56 cm 1. Additionally, when isotopic spectra are compared, the isotope shift for each pair of lines is always the same, confirming that they have the same vibrational numbering. These combined observations provide convincing evidence that this is indeed the 2 1/2,3/2 X 2 + band system. The isotopic spectra also provide unequivocal assignment of the vibrational quantum numbers. The 2 1/2 X 2 + progression begins with an 0-0 band at the weak line at 31456 cm 1, and there are a series of transitions of the form v,0 for v = 0 14. Vibrational analysis of these bands produces an excited state frequency of ω e = 271.8 cm 1 and an anharmonicity of ω e x e = 3.25 cm 1. Hotband vibrational transitions (not evident in Figure 2) provide the additional information that the ground state fundamental spacing ( G 1/2 ) is 96 cm 1. The excited state has a long vibrational progression in the 2 state, and the vibrational frequency of this excited state is about three times larger than that in the ground state. This spectrum appears strongly redshifted with respect to the atomic transition in isolated Mg +. All these observations are completely consistent with the picture of electrostatic bonding portrayed in Figure 1, indicating that the bonding in the pπ orbital configuration in the excited state is much stronger than it is in the sσ ground state.

METAL ION COMPLEXES 75 The measurement of an extended vibrational progression such as that shown in Figure 2 reveals not only the excited state vibrational frequency but also the shape and depth of the excited state potential well. If we assume that the excited state can be described by a Morse potential, then an extrapolation of the vibrational levels in a so-called Birge-Sponer plot allows the dissociation energy to be obtained (57). A plot of the spacings between subsequent vibrational levels, G v+1/2, versus the vibrational level, v + 1/2, is extrapolated to the zero intercept to obtain D 0. This treatment results in a binding energy of 5554 cm 1 for the excited 2 1/2 state of Mg + -Ar. Although the Morse analysis is simple and convenient, the determination of dissociation energies by this method may not always provide accurate results for weakly bound complexes, especially if the extrapolation involved is long. Birge-Sponer plots may deviate from linearity at high vibrational levels as a result of the effects of long-range forces at large bond distances. An alternative analysis is the Le Roy Bernstein (LB) treatment for the determination of asymptotic potentials and dissociation energies (58a,b). In this method, if the molecule dissociates according to a C/R n potential, then the vibrational first differences ( G v = [E(v 1) E(v + 1)]/2) are related by the equation G v [2n/(n+2)] D 0 E v, where D 0 is the dissociation energy and E v is the total vibrational energy of the state. A plot of G v [2n/(n+2)] versus E v, with the appropriate choice of the exponent reflecting the physics of the attractive forces, has an intercept equal to D 0. For the particular case here of a charge-induced dipole interaction, the exponent of choice is 4/3. The LB plot determines a dissociation energy for the Mg + -Ar excited 2 1/2 state of 7476 cm 1, which is substantially different from the Morse value. However, the LB method is actually only valid if vibrational levels at or very near the dissociation limit are measured. In the present system, we measure levels for the lower half of the potential, so the standard LB treatment is not expected to perform well. Le Roy has described a scaled version of the LB method to be used in these cases, and he has applied this method to the Mg + -Ar system (59). The scaled LB method produces a dissociation energy virtually the same as our Morse result, lending credence to the simple Morse assumption when levels low in the potential are measured. When excited state dissociation energies are measured in a system such as this with known asymptotic states, an energetic cycle can be used to determine the ground state dissociation energy. The excited state dissociation energy is combined with the measured (or extrapolated) origin band position and the atomic transition energy to obtain the ground state dissociation energy: D 0 = v 00( 2 1/2 2 + ) + D 0 [2 P 1/2 2 S].

