Advances in
ATOMIC,MOLECULAR,AND OPTICAL PHYSICS VOLUME 29
EDITORIAL, BOARD
P. R. BERMAN New York University New Yo...
14 downloads
790 Views
17MB Size
Report
This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
Report copyright / DMCA form
Advances in
ATOMIC,MOLECULAR,AND OPTICAL PHYSICS VOLUME 29
EDITORIAL, BOARD
P. R. BERMAN New York University New York, New York K. DOLDER The University of Newcastle-upon-Tyne Newcastle-upon-Tyne England
M. GAVRILA F.O.M. lnstituut voor Atoom- en Molecuulfysica Amsterdam The Netherlands
M. INOKUTI Argonne National Laboratory Argonne, Illinois S . J. SMITH Joinr Institute for Laboratory Astrophysics Boulder, Colorado
ADVANCES IN
ATOMIC, MOLECULAR, AND OPTICAL PHYSICS Edited by
Sir David Bates DEPARTMENT OF APPLIED MATHEMATICS AND THEORETICAL PHYSICS THE QUEEN'S UNIVERSITY OF BELFAST BELFAST, NORTHERN IRELAND
Benjamin Bederson DEPARTMENT OF PHYSICS NEW YORK UNIVERSITY NEW YORK, NEW YORK
VOLUME 29
@
ACADEMIC PRESS, INC.
Harcourt Brace Jovanovich, Publishers Boston San Diego New York London Sydney Tokyo Toronto '
This book is printed on acid-free paper. @ Copyright 0 1992 by Academic Press, Inc. All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopy, recording, or any information storage and retrieval system, without permission in writing from the publisher.
ACADEMIC PRESS, INC. 1250 Sixth Avenue, San Diego, CA 92101
United Kingdom Edition published by ACADEMIC PRESS LIMITED 24-28 Oval Road, London NWl 7DX
OF CONGRESS CATALOG CARDNUMBER: 65-18423 LIBRARY ISBN 0-12-003829-3 ISSN 1049-25OX
PRINTED IN THE UNITED STATES OF AMERICA 91929394 9 8 7 6 5 4 3 2 1
Contents
CONTRIBUTORS
ix
Studies of Electron Excitation of Rare-Gas Atoms into and out of Metastable Levels Using Optical and Laser Techniques Chun C . Lin and L. W. Anderson
I. 11. 111.
IV. V.
Introduction Excitation out of the Ground Level into Nonmetastable Levels Excitation out of the Ground Level into Metastable Levels Excitation out of Metastable Levels Conclusion Acknowledgments References
1 2 12 23 27 30 30
Cross Sections for Direct Multiphoton Ionization of Atoms M. V. Ammosov, N . B. Delone, M . Yu Ivanov, I . I . Bondar, and A . V. Masalov I. 11. 111.
IV. V.
VI.
Introduction Methods of Measuring the Principal Quantities That Characterize Multiphoton Ionization of Atoms The Procedure for Measuring the Quantities Needed to Find the Multiphoton Cross Sections and Its Accuracy The Results of Measuring the Multiphoton Cross Sections of Direct Atom Ionization Analytical Expression for Estimating the Multiphoton Cross Sections of Direct Atom Ionization Conclusion References
34 45 73 90
101 107 108
Collision-Induced Coherences in Optical Physics G . S . Aganval I. 11.
Introduction A General Framework for the Calculation of Nonlinear Optical Phenomena
114 116
Contents
vi
Second-OrderOptical Response and Collision-InducedCoherences Collision-InducedCoherences in Fluoresence and Ionization Spectroscopy V. Collision-InducedCoherences in Third-Order Nonlinear ResponseFour-Wave Mixing VI . Collision-InducedCoherences in Probe Absorption in the Presence of a Pump VII . Relation between Collision-InducedCoherences in Second-Order and Third-Order Responses VIII. Collision-Induced and -Enhanced Resonances in Fifth-Order Nonlinearities IX . Effect of Cross-Relaxationon Collision-Induced Resonances X. Dipole-Dipole Interaction-InducedResonances XI. Collision-Induced Resonances in Spontaneous Processes XI1. Narrowing and Enhancement of Signals Due to VelocityChanging Collisions XIII. Nonlinear Response, Collisions, and Dressed States XIV. Other Sources of Coherences Similar to Collision-InducedCoherences xv. Conclusion Acknowledgments References III. IV.
119 124
129 139 145 147 149 152 155 160 164 166 172 174 174
Muon-Catalyzed Fusion Johann Rafelski and Helga E. Rafelski I. 11.
