Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

Similar documents
Resonant enhanced electron impact dissociation of molecules

Supporting Information. Dynamics of the Dissociating Uracil Anion. Following Resonant Electron Attachment

Colliding Crystalline Beams

RESONANCE INTERFERENCE AND ABSOLUTE CROSS SECTIONS IN NEAR-THRESHOLD ELECTRON-IMPACT EXCITATION OF MULTICHARGED IONS

Transverse momentum of ionized atoms and diatomic molecules acquired in collisions with fast highly-charged heavy ion. V. Horvat and R. L.

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

National Accelerator Laboratory

Applications of Pulse Shape Analysis to HPGe Gamma-Ray Detectors

Using the X-FEL to understand X-ray Thomson scattering for partially ionized plasmas

Start-up Noise in 3-D Self-AmpMed

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

INTERMOLECULAR POTENTIAL FUNCTIONS AND HIGH RESOLUTION MOLECULAR SPECTROSCOPY OF WEAKLY BOUND COMPLEXES. Final Progress Report

Multicusp Sources for Ion Beam Lithography Applications

Uncertainty in Molecular Photoionization!

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

Modeling Laser and e-beam Generated Plasma-Plume Experiments Using LASNEX

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

Bandgap Photons to GaInP2

Imaging the indirect dissociation dynamics of temporary negative ion: N 2 O

Recoil ion and electronic angular asymmetry parameters for photo double ionization of helium at 99 ev

Simulation of Double-Null Divertor Plasmas with the UEDGE Code

SHAPE RESONANCE IN PHOTOELECTRON SPECTROSCOPY

PROJECT PROGRESS REPORT (03/lfi?lfibr-~/15/1998):

Photoelectron Spectroscopy of the Hydroxymethoxide Anion, H 2 C(OH)O

GA A26474 SYNERGY IN TWO-FREQUENCY FAST WAVE CYCLOTRON HARMONIC ABSORPTION IN DIII-D

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

arxiv: v1 [physics.chem-ph] 29 Jul 2010

arxiv: v1 [physics.atm-clus] 19 Aug 2016

Analysis of Shane Telescope Aberration and After Collimation

A lattice dynamical investigation of zircon (ZrSiOJ has been carried out to obtain a

Nonlocal model of dissociative electron attachment and vibrational excitation of NO

5.80 Small-Molecule Spectroscopy and Dynamics

National Accelerator Laboratory

GA A26785 GIANT SAWTEETH IN DIII-D AND THE QUASI-INTERCHANGE MODE

Scaling between K+ and proton production in nucleus-nucleus collisions *

DML and Foil Measurements of ETA Beam Radius

ELECTRON IMPACT IONIZATION OF HELIUM [(e,2e) & (e,3e)] INVESTIGATED WITH COLD TARGET RECOIL-ION MOMENTUM SPECTROSCOPY

The Gas Flow from the Gas Attenuator to the Beam Line

AC dipole based optics measurement and correction at RHIC

August 3,1999. Stiffness and Strength Properties for Basic Sandwich Material Core Types UCRL-JC B. Kim, R.M. Christensen.

Two-Screen Method for Determining Electron Beam Energy and Deflection from Laser Wakefield Acceleration

Absolute Integral and Differential Cross Sections for the Reactive Scattering of H - + D 2 and D - + H 2

Curvature of a Cantilever Beam Subjected to an Equi-Biaxial Bending Moment. P. Krulevitch G. C. Johnson

Electron Transfer of Carbonylmetalate Radical Pairs: Femtosecond Visible Spectroscopy of Optically Excited Ion Pairs

Vibrational Autoionization in Polyatomic molecules

4kU. Measurement of Storage Ring Motion at the Advanced Light Source. QSTt ERNESTORLANDO LAWRENCE BERKELEYNATIONAL LABORATORY

Plasma Response Control Using Advanced Feedback Techniques

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

Scaling of divertor heat flux profile widths in DIII-D

J. C. Batchelder', K. S. Toth2, D. M. Moltz3, E. F. Zganjarl, T. J. Ognibene3, M. W. Rowe3, C. R. Binghan12.~,J. Powell3, and B. E.