76 DUNCAN The ground state bond energy for Mg + -Ar resulting from this cycle is 1295 ± 100 cm 1. The excited 2 state is more than four times more strongly bound than the 2 ground state, again consistent with Figure 1 and the discussion above. This ground state dissociation energy has been measured independently in a recent experiment by Breckenridge and coworkers using their determination for the dissociation energy and the ionization potential of the neutral Mg-Ar van der Waals complex (60). The ground state dissociation energy obtained for Mg + -Ar by this method is 1237 ± 40 cm 1, which is in very good agreement with our value. Our value relies on the Morse approximation for the excited state, and this independent agreement again shows that the Morse approximation works surprisingly well for this system. Another aspect of the electrostatic bonding in Mg + -Ar is the bond distance in the ground and excited states, which can be probed via rotationally resolved spectroscopy (47). Although the weak bonding causes bond distances to be relatively long, which in turn causes small rotational level spacings, the mass of Mg + -Ar is light enough that rotationally resolved spectra can be obtained. Figure 3 shows the rotational structure of the 5,0 vibrational band and a simulation using the constants B = 0.1836 cm 1, B = 0.1409 cm 1, and T rot = 13 K. The bond distances resulting from these constants are r 0 = 2.825 Å and r 5 = 2.475 Å. Consistent with the discussion earlier, the bond distance in the excited 2 state is shorter than it is in the 2 ground state. The bond is 0.35 Å shorter in the more strongly bound 2 state, providing a dramatic illustration of the importance of the specific orbital interactions in electrostatic bonds. Significant insights are also obtained through the comparison of Mg + and Ca + complexes. Figure 4 shows the vibrationally resolved photodissociation spectrum of Ca + -Ar in the analogous 2 1/2,3/2 X 2 + band system (48). In this case, the vibrational frequency and the spin-orbit splitting are nearly the same, and these intervals overlap almost exactly at v = 0, but toward higher vibronic levels the different anharmonicities reveal the two separate spin-orbit progression components. The resulting excited state vibrational frequency is ω e = 164.6 cm 1, while the spin-orbit splitting in the complex is 156 cm 1. This spin-orbit splitting is only slightly larger than the 148 cm 1 value expected by comparison to isolated Ca +, which has a value of 222 cm 1 (42). This spectrum is located near 435 nm, at an energy about 2100 cm 1 to the red of the corresponding 2 P 1/2,3/2 2 S atomic transition in Ca +. As in the Mg + -Ar complex, the redshift indicates that the excited state is more strongly bound than the ground state, consistent with the expectations for pπ versus sσ orbital interactions. The vibronic progression is shorter here than it was for Mg + -Ar, but the Morse extrapolation can be used to determine approximate dissociation energies for the

METAL ION COMPLEXES 77 Figure 3 The high-resolution photodissociation spectrum of the Mg + -Ar (5,0) band, showing the rotational fine structure characteristic of a 2 2 + band. There is a good match between the observed spectrum and the simulation. The analysis of this band yields a ground state bond distance of 2.825 Å and an excited state bond distance of 2.475 Å. The rotational temperature is 13 K. excited state and ground state, as described above. The resulting values are D 0 = 2875 cm 1 and D 0 = 732 cm 1. The relative trend in these bond energies for the ground versus excited state is quite similar to that found for Mg + -Ar, but the absolute values are much less for Ca + -Ar than they are for Mg + -Ar (D 0 = 732 vs 1295 cm 1 ). At first glance, this comparison is surprising. In a chargeinduced dipole interaction, the polarizability of the rare gas is most important, but this is constant for the two complexes. Alkali cations bound to rare gases have been shown to have about 10% contribution to the bonding from the van der Waals interaction term (61, 62). Although the polarizabilities of cations are not generally known, the value for Ca + is probably greater than that for Mg +, which would suggest a stronger bond for the Ca + complex, contrary to our

78 DUNCAN Figure 4 The low-resolution photodissociation spectrum of Ca + -Ar, showing the vibrational structure and spin-orbit splittings in the 2 1/2,3/2 2 + band system. Isotopic shifts establish that the first band in the spectrum is the 0-0 band. The vibrational interval and spin-orbit splitting at low v are approximately the same, and therefore these are not resolved. At higher v levels, the different anharmonicities split the sub-bands into two resolved progressions. observation. However, an even more important consideration in electrostatic interactions is the charge density. Simply put, a metal cation that most closely resembles a point charge can most effectively polarize its bonding partner, and such polarization is essential for the induced dipole interaction. Mg + is a smaller ion and a better approximation to a point charge; therefore, it binds more strongly to argon than Ca + does. In fact, this same charge density effect results in stronger bonding for Mg + compared to Ca + with all the rare gases (Ar, Kr, Xe) (see Table 1) (48). Data are unavailable for such a direct comparison of bond energies for other complexes, but all electrostatic interactions should be enhanced by this charge density effect. M + -CO 2 and M + -N 2 Complexes The same methodology used for rare-gas complexes can be applied to complexes of metal cations with small molecules. However, the details of the electrostatic interactions and electronic spectra are now intimately connected with the structure of the complexes. Complexes with diatomic molecules (e.g. N 2,O 2,H 2 ) or linear triatomic molecules (CO 2 ) may result in linear or T-shaped structures, depending on the details of the interaction. Symmetric systems such as those indicated here have no permanent dipole moments, so the quadrupole moment