111. IV. V.
Introduction Nuclear Fusion dr Muon Catalytic Cycle and Reaction Rates Muon Sticking Conclusion Acknowledgments References
177 181
186 199 209 21 1 212
Multiple-ElectronExcitation, Ionization, and Transfer in High-Velocity Atomic and Molecular Collisions
J . H.McGuire I.
Introduction
II. Theory III. Observations and Analysis N. Conclusion Acknowledgments Appendix: Correlation References
217 219 263 312 313 314 315
CONTENTS
vii
Cooperative Effects in Atomic Physics J . P. Connerade
I. 11.
HI. IV. V. VI . VII. VIII .
Ix.
X. XI. XI.
Introduction Many-Body Effects and the Conservation of Angular Momentum Rydberg Series Non-coulombic Potentials and the Periodic Table Giant Resonances Atomic Giant Resonances in Other Environments-Controlled Collapse and Instabilities of Valence Giant Resonances in Nuclei and in Atomic Clusters Are Giant Resonances in the d and f Sequences Atomic Plasmons? Extending Mean Fields Beyond the Hartree-Fock Scheme Can One Blow Off Complete Shells by Laser Spectroscopy? Interactions Between Giant Resonances and Rydberg SeriesIntershell and Intersubshell Couplings Conclusion References
INDEX
CONTENTS OF PREVIOUS VOLUMES
325 328 330 333 337 34 1 344 348 352 355 358 364 365 369 373
This Page Intentionally Left Blank
Contributors
Numbers in parentheses indicate the pages on which the authors’ contributions begin.
G. S. Agarwal(113), School of Physics, University of Hyderabad, Hyderabad 500 134, India M. V. Ammosov (33), General Physics Institute, Vavilov str., 38, 117942, Moscow, USSR L. W. Anderson (l), Department of Physics, University of Wisconsin, Madison, Wisconsin 53706
I. I. Bondar (33), Uzhgorod State University, Oktyabrskaya str., 54 2924000, Uzhgorod, USSR J. P. Connerade (325), Blackett Laboratory, Imperial College, London SW7 2AZ, UK
N. B. Delone (33), General Physics Institute, Vavilov str., 38, 117942, Moscow, USSR M. Yu.Ivanov (33), General Physics Institute, Vavilov str., 38, 117942, Moscow, USSR Chun C.Lin (l), Department of Physics, University of Wisconsin, Madison, Wisconsin 53706 A. V. Masalov (33), Lebedev Physical Institute, Leninsky prosp. 53, 117924, USSR J. H. McGuire (217), Department of Physics, Tulane University, New Orleans, Louisiana 701 18 Johann Rafelski (177), Department of Physics, University of Arizona, Tucson, Arizona 8572 1 Helga E. Rafelski (177), 5250 N. Foothills Drive, Tucson, Arizona 85718
This Page Intentionally Left Blank
ADVANCES IN ATOMIC, MOLECULAR. AND OFTICAL PHYSICS, VOL. 29
STUDIES OF ELECTRON EXCITATION OF RARE-GAS ATOMS INTO AND OUT OF METASTABLE LEVELS USING OPTICM AND LASER TECHNIQUES CHUN C . LIN and L. W.ANDERSON Department of Physics University of Wisconsin Madison,Wisconsin
I. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 11. Excitation out of the Ground Level into Nonmetastable Levels. . . . . . . . . 2 111. Excitation out of the Ground Level into Metastable Levels . . . . . . . . . . 12 A. Excitation into the Metastable Levels of Helium . . . . . . . . . . . . . 12 B. Excitation into the Metastable Levels of Neon . . . . . . . . . . . . . . 15 C. Excitation into the Metastable Levels of Argon . . . . . . . . . . . . . 21 D. Excitation into the Metastable Levels of Krypton. . . . . . . . . . . . . 22 E. Excitation into the Metastable Levels of Xenon . . . . . . . . . . . . . 23 IV. Excitation out of Metastable Levels . . . . . . . . . . . . . . . . . . . 23 A. Excitation out of the Metastable Levels of Helium . . . . . . . . . . . . 24 B. Excitation out of the Metastable Levels of Neon . . . . . . . . . . . . . 26 C. Excitation out of the Metastable Levels of Argon . . . . . . . . . . . . 26 D. Excitation out of the Metastable Levels of Krypton. . . . . . . . . . . . 27 V. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . 30 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
I. Introduction Electron excitation of atoms is one of the most fundamental inelastic atomic collision processes. It provides a powerful means of studying the nature of the electron-atom interaction. The body of cross section data for electron excitation out of the ground level into various excited levels for different atoms that has been accumulated over the past few decades has led to a basic understanding of the excitation processes. Electron excitation is also important for many applications in areas such as gas-discharge lasers, fluorescence lighting technology, and the dynamics of the upper atmosphere. In this chapter, we discuss a new phase of electron excitation: excitation into or out of metastable levels of the rare-gas atoms. Since metastable atoms in laboratory experiments radiate at a very slow rate compared with the total de1 Copyright 0 1992 by Academc Ress. Inc. All nghts of rcpduclion in any form reserved ISBN 0-12-003829-3
2
Lin and Anderson