N = 2 String Amplitudes *

CONVERGENCE STUDIES OF THE DOUBLE PHOTOIONIZATION OF Li + AND He

Shock compression of precompressed deuterium

Resonances in Chemical Reactions : Theory and Experiment. Toshiyuki Takayanagi Saitama University Department of Chemistry

SMALL BIPOLARONS IN BORON CARBIDES: PAIR BREAKING IN SEMICLASSICAL HOPPING* David Emin Sandia National Laboratories Albuquerque, NM , USA

Design of Current Leads for the MICE Coupling Magnet

ADSORPTION ON NANOSURFACES: A DETAILED LOOK AT METAL CLUSTERS USING INFRARED SPECTROSCOPY

Undulator Interruption in

UPGRADED CALIBRATIONS OF THE THOMSON SYSTEM AT DIII D

GA A26686 FAST ION EFFECTS DURING TEST BLANKET MODULE SIMULATION EXPERIMENTS IN DIII-D

Energy and angular dependence of H - (D - ) ions produced by dissociative electron attachment to H 2 O(D 2

GA A23713 RECENT ECCD EXPERIMENTAL STUDIES ON DIII D

Optimization of NSLS-II Blade X-ray Beam Position Monitors: from Photoemission type to Diamond Detector. P. Ilinski

AuttWr(s): A. Blotz, Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA

ION EXCHANGE SEPARATION OF PLUTONIUM AND GALLIUM (1) Resource and Inventory Requirements, (2) Waste, Emissions, and Effluent, and (3) Facility Size

9 7og$y4- International Conference On Neutron Scattering, Toronto August Spin Dynamics of the reentrant spin glass Fe0.7A10.3.

Alex Dombos Michigan State University Nuclear and Particle Physics

SOME ENDF/B-VI MATERLALS. C. Y. Fu Oak Ridge National Laboratory Oak Ridge, Tennessee USA

CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions

Complete analysis of the Na q fragmentation in collision with He

J. R Creighton and C. M. Truong Org. 1126, MS 0601 P.O. Box 5800 Sandia National Laboratories Albuquerque, NM

GA A27235 EULERIAN SIMULATIONS OF NEOCLASSICAL FLOWS AND TRANSPORT IN THE TOKAMAK PLASMA EDGE AND OUTER CORE

Multi-scale modeling with generalized dynamic discrepancy

Survey, Alignment and Beam Stability at the Advanced Light Source

The temperature dependence of the C1+ propylene rate coefficient and the Arrhenius fit are shown in Figure 1.

To be published in the Proceedings of ICEC-22, Seoul Korea, July 2008 MICE Note 232 1

Capabilities for Testing the Electronic Configuration in Pu

Electron impact ionization of diatomic molecules

Electron attachment to SF 6 and lifetimes of SF 6 negative ions

Determine the Inside Wall Temperature of DSTs using an Infrared Temperature Sensor

GA A22722 CENTRAL THOMSON SCATTERING UPGRADE ON DIII D

ust/ aphysics Division, Argonne National Laboratory, Argonne, Illinois 60439, USA

Bulk Modulus Capacitor Load Cells

MAGNETIC RESONANCE IMAGING OF SOLVENT TRANSPORT IN POLYMER NETWORKS

GA A25853 FAST ION REDISTRIBUTION AND IMPLICATIONS FOR THE HYBRID REGIME

FUSION WITH Z-PINCHES. Don Cook. Sandia National Laboratories Albuquerque, New Mexico 87185

GA A27857 IMPACT OF PLASMA RESPONSE ON RMP ELM SUPPRESSION IN DIII-D

Cross sections for eV e-h 2 resonant collisions: Vibrational excitation

Absolute differential cross sections for electron elastic scattering and vibrational excitation in nitrogen in the angular range from 120 to 180

CQNl_" RESPONSE TO 100% INTERNAL QUANTUM EFFICIENCY SILICON PHOTODIODES TO LOW ENERGY ELECTRONS AND IONS

PROJECT PROGRESS REPORT (06/16/1998-9/15/1998):