METAL ION COMPLEXES 79 Table 1 A comparison of the spectroscopic constants for Mg + and Ca + complexes with rare-gas atoms and small molecules a Experiment Theory Complex ω e (cm 1 ) D 0 (kcal/mol) ω e D 0 References Mg + -Ar 272 3.70 266 2.98 34a,b, 45a,b, 47 Mg + -Kr 258 5.18 257 5.33 34a,b, 45a,b Mg + -Xe 258 11.96 45a,b Mg + -CO 2 382 14.7 359 16.4 29, 43a c Mg + -N 2 558 9.7 33, 51 Mg + -H 2 O 517 25.0 505 30.4 30, 31a, 44a,b Mg + -D 2 O 503 44a,b Ca + -Ar 165 2.10 2.26 31d, 48 Ca + -Kr 149 4.38 3.39 31d, 48 Ca + -Xe 142 6.18 48 Ca + -CO 2 259 50 Ca + -N 2 276 4.49 52 Ca + -H 2 O 358 26.6 31a, 35, 49 Ca + -D 2 O 351 49 a Vibrational frequencies are for the excited state, while dissociation energies are for the ground state. determines the distribution of charge around the molecule and the eventual electrostatic interaction that takes place with a metal cation. O 2,N 2, and CO 2 all have negative quadrupole moments ( 1.33, 4.90 and 14.9 10 40 C m 2, respectively), while H 2 has a positive value (2.1 10 40 C m 2 ) (63). Therefore, simple electrostatics predict that M + -L complexes will be bound primarily through the charge-quadrupole interaction in a linear configuration for N 2,O 2, and CO 2 and in a T-shaped configuration (C 2v symmetry) for H 2. Ab initio calculations predict these same configurations for Mg + complexes (27a,b, 29, 33). Stwalley and coworkers have measured photodissociation spectra for Mg + -D 2, and their analysis is consistent with a C 2v structure (23a,b). Calculations predict that the ground state of Mg + -O 2 undergoes charge transfer to a Mg 2+,O 2 ion pair configuration (27b). We have been able to measure photodissociation spectra for complexes of Mg + and Ca + with CO 2 and with N 2, and all four of these complexes are linear (43a c, 50 52). Figure 5 shows the photodissociation spectrum for Ca + -CO 2 in the region to the red of the Ca + 2 P 2 S atomic resonance (50). If this complex is linear, it will have the same pattern of electronic states as those shown in Figure 1 for Mg + -Ar. The strong spectrum to the red of the atomic transition is therefore assigned as the 2 1/2,3/2 X 2 + band system, analogous to the same systems observed for M + -Ar complexes. Confirmation of this assignment relies on the

80 DUNCAN Figure 5 The photodissociation spectrum of Ca + -CO 2 assigned to the 2 1/2,3/2 2 + band system. The observation of the spin-orbit splitting at approximately two thirds the atomic value confirms that the structure is linear. observation of the required spin-orbit structure. If the complex is linear, the spin-orbit interaction will be observed with about the same magnitude as in the diatomic M + -RG complexes (near 148 cm 1 ), while nonlinear structures would have much smaller spin-orbit splittings probably not noticeable with our resolution (0.5 cm 1 ). To assign spin-orbit structure, we measure isotopic shifts for vibronic lines in the spectrum. This is easily done for Mg + complexes, where the minor isotopes have natural abundances near 10% and where the reduced mass change is greater. However, for 40 Ca + and 42 Ca +, the reduced mass change is smaller and the abundance of the 42 Ca is about 2%. Nevertheless, we have been able to obtain isotopic spectra that show that corresponding spin-orbit components, as labeled in the figure, have the same isotopic shifts. The spinorbit splitting is 136 cm 1, which is slightly lower than the expected value of 148 cm 1, and the excited state vibrational frequency is ω e = 258.9 cm 1. The observation of this spin-orbit splitting confirms that the complex is linear and that this is indeed the expected 2 1/2,3/2 X 2 + system. There is such a short vibrational progression here that extrapolation to determine dissociation energies is unwarranted. However, the observation of this transition with a strong redshift from the atomic resonance line indicates that Ca + -CO 2 has an