struction rate, the detection of the metastable atoms is difficult and experiments to measure the electron excitation cross sections into or out of metastable levels require novel experimental techniques. These experiments are, however, very important because they allow one to test whether the physical description of electron excitation derived from the experiments on excitation out of the ground levels is also valid for excitation out of the metastable levels. Furthermore, metastable atoms and molecules play a prominent role in many natural phenomena such as atmospheric processes and gas discharges and plasmas. For instance, in gas discharges containing a rare gas at pressures above about 1 mtorr, the primary mechanism to produce the electron-ion pairs that sustain the discharge is a two-step one involving electron excitation into a metastable level followed by a second electron collision with an atom in a metastable level that results in ionizing the metastable atom. There are usually large numbers of metastable atoms present in a discharge and the processes that lead to the formation or the destruction of these metastables are important. This chapter contains discussions of experiments on the electron excitation of rare-gas atoms into or out of metastable levels. In Section I1 we discuss briefly the electron excitation out of the ground level to nonmetastable levels of rare-gas atoms as background, and then in Section I11 we discuss the electron excitation into metastable levels of rare-gas atoms. There are various methods of measuring total (integrated) cross sections or differential cross sections into metastable levels including methods based on metastable detection, using laser-induced fluorescence (LIF) or absorption, using Auger-electron or channel-electron multipliers, or using electron energy loss measurements. We discuss these various methods but with emphasis on the measurements of total cross sections, especially those measurements using optical detection because of the reliability and the relative simplicity of the measurements based on optical methods. Discussions of other recent measurements of total cross sections using nonoptical methods are also included. In some cases the differential cross section data are used to provide absolute calibration for measurements of total cross sections. Our discussion of differential cross section measurements will focus mainly on those used for calibration of total cross sections. In Section IV we discuss measurements of total cross sections out of metastable levels for rare gases. We do not include measurements of differential cross sections out of metastable levels since they were not used for calibration of total cross sections in the experiments discussed in this chapter.
II. Excitation out of the Ground Level into Nometastable Levels In a typical experiment using the optical method to measure the total excitation cross sections out of the ground level of an atom, an electron beam passes
ELECTRON EXCITATION OF RARE-GAS ATOMS
3
Collision Chamber
Monochromator
Standard Lamp FIG.1. Schematic diagram of the apparatus for measuring optical-emission cmss sections. The rotating mirror allows one to direct separately the radiation from the collision chamber and the radiation from the standard lamp to the monochromator via the same optical path.
through a gas inside a collision chamber and is collected by a Faraday cup as shown in Fig. I . Emission from the atoms excited by the electron beam emerges from a slot in the Faraday cup and passes through a window in the collision chamber. The light emitted from a segment Ax of the electron beam into a known solid angle is directed to a monochromator and the optical signal for a particular i --f j transition is detected by a photomultiplier tube (PMT). Absolute calibration obtained by comparing the PMT current due to the atomic transition with that due to a standard lamp at the same wavelength, gives the number of photons emitted per unit time and per unit length of the electron beam, N ( i + j ) . A detailed description of the experimental apparatus and procedures has been given by Filippelli et al. (1984). The gas density and electron-beam current are kept sufficiently low that one can neglect all the secondary processes that affect the excited-level population such as collisions of excited atoms with electrons or other atoms.' The optical emission cross sections for the i -+j transition, QOpt(i + j), is defined by the equation (I/e)p,Qo,, (i
j) =N i
-+
+
j),
(1)
I In some experiments secondary processes may be a complicated issue and cause considerable errors in the cross sections. See, for example, Filippelli er al. (1984).
Lin and Anderson
4
where I is the electron-beam current, e is the proton charge, and pg is the number density of the atoms in the ground level. Summing the optical-emission cross sections over all the j-levels below a given i-level gives the apparent excitation cross section for the i-level,
which is related to the steady-state number of atoms in the i-level per unit length of the beam, N ( i ) ,
( I / e)p,QaPp(i ) = N(i
(3)
),
where A(i) =
A(i+j),
(4)
j
and A( i +j ) is the Einstein A coefficient for the i +j transition. The notation < i under the summation sign covers all the j-levels that are below the i-level. Equation (3) results from Eqs. (1) and (2) since
j
N ( i + j ) = N(i)A(i + j ) .