PRINCETON PLASMA PHYSICS LABORATORY PRINCETON UNIVERSITY, PRINCETON, NEW JERSEY

Novel Nanoparticles for Ultrasensitive Detection and Spectroscopy

5.61 Physical Chemistry Final Exam 12/16/09. MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Chemistry Chemistry Physical Chemistry

A Distributed Radiator, Heavy Ion Driven Inertial Confinement Fusion Target with Realistic, Multibeam Illumination Geometry

How Small Can a Launch Vehicle Be?

e, )2 Crossing Angles In The Beam-Beam Interaction R.H. Siernann* II. PERTURBATION FORMALISM SLAC-PUB-7225

On the HalfiLife of LABORATORY ERNEST ORLANDO LAWRENCE BERKELEYNATIONAL. and Particle Astrophysics Nuclear Science and Physics Divisions

SciDAC CENTER FOR SIMULATION OF WAVE-PLASMA INTERACTIONS

Clifford K. Ho and Michael L. Wilson Sandia National Laboratories. P.O. Box Albuquerque, NM

Transcription:

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory Title Dissociative Electron Attachment to Carbon Dioxide via the 8.~;;eV Feshbach resonance Permalink https://escholarship.org/uc/item/59t9171b Author Slaughter, Dan Publication Date 1-9-6 Peer reviewed escholarship.org Powered by the California Digital Library University of California

Dissociative Electron Attachment to Carbon Dioxide via the 8. ev Feshbach resonance D. S. Slaughter, 1 H. Adaniya, 1 T. N. Rescigno, 1 D. J. Haxton, 1 A. E. Orel, C. W. McCurdy, 1, 3 and A. Belkacem 1 1 Lawrence Berkeley National Laboratory, Chemical Sciences, Berkeley, California 947, USA Department of Applied Science, University of California, Davis, CA 95616, USA 3 Departments of Chemistry and Applied Science, University of California, Davis, CA 95616 Momentum imaging experiments on dissociative electron attachment (DEA) to CO are combined with the results of ab initio calculations to provide a detailed and consistent picture of the dissociation dynamics through the 8. ev resonance, which is the major channel for DEA in CO. The present study resolves several puzzling misconceptions about this system. PACS numbers: 34.8.Ht Negative ion resonances are ubiquitous in low-energy electron-molecule collisions and provide an efficient vehicle for the transfer of electronic energy to nuclear motion either through vibrational excitation or dissociative electron attachment, the latter process resulting in the formation of both charged and neutral fragments. Recent dynamical studies [1] have shown that DEA to fundamental polyatomic systems can exhibit complex electronic and nuclear dynamics involving symmetry breaking target deformations [] and, in some cases, conical intersections [3, 4]. Mechanistic studies of the DEA process may give insight into their behavior in the condensed phase [5] and in biological environments [6]. Carbon dioxide offers an interesting case in point. The inverse of DEA to CO, i.e. associative detachment, is thought to be important in the catalytic oxidation of CO on a metal surface [7]. In light of its fundamental importance to the understanding of such processes, it is noteworthy that the electronic structure of CO and its metastable anions has not been completely characterized. Most of the extant literature on low-energy electron-co scattering deals with the short-lived