METAL ION COMPLEXES 81 electrostatic interaction much like that in the M + -RG complexes. The pπ orbital configuration in these complexes makes the charge-quadrupole interaction more effective than in the sσ configuration, resulting in stronger excited state bonding. This same pattern has now been observed in our lab for Mg + -CO 2 (43a c), Ca + - CO 2 (50), Mg + -N 2 (51), and Ca + -N 2 (52). In each case we observe a 2 1/2,3/2 X 2 + band system red shifted from the corresponding atomic resonance line. All of these complexes are linear, as evidenced by clear excited state spin-orbit doublets. Dissociation energies have been determined for Mg + - CO 2 (14.7 kcal/mol) and Ca + -N 2 (4.49 kcal/mol) (see Table 1). In the Ca + - N 2 complex, the Ca + -N 2 and Ca + N-N stretching modes are both active in the excited state vibronic structure (52). The N-N stretch fundamental is 2487 cm 1, which is higher than the stretching mode frequency (ω e = 2359 cm 1 )in isolated N 2. M + -H 2 O Complexes Perhaps the most important electrostatic interactions in chemistry are those involving the solvation of metal ions by water. Experimental and theoretical investigations of metal ion-water interactions began many years ago. Thermochemical measurements of gas-phase complexes produced in mass spectrometry have determined the energetics of binding to a metal ion center as a function of solvation (8, 10, 11). However, there are still virtually no measurements of the structures of metal ion-water complexes. Our recent data on Mg + -H 2 O and Ca + -H 2 O are the only examples of such measurements to our knowledge. These are admittedly only mono-solvated complexes, but there are no measurements of the structures of complexes with multiply solvated metal ions. Simple electrostatics predicts that a metal cation should bind to a water molecule through a charge-dipole interaction, with the metal cation bound on oxygen in a planar C 2v structure. This is the simple picture portrayed throughout chemistry, and it is upheld by high-level ab initio calculations (30, 31a d, 35 40). We must consider this structure in order to predict electronic states and electronic spectroscopy in M + -H 2 O complexes. Figure 6 shows the electronic states that result for a magnesium or calcium cation in a complex with H 2 O. The ground electronic state of the complex correlating to the 2 S state of the metal atom is 2 A 1. It has a metal ion with one valence electron in an s orbital bound adjacent to oxygen on the C 2 axis. The electrostatic interaction here is relatively weak, analogous to that in the 2 + ground states of linear Mg + and Ca + complexes. In the excited electronic state correlating to the 2 P atomic state, one valence electron in a p orbital on the metal has three possible orientations with respect to the molecule. The 2 A 1 state has the p orbital oriented along the C 2 axis, with significant valence electron density screening the attractive forces. This configuration is analogous to the excited 2 state (pσ orbital

82 DUNCAN Figure 6 The electronic states of Mg + -H 2 OorCa + -H 2 O complexes. configuration) discussed for linear complexes, and the resulting bonding is expected to be extremely weak. The excited p orbital is perpendicular to the C 2 axis and therefore also perpendicular to the dipole moment in the 2 B 1 and 2 B 2 excited states. These states are analogous to the excited 2 state discussed for linear complexes, and the electrostatic interaction should be strongest here. The 2 B 1 state has the p orbital aligned in the same plane as the lone pair orbitals on the oxygen, while the 2 B 2 state has this orbital perpendicular to the plane of the oxygen lone pairs. The latter configuration minimizes the repulsion between the p orbital and the lone pairs, and it is therefore expected to lie at slightly lower energy. The 2 B 1 X 2 A 1 and 2 B 2 X 2 A 1 band systems are therefore expected to be found to the red of the metal atomic transitions, and the 2 A 1 X 2 A 1 transition is expected to lie to the blue of the atomic transition.