(5)
The i-level is populated both by direct excitation from the ground level and by radiative cascade from the atoms that are excited by the electron beam to the various levels above i. To obtain the direct excitation cross section, Q d i r , the cascade contribution from all the higher levels must be subtracted from the apparent cross section, i.e., Qdir(i)
=
Qapp(i>
-
2 ksr
Qopt(k+
i).
(6)
The optical-emission cross sections are the experimentally measured quantities from which the apparent and direct excitation cross sections are deduced. According to Eqs. (1) and (3,the ratio of Qopt(i+ e) to Qopt(i+ k) is equal to the ratio of the corresponding Einstein A coefficients so that Eq. (2) can be rewritten as
If the branching ratio, A(i + e ) / A ( i ) ,is known through spectroscopic measurements or theoretical calculations, then the apparent cross section for level i can be obtained by measuring the optical cross section for only one transition of the type i +j. To determine the direct excitation cross section from Eq. (6) requires optical cross sections for all the k + i emissions, some of which are usually in the infrared or far infrared region and are difficult to measure because of the lower sensitivity of infrared detectors as compared with PMT. However, if one can measure the optical cross section for a transition of the type k + m , which
5
ELECTRON EXCITATION OF RARE-GAS ATOMS
lies in a more favorable part of the spectrum, then one can infer the k -+ i cross section from the relation Qo,,(k * i) = Q,,(k + m)[A(k + i ) / A ( k+ m ) ] .
(8)
Determination of the direct excitation cross sections for the nIS, nlP, nlD, n3S, n3P, and n3D series of He by means of the optical method have been reported by several groups, e.g., St. John et al. (1964), Moustafa Moussa ef al. (1969), van Raan et al. (1974), and Showalter and Kay (1975). Optical cross sections for the nIS * 2’P, nlP + 2’S, nlD ---* 2lP, n3S + 23P, n3P+ 23S, and n3D 4 23P transitions with n 1 3 were measured. All these lines have wavelengths suitable for PMT detection. The transition probabilities are known with a reasonable degree of accuracy so that Eq. (7) gives the apparent excitation cross sections. Likewise, Q. (8) can be used to determine cascade from the higher levels for which the k + m optical cross sections have been measured and for which the remaining cascade is then calculated. To illustrate the main features of the cross section data, in Fig. 2 we show plots of direct excitation cross sections for the 3IS, 3’P, 3ID, 33S, 33P, and 33D levels of He versus the incident electron energy, called the excitation functions, measured by St. John ef al. (1964). Each curve is separately normalized so that all curves have the same maximum height. Levels of the same spin, same L, and different n exhibit a very similar shape for the excitation function except for the different peak values of the cross sections, which decrease with increasing n, and for a slight shift in the position of the peaks due to the slight difference in
0.0
0
50
100
160
200
INCIDENT-ELECTRON ENERGY (eV) FIG. 2. Energy-dependence of the direct cross sections for electron excitation out of the ground level of helium to the 3IS, 3’P, 3ID, 33S, 33P, and 3’D levels measured by St. John er al. (1964). All curves are separately normalized to give the same peak height.
6
Lin and Anderson
the threshold excitation energies. The shape of an excitation function generally falls into one of three categories. A very broad maximum followed by a slow decline is seen for the excitation functions of the nlP levels, which are optically (electric dipole) connected to the ground level. Excitation functions for all the triplet levels exhibit a narrow peak and decrease rapidly at higher energies. The change in the total spin of the target atom in these processes might at first sight seem to require a magnetic spin-dependent interaction between the incident and atomic electrons which is much smaller than the Coulomb interaction. However, instead of a direct spin flip, the mechanism responsible for excitation of the triplet levels is that the incident electron takes up a bound excited orbital and the atomic electron with spin opposite to that of the incident electron is ejected from the atom. This mechanism requires no magnetic spin-dependent interaction but only an exchange between the incident and atomic electron. Such an electronexchange process is efficient if the incident electron is slow enough to remain in the vicinity of the target for an appreciable amount of time. As one increases the electron energy, the duration of interaction decreases, causing a drastic decline in the excitation function. Intermediate between the very broad and very narrow excitation functions is the group of excitation functions for the nlS and n'D levels which correspond to spin-allowed but dipole-forbidden optical transitions to the 1 S ground level. The three kinds of energy dependence of the cross sections are also predicted by theoretical calculations based on the Born-type approximations as illustrated in Fig. 3 where the curves are individually normalized to the same peak height. Furthermore, with the Born-Bethe approximation, 1.0
5
E
w
0.8
0.0
0
50
100
150
200
INCIDENT-ELECTRON ENERGY (eV) FIG. 3. Energy-dependence of the direct cross sections for electron excitation out of the ground level of helium to the 3'5, 3'P, and 3'D levels calculated by the Bornapproximation and to the 33S, 33P, and 33D levels by the Born-Rudge approximation using Hartree-Fock wave functions. All curves are separately normalized to give the same peak height.