Π u shape resonance near 4 ev, which provides the dominant mechanism for vibrational excitation, while experimental studies of DEA to CO [8 14] have focused mainly on total cross sections and their dependence on electron energy and ion kinetic energy release. The Π u resonance also feeds the CO( 1 Σ + ) + O ( P) DEA channel whose thermodynamic threshold lies at 3.99 ev. Scattering calculations [15] show that the Π u resonance becomes sharper and finally electronically bound as the CO bonds are increased along the symmetric stretching coordinate. It is also known that the CO ion becomes stable upon bending. It was a long-held belief [16 18] that one component ( A 1 ) of the Π u resonance, which splits into A 1 and B 1 components upon bending, correlates with the stable anion, but as Sommerfeld et al. [19] have shown, this is not the case. The stable anion correlates with a second, lower energy A 1 state which, in linear geometry, becomes a virtual state that dominates electron scattering below 1 ev []. By symmetry, there are only three electronic states that can correlate to the CO( 1 Σ + ) + O ( P) asymptote. One of these is the A 1 state that, in linear geometry, becomes the virtual state. If, as is commonly believed [5, 1], the Π u resonance accounts for the other two states, then we are led to the puzzle whose resolution is a subject of this Letter. The dominant DEA channel in CO is observed at 8. ev. Since this energy is less than the 1. ev required to produce electronically excited CO* + O, the 8. ev resonance must necessarily result in electronic ground-state products. So how is this possible when, according to current thinking, all three states arising from this asymptote have already been accounted for? The early theoretical work of Claydon et al. [1] and England et al. [] assigned the 8. ev peak to a Σ + g shape resonance. This assignment has since been disputed. Srivastava and Orient [13], having found little or no dependence of the 8. ev DEA peak on vibrational excitation of the target, suggested it was a Feshbach resonance, citing unpublished theoretical work by Winter supporting their conclusion. Dressler and Allan [14] and, more recently, Huels et al. [5] reached a similar conclusion, although there is no consensus about the symmetry of the Feshbach state nor its parent target state and the question of how such a state could feed ground-state products was never addressed. Our experimental setup, consisting of a momentum spectrometer and an ion detection scheme similar to that used in cold target recoil ion momentum spectroscopy (COLTRIMS) [3], is the same one used in our earlier study of DEA to water [1] and so will not be described in detail here. A stainless steel capillary was used to produce an effusive jet of CO molecules which was crossed at 9 with a pulsed electron beam. The absolute electron energy was determined and checked periodically by measuring the thermodynamic threshold for O production from CO, while the momentum spectrometer was calibrated against the well-known O momentum distribution from DEA to O. The ion kinetic energy and angular resolution were limited by thermal broadening of the effusive target beam which increases with the square