METAL ION COMPLEXES 83 We have measured photodissociation spectra for Mg + -H 2 O, Mg + -D 2 O, Ca + - H 2 O, and Ca + -D 2 O in the region to the red of the respective atomic transitions where the 2 B 1 X 2 A 1 and 2 B 2 X 2 A 1 band systems are expected. For both Ca + and Mg + complexes, sharp structured spectra are obtained. The electronic origins of the two expected 2 B 1 X 2 A 1 and 2 B 2 X 2 A 1 band systems are assigned via metal isotope and H 2 O-D 2 O isotope shifts. The spacing between these two electronic states for Mg + -H 2 O is 1869 cm 1. This spacing and the absolute energy of both origins is in good agreement with the predictions of theory (30, 31a d, 37). We therefore assume that the state ordering predicted by theory is correct. Six vibrational modes are possible for these complexes, including three H 2 O-based modes and three metal-h 2 O modes (one symmetric stretch and two bends). For a weakly bound complex, the H 2 O-based modes are not expected to shift dramatically in frequency from the well-known modes for an isolated H 2 O molecule. Any low-frequency intervals are therefore likely to be associated with metal-based modes. This kind of reasoning and the rotational contours described below make it possible to assign all of the vibrational structure in these spectra. The Mg + -H 2 O spectrum is rich in vibronic structure, with a metal-h 2 O stretch progression in both electronic states, an H-O-H scissors mode progression, combinations of these two, and single interval modes assigned to the symmetric and asymmetric O-H stretches and the metal-h 2 O in-plane and out-of-plane bending modes. The scissors motion in the excited state of the complex has about the same frequency that it does in isolated H 2 O (1614 cm 1 ), while the O-H stretches are slightly lower in frequency in the complex than they are in isolated H 2 O. The metal-h 2 O stretching mode frequency is 518 cm 1 in the 2 B 2 excited state, and 489 cm 1 in the 2 B 1 state, compared to frequencies of 505 and 471 cm 1, respectively, predicted by theory (30). Although only five members are observed in the Mg + -H 2 O stretch progression, we can extrapolate this progression to make an estimate for the excited state dissociation energy, and through the cycle described earlier we can estimate the ground state bond energy. The result obtained is that the ground state bond energy for the charge-dipole interaction is 25.0 kcal/mol. This compares with a value calculated by theory of 30.4 kcal/mol (30) and a value from collision-induced dissociation (CID) experiments of 28.4 kcal/mol (14c). Our value is understandably lower than the others, since our extrapolation necessarily makes a pseudo-diatomic approximation and neglects differences between the ground and excited states in the zero-point motions of all the other modes in the system. Reasonable values for the neglected zero-point frequencies suggest that the bond energy is underestimated by about 2 kcal/mol. We therefore believe that the CID result of Armentrout and coworkers is very close to the exact value for the dissociation energy (14c).

84 DUNCAN Figure 7 A portion of the low-resolution photodissociation spectrum of Ca + -H 2 O and Ca + -D 2 O. The rotational contour differences due to the two isotopomers are clearly evident. A similar pattern of two electronic band systems is also observed for Ca + - H 2 O, except that there is very little vibronic activity. Vibronic intervals are observed in the Ca + -H 2 O coordinate, but the progressions are far too short to risk an extrapolation to a dissociation energy. Figure 7 shows a portion of the spectrum measured for Ca + -H 2 O and Ca + -D 2 O in the higher energy 2 B 1 X 2 A 1 system. Vibrational intervals of 358 and 336 cm 1 are observed for the 2 B 2 and 2 B 1 excited states. The splitting between the two electronic states is again about 1800 cm 1, and this spacing is essentially constant for H 2 O and D 2 O complexes. Less vibrational information is therefore obtained for the calcium complexes, but the signals obtained are much stronger than in the Mg + -H 2 O spectrum. This fact, coupled with a newer laser system purchased in the interim, made it possible to investigate the Ca + -H 2 O spectroscopy with significantly improved resolution. It is evident from Figure 7 that there is a significant amount of rotational fine structure associated with the electronic origin and with the weaker vibronic bands. This rotational structure makes it possible to identify the nature of the vibronic bands observed and to eventually determine the structure of both of these metal-water complexes. In the C 2v structure predicted for both Mg + -H 2 O and Ca + -H 2 O, the calculated rotational constants indicate that these species are very close approximations to a symmetric top in both the ground and excited states. This is understandable because the mass is localized on the C 2 axis, with only the hydrogen atoms lying off this axis. The broad rotational contours