ELECTRON EXCITATION OF RARE-GAS ATOMS
7
TABLE 1 PEAKVALUESOF THE DIRECTELECTRON EXCITATION CROSSSECTIONS (in l O - l * cmz) out of the Ground Level into Various Excited Levels of Helium ~
Levels 3's 3'P 31D
~
Qd 0.47 3.2 0.35
Levels 3's 33P 3'D
Q!? 0.89 0.67 0.25
it is shown that at high energies the cross sections for excitation to the n'P levels are proportional to E - I In E where E is the incident electron energy and the cross sections for excitation to niS and n i D are proportional to E - ' (Mott and Massey, 1965; Moiseiwitsch and Smith, 1968). For the triplet excitation both the Born-Ochkur (Ochkur, 1963) and Born-Rudge (Rudge, 1965) approximations predict an E -3-dependence at high energies. The overall agreement between Figs. 2 and 3 is considered satisfactory in view of the experimental uncertainty and the fact that the Born approximation is valid only at high energies. Parenthetically, we should mention that the theory of electron excitation of atoms has progressed much beyond the stage of the Born approximation. In this chapter we use the Born approximation for illustration because of its simplicity. A discussion of the current status of the theory is outside the scope of this chapter. The magnitude of the cross sections also depends on the nature of the transitions. Excitation to the n' P series corresponding to dipole-allowed transitions shows much larger peak values of the cross sections than the n l S and n i D series as exemplified by the results of St. John et al. (1964) shown in Table I. The peak values of the cross sections for the n3S and n3P series are larger than those for the n3D series, but even the former are smaller than the peak values of the cross sections for the n'P series. A simple, intuitive way to understand the behaviors of the excitation cross sections is to think of the excitation process as an absorption-like transition caused by the electromagnetic field associated with the incident electron. If we apply a multipole-type analysis to the electromagnetic field, the dipole component (k = l), which is the strongest one, is responsible for the excitation from the 1's level ( L = 0) to the n'P levels ( L = 1). Excitation to the n'D levels requires the quadrupole component ( k = 2), hence smaller cross sections. To account for the excitation into the n ' S level, one has to introduce the k = 0 component which has no direct optical analog. The essence of this simple picture can be traced to the studies of the stopping power of matter for charged particles by Bohr (1913, 1915) in which he suggested a qualitative similarity between the effect of a moving charged particle on an atom and the effect of an electromagnetic wave in producing the absorption and dispersion (Merzbacher 1984), and to the work of Purcell (1952) on the lifetime of the 2*S levels of hydrogen in
8
Lin and Anderson
an ionized atmosphere. One may wonder whether the implicit use of a projectile path is compatible with the wave nature of the incident electron. However, this simple picture can be placed on a more rigorous foundation by means of the Born approximation which gives the cross section as being proportional to the square of an integral containing the wave functions of the initial and final stages of the projectile-target system and the Coulomb interaction of the incident electron (coordinates r:r, 8 , @) with the atomic electrons (coordinates ?;:ri,ei, Qi).Upon expanding the Coulomb term using the spherical harmonics as
where r , is the greater of r and r i , and r< is the lesser, it is easy to see that the k = 0, 1, and 2 terms are responsible for excitation to the S, P, and D levels, respectively. This picture applies to collisions at high energies in which the impact duration is small compared to the classical periods of the target electrons. For slow collisions the orbital motion of the target electrons may be sufficiently altered by the projectile that modifications are needed for the arguments given in this paragraph. Nevertheless, it often turns out that this simple picture serves as a useful guide for qualitative considerations even for collisions of intermediate energies. The classifications of excitation cross sections and excitation functions of He are based on the quantum numbers of the excited states under study according to the LS coupling. This scheme can be extended to atoms that do not conform to the LS coupling. Consider the Ne atom in the excited 2p5ns configuration. The 2p5 core has angular momenta el = I and s, = 1/2, and the outer ns electrons, e2 = 0 and s2 = 112. The LS coupling gives the 'P, , 3P0,3 P l , and 3P2terms. To describe a non-LS-coupling case, one starts with the LS eigenstates and allows for the deviation from the LS coupling by mixing the various LS eigenstates of the same J since the total angular momentum is rigorously a good quantum number even for the intermediate coupling cases. If we neglect mixing of configurations, the wave functions, $. of the 2p5ns levels can be expanded using the LS eigenfunctions, 4, as = 4 ( 2 p 5 n s , ' P I ) + P4(2p5ns, 3PI), JlZ = P4(2p5ns, ' P I ) - 4 ( 2 p 5 n s , 3PI),
$1
$3
=
$,
=
(10)
4(2p5ns, 'Po), 4(2p5ns, 3P2).