ion counts (arb. units)..4.6.8 1. 1. 1.4 ion kinetic energy (ev) FIG. 1: (Color online) Measured O kinetic energy distribution for an incident electron energy of 8.1 ev (solid curve). The experimental data of Chantry [1] (circles) and the uncorrected (triangles) and corrected (squares) data of Dressler and Allan [14] are shown for comparison. root of the mean kinetic energy[4]. For the present measurements we estimate the overall kinetic energy resolution to be. ev FWHM, for a mean O kinetic energy of.7 ev, while the overall angular resolution at this energy was estimated to be 4 FWHM. The kinetic energy distribution of O, for electron attachment energy just below the resonance peak of 8. ev, is displayed in Fig. 1, along with previous experimental data found in the literature. The present measurements and the data of Chantry [1] were normalized to the 8.3 ev electron energy corrected data of Dressler and Allan [14] at the.6 ev O kinetic energy peak, while the uncorrected measurements (triangles in Fig. 1) of the latter work have been maintained on the same scale as their corrected data. The kinetic energy distribution measured by Dressler and Allan depended on the transmission function of their spectrometer, which decreased like 1/E k with increasing ion kinetic energy E k. They attempted to remove this dependence by weighting the distribution with E k. Qualitatively, the present data lie between their corrected and uncorrected data, indicating that their correction was somewhat overestimated. The distribution of the present data consists of two peaks, the larger peak having a maximum at.6 ev and the smaller one peaking at. ev. Such a two-peaked structure was calculated for the CO ( Σ + g ) potential energy surface by Sizun and Goursaud [5], but as stated above, more recent evidence [5, 13, 14] suggest that the Σ + g state is not responsible for the 8. ev resonance and that the correct assignment is a doubly excited (Feshbach) state. Since -body breakup is the only open channel at these energies, we can determine the kinetic energy distribution of the neutral fragment and its occupied vibrational states directly from the ion kinetic energy distribution using conservation of energy. The nominal electron beam energy is 8.1 ev, which leaves 4.5 ev of excess energy above the thermodynamic threshold, which is the differ- ion counts (arb. units) 3 6 9 1 15 18 ion ejection angle (deg.) FIG. : (Color online) Left: laboratory-frame O momentum distribution from DEA to CO for a incident electron energy of 8.1 ev. The arrow indicates the incident electron direction and the intensity scale is the ion yield in arbitrary units. Right: Laboratory-frame O angular distribution, with respect to the electron beam direction ( o ), for the low-energy (circles) and high-energy (triangles) peaks in the O kinetic energy distribution. Error bars represent an estimate of the statistical uncertainty. ence between the CO + O dissociation energy and the electron affinity of O [4]. The small peak with zero kinetic energy release is a result of dissociation leading to highly vibrationally excited (ν = 16) CO(X 1 Σ + ) fragments. The main peak spanning O kinetic energies of.4 ev to 1.1 ev corresponds to a total kinetic energy release, shared between the two fragments, of.6 ev and 1.7 ev, which implies population of mostly the ν = 9 to ν = 13 vibrational levels of the CO ground state. To investigate the origin of the main peak in the O kinetic energy distribution, we examine the ion momentum and angular distributions with respect to the electron beam direction. The momentum distribution, displayed in the left panel of Fig., shows an asymmetric angular distribution, peaking at wide backward angles with respect to the electron beam direction. Ion angular distributions for each of the two peaks in the kinetic energy distribution are ldisplayed in the right panel Fig., where we have integrated over two subsets of the ion kinetic energy in order to separate the two distributions. The lowkinetic energy angular distribution tends to considerably wider angles than the main kinetic energy peak, which is consistent with the slower ions following very different trajectories compared to the faster ions. To assist with the interpretation of the measured data, we carried out both ab initio electronic structure and fixed-nuclei electron scattering calculations. Neutral CO is a linear closed-shell molecule nominally described, near its equilibrium geometry, by the electronic configuration (core) 1 (σ g ) (σ u ) (π g ) 4 (π g ) 4. The neutral target states were described by complete-activespace (CAS) configuration-interaction (CI) calculations with state-averaged multi-configuration-self-consistentfield (MCSCF) orbitals, in which we doubly occupied the first five orbitals, and included an additional π and a Rydberg σ orbital in the active space. The negative