METAL ION COMPLEXES 85 observed are therefore associated with the rotational motion of the light hydrogen atoms about the C 2 axis. The rotational spacings are therefore quite large, and the rotational constant should be close to the corresponding value in isolated H 2 O, which is about 14 cm 1. The smaller rotational constants are on the order of 0.1 cm 1 for Ca + -H 2 O and 0.3 cm 1 for Mg + -H 2 O. The qualitative appearance of the rotational contours and the significant change when D 2 O is substituted for H 2 O are consistent with this general picture, but we need a quantitative description of the rotational structure to see if the idealized C 2v structure is indeed accurate. We have therefore modeled the spectrum as a symmetric top, beginning with the rotational constants calculated from the theoretical structure. The simulation includes the standard line intensity formulae for the symmetric top and the additional 3:1 intensity factor alternation for K a odd:even levels caused by the nuclear-spin statistics. For the D 2 O complexes, the rotational constant about the C 2 axis is reduced by a factor of two and the nuclear-spin factor changes to 1:2 for K a odd:even. Rotational temperatures near 10 K are used for both the Mg + and Ca + complexes. The rotational structure expected for vibronic bands in these spectra should also vary with the vibronic symmetry. For totally symmetric (a 1 ) vibronic levels, a perpendicular-type rotational band structure is expected, while parallel-type bands are expected for non totally symmetric bands. The origin band of each of these electronic systems will therefore have a perpendicular-type band structure. The simulation of a perpendicular-type band for the rotational constants from the calculated structures immediately produces a rotational contour that matches the qualitative appearance of our spectra. The contours in Figure 7 are therefore recognizable as perpendiculartype bands at relatively low temperature. Since we do not have complete rotational resolution, we cannot assign specific lines for a complete fit of the spectra. Instead, we adjust the structural parameters to obtain the best match to the partially resolved band contours measured. This procedure eventually produces a single molecular structure for the metal-h 2 O complex that matches the observed spectra for both the H 2 O and D 2 O isotopomers. Figure 8 shows the rotational structure measured for the Ca + -D 2 O complex at the origin band and the comparison to the simulated contour for a symmetric top perpendicular band. The constants for the simulation are A = 7.2 cm 1, A = 6.9 cm 1, (B + C) /2 = 0.23 cm 1, and (B + C) /2 = 0.242 cm 1, and the rotational temperature is about 12 K. A similar analysis was performed for Mg + -H 2 O, albeit it with lower-resolution spectra. The most significant outcome of this rotational analysis is that it becomes possible for the first time to determine the structure of these two metal cation-water complexes. The structures of metal cation-water complexes have been calculated numerous times, but there have been no previous direct measurements of such information. Figure 9 shows the structures determined for Mg + -H 2 O and

86 DUNCAN Figure 8 The high-resolution photodissociation spectrum of Ca + -D 2 O and a corresponding simulation for a symmetric top system. Figure 9 The structures determined for Mg + -H 2 O and Ca + -H 2 O.

METAL ION COMPLEXES 87 Ca + -H 2 O. Both complexes have the planar C 2v structure expected initially for a charge-dipole complex. The more interesting aspects of the structures, however, concern the details of bond distances and angles. The M-O bond distance is 2.03 Å for the Mg + complex and 2.22 Å for the Ca + complex. The shorter bond distance for magnesium is consistent with our previous discussion about the charge density effect in electrostatic bonding. Mg + is a smaller ion and it binds at a shorter distance. A more remarkable effect is that the water molecule in both complexes is distorted from its equilibrium structure. The H-O-H bond angle in isolated water is 104.5. As is often discussed in freshman chemistry, the oxygen atom in water has four regions of electron density around it from the two O-H bonds and the two lone pairs. Unlike methane, however, the bond angle is not 109 because the lone pair electrons sterically crowd the O-H bonds into a smaller region of space. When a metal cation binds to the water molecule along the C 2 axis, it polarizes the lone-pair electron density away from the O-H bonds, relieving this steric crowding, and consequently the H-O-H bond angle expands. This same polarization effect occurs for both the Mg + and Ca + complexes. However, as shown in Figure 9, the H-O-H angle is 110.6 in the Mg + complex and 106.8 in the Ca + complex. The polarization in the Mg + complex is greater, again because of the charge density effect. Because of its smaller size, Mg + binds closer to the oxygen atom. It is also more like a point charge. The combined effect is a much greater polarization-induced bond angle expansion than that seen for Ca +. The polarization effect on the water molecule is also different in different electronic states of the complexes, corresponding to different valence orbital interactions. Figure 9 shows the structures for the respective 2 A 1 ground state complexes, which have the metal valence electron in an s orbital localized on the metal. As discussed earlier, this kind of configuration is not the most favorable for electrostatic interactions because the valence shell repulsion is significant. In the excited 2 B 1 and 2 B 2 states, the metal valence electron resides in a p orbital that is perpendicular to the C 2 axis. This configuration removes the valence electron repulsion and enhances the attractive interaction by exposing the positive metal core. This interaction is most effective in the 2 B 2 state, which has the p orbital perpendicular to the oxygen lone pairs, thus producing a slightly greater binding energy, as noted above. These excited state orbital interactions are also evident in the complex structures. The Ca + -O bond distance is 2.13 Å in the 2 B 2 state and 2.17 Åinthe 2 B 1 state; both are shorter than in the ground state. The polarization distortion on the water molecule is also greater in the excited state. The H-O-H angle is 110.7 in the 2 B 2 state and 110.3 in the 2 B 1 state. The enhanced excited state interactions lead to shorter bonds and more polarization of the water molecule.