Since 3P0 and 3P2are the only members with the particular value of J within the 2pSns configuration, neither member mixes with any other members. In Paschen's notation the levels corresponding to G I , J13, and 14, for the 2p53s configuration are called I s2, 1s,, 1s3, and 1ss , respectively. For the 2p54s con-
ELECTRON EXCITATION OF RARE-GAS ATOMS
9
figurations the designation of 2s2 through 2s, are used and so on for higher configurations. Since the ns3 and ns5 levels are purely triplet, we expect their direct excitation functions to show a sharp peak near the onset and a steep decline at high energy based on the results of He discussed earlier. The direct excitation functions for ns, and ns4, on account of the mixed singlet and triplet characters of their wave functions, are expected to be a combination of a broad-peak and a narrow-peak component. Unless one of the mixing coefficients is exceptionally small, the large broad peak dominates so that the excitation functions of the nsz and ns4 levels have the shape characteristic of the dipole-allowed excited states, i.e., a very broad peak with a slow decline (E - I In E dependence) at high energies. The 2p5np configuration can be analyzed in the same way. The ten levels of the 2p5np configuration can be considered as superpositions of the ten ISo, 3 S l , IPI, 3P0,3PI, 3 P ~ ID2, . 3D1, 3D2,3D3LS eigenstates with mixing only among members of the same J. Here the 3D3term, being the only member with J = 3, remains unmixed, whereas all the other nine levels have both singlet and triplet characters. Thus, only the level with J = 3 has a sharply peaked excitation function. The other nine levels have excitation functions characteristic of spin-allowed and dipole-forbidden excited levels, i.e., a peak of intermediate width and (l/E)-dependence at high energies. Notice the difference between this and the very broad peak and E-I In E dependence for the J = 1 levels of the 2p5ns configuration because transitions from the ground level (2p6 IS,) of Ne to the J = 1 levels of the 2p53s configuration (Is, and 1s4) are dipole-allowed whereas transitions from the ground level to all levels of the 2p5np configuration are dipole-forbidden. The apparent excitation functions for the four levels of the 2p55sconfiguration (3s in the Paschen notation) and for 9 of the 10 levels of the 2p53pconfiguration (2p in the Paschen notation) measured by Sharpton et al. (1970) are shown in Figs. 4 and 5 , respectively. The apparent excitation functions include the cascade contributions, and the deviation of the shape of the apparent excitation functions from the shape of the direct excitation functions varies from one level to another. Nevertheless, these graphs show qualitatively the three kinds of excitation functions: the very broad ones for the optically allowed levels of the 2p55s configurations (3s, and 3s,); the narrow peaks for the 2p53p, J = 3 (2p9) level, 2p55s, J = 2 (3s5) level, and 2p55s, J = 0 (3s3) level characteristic of spin-forbidden levels; and the intermediate shape for the remaining eight levels of the 2p53p configuration (spin-allowed but dipole-forbidden). The 2p3 curve is not included in Fig. 5 because the apparent excitation cross sections of the 2p3 level have a very large cascade component so that the shape of the apparent excitation function differs significantly from that of the direct excitation function. The 2p5nd configurations can be treated in a similar way (Sharpton et al., 1970) and are not detailed here. The different shapes of the excitation functions for different levels deduced from these analyses are in full agreement with experiments.
Lin and Anderson
10
fTym
"iNm
3c 0
v)
J-1
J-1
0
100
0
200 0
100
200
Electron Energy, eV FIG. 4, Apparent excitation functions for the four levels of the 2ps5s configuration of Ne. All curves are separately normalized to give the same peak height. The levels are labeled by the Paschen notation and the value of J .
U
mm J=2
J=2
40 8o 12Ol6O
40 8o 12Ol6O
40 8o 120160200
Electron Energy, eV FIG. 5 . Apparent excitation functions for nine levels of the 2p53pconfigurationof Ne. All curves are separately normalized to give the same peak height. The levels are labeled by the Paschen notation and the value of J .