3 Energy (ev) 1 8 6 4 3 Πg Πg Πu Σg + CO - ( Π) + O( 1 D) CO + O( 1 D) CO + O( 3 P) CO + O - ( P) X 1 + Σ g.5 3 3.5 4 4.5 5 5.5 6 OC --- O (Bohr) A Elastic Cross Section (atomic units) 5 15 1 5 7.4 7.5 7.6 7.7 Electron Energy (ev) FIG. 3: (Color online) Left: collinear potential energy curves for CO (dashed) and CO (solid). One CO distance is fixed at.195 bohr. Vertical line indicates equilibrium geometry. Right: Π g component of e - CO elastic cross section at equilibrium geometry. ion states were then obtained by carrying out multireference CI calculations consisting of a CAS CI + all single-excitations into virtual orbitals. Such a treatment attempts to strike a balanced description of correlation in the neutral and negative ion states, although it is admittedly biased toward the anion states. The results of these calculations are summarized in Fig. 3a, which shows a cut through the potential energy surfaces in linear geometry where one CO distance is fixed and the other varied. In plotting the results, the anion states were all shifted upward by.7 ev relative to the neutral ground state, said shift chosen to make the Σ + virtual state coincide with the neutral ground state in the region where the former is unbound. In addition to the Π u anion, we find a doubly excited (π 3 g σ ), Π g negative ion state, whose parent is the 3 Π g excited state of the neutral target. We note that near the equilibrium geometry of the neutral, the Π g state lies below all the electronically excited CO states and can thus only autodetach into the e + CO (X 1 Σ + g ) continuum. The surprising result is that it is the Π Feshbach state, along with the Σ + virtual state, that correlate with CO( 1 Σ + ) + O ( P) in linear geometry. There is a sharp avoided crossing between the Π shape resonance and the Π doubly excited state in linear geometry, the former correlating with a short-lived CO, Π anion + O ( 1 D). There is in fact a conical intersection between the Π states close the to point where they avoid in linear geometry. To further characterize the doubly excited state, we carried out fixed-nuclei complex Kohn scattering calculations with the same prescriptions for constructing the N- electron target states and N+1-electron scattering states that were employed in the structure calculations. Figure 3b shows the Π g component of the elastic cross section at equilibrium geometry near the resonance energy. A Breit-Wigner fit to the results confirmed the resonance to be extremely long-lived, with a width of.4 ev. We must emphasize that these results are for a single geometry. When convoluted over vibrationally weighted B geometries in the Franck-Condon region, no sharp features would be expected to be seen in the elastic cross section, which explains why the observed 8. ev DEA peak is not visible in the transmission spectrum [14]. To connect the theoretical results to the observed laboratory-frame angular distributions, we calculate the entrance amplitude, formally defined as V (θ, φ; S) = Ψ + bg (θ, φ;s) H el Ψ res (S), where Ψ + bg is a background scattering function with a plane-wave incident on the target in the direction θ, φ, Ψ res is the resonance wavefunction, H el is the electronic part of the Hamiltonian and S labels the internal coordinates of the molecule. The electron attachment probability, a function of θ and φ expressed relative to the dissociation axis in the molecular frame, is computed from the squared modulus of the entrance amplitude. In practice, we can evaluate the entrance amplitude in terms of quantities obtained from an analysis of the calculated fixed-nuclei S-matrix, as outlined in ref. [6]. The attachment probability can be directly related to a laboratory frame angular distribution when the axial recoil condition is met, requiring in the present case that the recoil axis which connects the atom ion and the diatom center of mass does not rotate during the dissociation. Under these assumptions, the laboratory angular distribution is obtained by averaging the attachment probability over initial and final target rotational states, which eliminates its dependence on φ. Since asymmetry is clearly observed in the angular distributions, it is important to incorporate the effects of zero-point bending and asymmetric stretch motion into the calculation of the entrance amplitudes. In order to simplify the calculations while gauging the importance of these effects, we obtained the entrance amplitudes from scattering calculations with the nuclei located at their root-mean square (RMS) values, assuming harmonic asymmetric stretch and bending potentials. Figure 4 shows the calculated entrance amplitudes and angular distributes at both the equilibrium and RMS geometries. While the angular distribution computed at the RMS geometries gives results in better agreement with experiment than the equilibrium geometry result, the magnitude of the observed asymmetry clearly implies that there must be post-attachment bending involved in the dissociation dynamics. To show this effect, the rightmost panel of Fig. 4 shows results (labelled convolved ) in which the angular distribution is calculated by adding 3 degrees to the axial recoil angle and convolving the computed values with a 55 FWHM Gaussian distribution to simulate the finite resolution of the experiment. The agreement with the experimental distribution at.7ev O KE is quite good. These results indicate that it is likely that the electron attaches preferentially on the stretched side of the molecule and subsequent dynamics tends to favor dissociation in linear geometry. The present results provide new insight into the topol-