88 DUNCAN CONCLUSIONS The systematic studies of these related metal complexes, which are summarized in Table 1, reveal many interesting details of electrostatic bonding and several fundamental concepts. The isoelectronic Mg + -L and Ca + -L complexes have the same patterns of ground and excited electronic states for corresponding ligands. In every complex, the band systems studied are strongly redshifted from the atomic resonance line, indicating that the excited state bond energy is significantly greater than the ground state bond energy. The excited states studied all correspond to a metal atom with a valence electron in a p orbital oriented perpendicular to the ligand. This configuration is compared to ground states with metal valence electrons in an s orbital. Metal orbital configurations that minimize valence electron repulsion are therefore a major consideration in the electrostatic bonding. Similar arguments based on valence shell orbital interactions explain the stronger excited state bonding observed for all of these complexes, the shorter excited state bond distances, and the differences in ligand polarization in the M + -H 2 O complexes as a function of electronic state. The next consideration readily apparent throughout these studies is the order of the electrostatic interaction. The metal rare-gas complexes are the most weakly bound, where the interaction is charge-induced dipole. Over the series of Ar, Kr, Xe complexes the rare-gas polarizability increases, and consequently the induced dipole bond energy increases. The charge-quadrupole complexes are more strongly bound than the rare-gas complexes. The negative quadrupole moments of N 2 and CO 2 make these complexes linear, and the smaller quadrupole moment of N 2 makes these complexes more weakly bound than the corresponding CO 2 complexes. The most strongly bound complexes are the charge-dipole M + -H 2 O systems, which have C 2v structures. It will be interesting in the future to study additional charge-dipole systems such as M + -NH 3 or M + -CH 3 OH to investigate the hierarchy of bond strengths and the additional variety of structures expected. The final underlying concept revealed in these studies is the importance of metal charge density. Although all dissociation energies have not yet been determined, it is apparent that the magnesium ion complexes are uniformly more strongly bound than the corresponding calcium complexes. Where bond distances have been measured, the magnesium-ligand bond distance is shorter than the corresponding calcium-ligand distance. In the water complexes, where the first evidence for ligand polarization and distortion is observed, the effect is greatest for magnesium. There are few other opportunities for comparison of this effect. However, Brucat and colleagues have measured the bond distance in Co + -Ar (2.385 Å) (22a), where the metal ion is smaller than Mg +. Miller and coworkers have measured the bond distance for Ba + -Ar (3.36 Å) (25),