A very interesting trend in the magnitude of the excitation cross sections of Ne has been observed experimentally (Sharpton et al., 1970). Within the 2p53p configuration the levels with even values of J generally have larger cross sections than those with odd J , whereas the reverse is true for the 2ps4d configuration. This observation can be explained by the multiple picture described earlier. Excitation from the 2p6 IS, ground level to a level in the 2p53p configuration entails a 2p +-3p transition, or t? = 1 + t? = 1. This can be accomplished by
ELECTRON EXCITATION OF RARE-GAS ATOMS
11
means of the k = 0 and k = 2 components of the multipole interaction. However, the final levels are further characterized by J that ranges from 0 through 3. The k = 0 and k = 2 components just mentioned are effective for transitions from the ground level (J = 0) to a J = 0 level and a J = 2 level, respectively, of the 2p53p configuration, but do not match the parity requirement for transitions from J = 0 to J = 1 or J = 3. While transitions to the o d d 4 levels can take place through higher-order effects (or exchange for J = 3), their cross sections are expected to be much smaller than the cross sections for the even-.! levels. For excitation from the 2p6 configuration to the 2p54d configuration, the p .+d transition (A1 = 1) limits us to the k = 1 and k = 3 components which are compatible with transitions from J = 0 to J = 1 and J = 3, but not with the J = 0 0, 2, 4 transitions. Therefore, in this case we expect to find larger cross sections for the odd4 levels. A rigorous treatment based on the Born approximation and group-theoretical arguments is given by Sharpton et al. (1970) for Ne and by Ballou et al. (1973) for Ar. Fabrikant et al. (1988) and Teubner ef al. (1985) have expressed concern about the reliability of describing the ls3(,PO)and ~ S ~ (levels ~ P as ~ purely ) triplet levels in electron-excitationexperiments. In principle the ,Po (or ,P2) level could be mixed with a singlet level of the same J from a different configuration of the same parity. The usual configuration interaction that arises from the r i k Coulomb interaction term in the Hamiltonian, however, does not produce triplet-singlet mixing because r;' is diagonal in the total spin quantum number. To spoil the pure triplet character of the Is, and Is, levels of Ne, one must resort to mixing of the 2p53s configuration with a higher configuration through interactions of electron spins with orbital motion. Considering the relative magnitude of the spin-orbit-type interaction versus the energy spacings between the 2p53s and the higher configurations, it is clear that such mixings must be very small. This is confirmed experimentally by the gJ factor which is sensitive to the singlet admixture of these nearly pure triplet levels except for the case of J = 0 where there is no Zeeman splitting. The measured value of the gJ factor for the Is, level of Ne is 1.503 (Moore, 1949), which is very close to the theoretical value of 1.501 for the pure ,P2 level. Likewise, the measured gJ factors for the corresponding Is( 3P2)levels of Ar, Kr, or Xe indicate that these levels are also pure LS levels. Of course, one must distinguish the triplet-singlet mixing in the 1s3(3P0)or ls,t3P2) levels from the triplet-singlet mixing in the Is, and Is, levels. In the latter case the mixing is quite appreciable because it occurs between LS eigenstates ()P, and ' P I ) within the same configuration ( 2 ~ ~ 3 sHowever, ). the fact that the 1s2and 1s4 levels are not pure LS levels does not mean that other levels of Ne cannot be pure LS levels. As explained earlier, the Is, and Is, levels have the unique J values within the 2p53s configuration so that they do not mix with any LS eigenstates of the same configuration and therefore can be described as pure LS levels to a high degree of accuracy. In studying electron excitation of the rare-gas atoms, recognition of the pure LS levels among the majority of non-
-
12
Lin and Anderson
LS levels is a key step in the progress of understanding the cross section data from a unified viewpoint in terms of the atomic structure. In addition to measurements of total-excitation cross sections using the optical method, differential cross sections for electron excitation are of interest. Differential cross sections can be measured by detecting the electrons that have lost the exact energy necessary to excite a given atomic level and that are scattered at an angle 8.These experiments provide information on the angular distribution of the scattered electrons that cannot be obtained using the straightforward optical method to measure total cross sections. The differential excitation cross section multiplied by sin 8 and integrated over all angles 8 is equal to the totalexcitation cross section. In order to obtain accurate integrated cross sections, it is necessary to have differential cross sections that are measured over a wide range of angles, and if the differential cross section is strongly peaked in the forward direction, the scattering angle must be measured very accurately. It is usually preferable to measure the total cross sections using the optical method.