4 Equilibrium Geometry RMS Counts Counts 45 9 135 18 Scattering angle (degrees).7ev KE Axial recoil Convolved 45 9 135 18 Scattering angle (degrees) FIG. 4: (Color online) Entrance amplitudes and angular distributions for CO DEA through the Π Feshbach resonance. Left to right: real and imaginary parts of the entrance amplitude calculated at RMS values of bend and asymmetric stretch; calculated O angular distributions at equilibrium and RMS geometries; comparison of calculated and measured angular distributions (see text). ogy of the CO anion states and the dynamics of dissociative attachment in this system. This study again illustrates that, even with small polyatomic targets, an understanding of anion dissociation dynamics beyond simple one-dimensional models can be crucial in interpreting measured data. By combining the results of momentum imaging spectroscopy with ab initio theory we are able to clearly show that the 8. ev DEA peak in CO is initiated by electron attachment to a dissociative, doubly excited Π state that interacts with a lower Π shape resonance through a conical intersection and dissociates to electronic ground-state products. Mapping out the conical intersection(s) is complicated since bending breaks the degeneracy of the Π states, resulting in a pair of A and A states with different topologies. This topology, and the role it plays in understanding DEA through the 4 ev shape resonance, will be the subject of a longer paper. This work was performed under the auspices of the US DOE by LBNL under Contract DE-AC-5CH1131 and was supported by the U.S. DOE Office of Basic Energy Sciences, Division of Chemical Sciences. This document was prepared as an account of work sponsored by the United States Government. While this document is believed to contain correct information, neither the United States Government nor any agency thereof, nor the Regents of the University of California, nor any of their employees, makes any warranty, express or implied, or assumes any legal responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by its trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof, or the Regents of the University of California. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof or the Regents of the University of California. [1] H. Adaniya, B. Rudek, T. Osipov, D. J. Haxton, T. Weber, T. N. Rescigno, C. W. McCurdy, and A. Belkacem, Phys. Rev. Lett. 13, 331 (9). [] T. N. Rescigno, C. S. Trevisan, and A. E. Orel, Phys. Rev. Lett. 96, 131 (6). [3] D. J. Haxton, T. N. Rescigno, and C. W. McCurdy, Phys. Rev. A 75, 1711 (7). [4] H. Adaniya, B. Rudek, T. Osipov, D. J. Haxton, T. Weber, T. N. Rescigno, C. W. McCurdy, and A. Belkacem, in preparation (11). [5] M. A. Huels, L. Parenteau, P. Cloutier, and L. Sanche, J. Chem. Phys. 13, 6775 (1995). [6] B. Boudaïffa, P. Cloutier, D. Hunting, M. A. Huels, and L. Sanche, Science 87, 1658 (). [7] S. N. Maximoff and M. P. Head-Gordon, PNAS 16, 1146 (9). [8] G. J. Schulz, Phys. Rev. 18, 178 (196). [9] D. Rapp and D. D. Briglia, The Journal of Chemical Physics 43, 148 (1965). [1] P. J. Chantry, J. Chem. Phys. 57, 318 (197). [11] R. Abouaf, R. Paineau, and F. Fiquet-Fayard, Journal of Physics B: Atomic and Molecular Physics 9, 33 (1976). [1] M. Tronc, L. Malegat, and R. Azria, Chemical Physics Letters 9, 551 (198). [13] S. K. Srivastava and O. J. Orient, Phys. Rev. A 7, 19 (1983). [14] R. Dressler and M. Allan, Chemical Physics 9, 449 (1985). [15] T. N. Rescigno, W. A. Isaacs, A. E. Orel, H. -D. Meyer and C. W. McCurdy, Phys. Rev. A 65, 3716 (). [16] D. G. Hopper, Chem. Phys. 53, 85 (198). [17] W. B. England, Chem. Phys. Lett. 78, 67 (1981). [18] Y. Yoshioka and H. F. Schaefer III and K. D. Jordan, J. Chem. Phys. 75, 14 (1981). [19] T. Sommerfeld, H. -D Meyer and L. S. Cederbaum, Phys. Chem. Chem. Phys. pp. 4 45 (4). [] W. Vanroose, Z. Zhang, C. W. McCurdy, and T. N. Rescigno, Phys. Rev. Lett. 9, 531 (4). [1] C. R. Claydon, G. A. Segal, and H. S. Taylor, J. Chem. Phys. 5, 3387 (197).

5 [] W. B. England, B. J. Rosenberg, P. J. Fortune, and A. C. Wahl, The Journal of Chemical Physics 65, 684 (1976). [3] R. Dörner, V. Mergel, O. Jagutzki, L. Spielberger, J. Ullrich, R. Moshammer, and H. Schmidt-Böcking, Physics Reports 33, 95 (). [4] P. J. Chantry and G. J. Schulz, Phys. Rev. 156, 134 (1967). [5] M. Sizun and S. Goursaud, J. Chem. Phys. 71, 44 (1979). [6] D. J. Haxton, C. W. McCurdy, and T. N. Rescigno, Phys. Rev. A 73, 674 (6).