METAL ION COMPLEXES 89 where the metal ion is larger than Mg +. These values compare to the ground state Mg + -Ar value of 2.825 Å (47). The bond energies for the series Co + -Ar, Mg + -Ar, Ca + -Ar, Sr + -Ar, and Ba + -Ar are, respectively, 4097, 1295, 732, 800, and 650 cm 1 (21a h, 25, 45a,b, 48, 64). For this series, the bond distance and the bond energy both follow the trend expected on the basis of the metal ion size and charge density. The relative importance of these three factors is also interesting to consider. For example, the excited state 2 bond energy for Mg + -Ar is 5554 cm 1 (45a,b), while the ground state bond energy for Mg + -CO 2 is 5140 cm 1 (43a c). A favorable orbital interaction therefore outweighs a higher-order electrostatic interaction. The excited 2 state bond energy for Ca + -Ar (48) is 2875 cm 1 compared to the ground state bond of 1295 cm 1 for Mg + -Ar (45a,b). In this comparison, a favorable orbital interaction outweighs the greater charge density in the smaller ion. Thus, in these two examples, the valence orbital interactions that reduce repulsion are notably more important than other considerations in electrostatic bonding. A similar conclusion has been reached by Breckenridge et al in their consideration of neutral metal van der Waals complexes (55). These various properties of ground and excited Mg + -L complexes have been investigated extensively in theoretical calculations by Bauschlicher and coworkers (26 34b). As shown in Table 1, there is excellent agreement between the structural properties, bond energies, electronic state energies, and vibrational frequencies calculated by theory and those measured. Theoretical studies have also been carried out on complexes that have not yet been studied spectroscopically (Mg + -NH 3,Mg + -CH 4,Mg + -CH 3 OH) (26, 28, 30). Some very interesting predictions have been made regarding double ligand complexes (29, 31a d). Mg + -(Ar) 2,Mg + -(CO 2 ) 2, and Mg + -(H 2 O) 2 are all predicted to be bent complexes. This arises because the first ligand induces a back-polarization of valence electron density on the metal, producing a lobe of electron density protruding out the back side of the metal atom. This electron density causes additional valence shell repulsion on the side of the metal ion opposite the first ligand, and so the second ligand adds at a bent position. In contrast to this situation for magnesium, di-ligand complexes of calcium and strontium are predicted to be linear. These are very specific predictions of complex structure that could be investigated by spectroscopy of multiple-ligand complexes. Theoretical studies of calcium and strontium ion complexes are limited, and they have focused for the most part on the ground electronic state. Additional work in this area is particularly needed because of the new experimental work now becoming available. If metal complexes in the gas phase are to be viable models for metal ion solvation, it is essential to extend these studies to larger complexes. However, this has not yet been done with any great success. Both our group and those

90 DUNCAN of Farrar (20a f) and Fuke (24a e) have investigated metal ions in multipleligand complexes. Low-resolution spectra have been obtained that show the position of electronic states, but no vibrationally resolved spectra have been obtained, and no detailed structures or energetics could be derived. There are several reasons for these difficulties, including poor Franck-Condon factors, congestion caused by low-frequency vibrational modes, and lifetime broadening due to excited state insertion chemistry. The latter issue is pervasive in multiple-ligand complexes and is also seen for many mono-ligand complexes. Many metal ions including magnesium, calcium, and strontium have endothermic insertion chemistry with small molecules in the ground electronic state. This makes it possible to form electrostatic complexes as opposed to covalent insertion complexes. However, upon optical excitation at visible and ultraviolet wavelengths, the complexes lie at energies above the thresholds for reaction. Reactive channels are already seen in the photodissociation of, for example, Mg + -H 2 O, Mg + -CO 2,Ca + -CO 2,Ca + -H 2 O, Mg + -H 2, and Sr + -CH 3 OH monomer complexes, where, for example, metal oxide, hydroxide, and hydride channels are observed. In most cases, the same sharp spectra are observed in the M + and MO + fragment ion channels, confirming that the reaction occurs after absorption to the excited state resonances. In multiple-ligand complexes, these reactive channels often begin to dominate, and no sharp structure is observed. The additional complication of metal-ligand charge transfer also produces broad spectra in complexes in which the metal and ligand have closely spaced ionization potentials (19a,b). To avoid excited state reactive processes, it would be useful to probe these complexes at levels in the ground electronic state where photoinduced chemistry is still endothermic. However, the bond energies are so great and vibrational frequencies so low in general that infrared excitation in the ground state cannot produce photodissociation. The density of ions is too low for direct absorption techniques, and such techniques would also lose the advantage of mass selection. To circumvent these difficulties, we have attempted mass-analyzed threshold photoionization spectroscopy (MATI), which is a close relative of ZEKE spectroscopy. This technique has been successful for Al + -Ar and Al + -Kr, where photoionization of the corresponding neutral van der Waals complexes produces ground state cation complexes with vibrationally resolved structure in the spectrum (65). However, MATI and ZEKE are still problematic for metal complexes and have not been applied to multiple-ligand species (66). It is also interesting to probe the ground electronic states of metal ion complexes, even those whose excited state spectra have already been measured. Dispersed fluorescence and stimulated emission pumping schemes are being actively pursued to reveal the detailed nature of the electrostatic potential surfaces in these systems.