III. Excitation out of the Ground Level into Metastable Levels The measurement of total electron excitation cross sections into metastable levels has been studied for all the rare-gas atoms (He, Ne, Ar, Kr, and Xe) using various techniques. For He the methods utilizing absorption out of the metastable levels, sensitized fluorescence, or time of flight have been used. For Ne, there are extensive measurements including measurements using optical and laser techniques as well as nonoptical methods. Measurements for Ar, Kr, and Xe are somewhat more limited. A. EXCITATION INTO THE METASTABLE LEVELS OF HELIUM The ground level of He is (Is*) ‘ S o and the two lowest excited levels, (ls2s)2’S and ( l s 2 ~ ) 2 ~are S , both metastable. The LS coupling gives an excellent description for the low-lying levels of He. Since the metastable levels do not radiate, they are difficult to detect. An early optical experiment (Woudenberg and Milatz, 1941) to measure the apparent excitation cross section of the 23S level of He used optical absorption out of the 23S level. They produced the Z3Smetastables by an electron beam in a He gas and used the absorption out of the z3S level to determine the rate of production of He atoms into the 23S level. They carefully studied the effects of pressure and of other experimental factors. They determined the relative values of the apparent cross section and found that QaPp(2)S) has a nmow peak near an electron energy of 25 eV.
ELECTRON EXCITATION OF RARE-GAS ATOMS
13
Another optical method for detecting the metastable levels of He is sensitized fluorescence. An excitation transfer collision of a He atom in the 23S level produces a Cd+ ion in an excited (4d95s2)2D,, level by the reaction He(23S)
+ Cd('S) + He(1lS) + Cd+(
'D5/2)
+ e-.
(11)
The excited Cd+( 'D5J ion is detected by its emission of radiation of wavelength 441.6 nm. Bogdanova and Marusin (1975) used the radiation from the Cd+ ions to detect He atoms in the 2 ) s level. The 441.6-nm radiation emitted by the Cd+( 2D5/2) ions is measured first in the absence of He and then in the presence of He. The rate equations for the formation and destruction of He(23S) and Cd+( 2D5/2)levels can be solved in the steady state to give the following expression for the apparent electron excitation cross section for He(23S) atoms,
where Z is the intensity of radiation emitted by Cd+(2D5,2)in the absence of He, AZ is the change in the intensity of radiation emitted when He is present, Qapp[Cd+ ( 'DY2)] is the apparent cross section for production of Cd+(2DJ12) by electron impact, QTis the excitation transfer cross section for the reaction given by Eq. ( l l ) , (QTv) is the thermal average of QT times the relative He-Cd velocity, nHeis the ground-level He density, and ~ ( 2 ~iss )the reciprocal of the steady-state destruction rate of the He(23S) atoms by all mechanisms. Equation (12) is based on the assumptions that AZ is entirely due to the transfer process in Eq. (11) and that the rate of this transfer process is much smaller than [ ~ ( 2 ~ S ) ] -Inl . addition to measurements of A M and Qa,[Cd+ ( *Ds,2)],one must have accurate values of QTand ~ ( 2 ~ins order ) to determine the metastable cross section. Thus, a comprehensive knowledge of the excitation transfer mechanisms is required to obtain reliable results using the method of sensitized fluorescence. Bogdanova and Marusin (1975) used the value of QT as roughly equal to 10-l5 cm2 estimated by Ivanov et al. (1972), and they measured the lifetime of He(23S) from the duration of the afterglow to obtain ~ ( 2 ~ sIn) .a similar manner Bogdanova and Marusin ( 1969) have measured the apparent electron excitation cross section of He(2IS) using the sensitized fluorescence of Ne(3s2). They found a peak value for Qa,(2'S) of 1.6 x lo-'' cm2 at an electron energy of 35 eV, and a value for Qapp(21S)of 1.3 X lo-'* cm at an electron energy of 45 eV. Their results are in serious disagreement with other measurements. An interesting optical experiment using laser-induced fluorescence has been performed by Zetner et al. (1986) to study the electron excitation of the 2IS level of He for energies near threshold. The LIF technique for determining metastable level cross sections was developed by Phillips er al. (1981a, 1985) in an experiment to study the excitation of Ne metastables and is discussed in Section 1II.B dealing with Ne. Zetner et al. (1986) have observed resonances in the 2 ' s excitation cross section.
14
Lin and Anderson
In addition to the experiments using optical methods, there are a number of nonoptical experiments that give electron excitation cross sections into the metastable levels. The sum of the apparent cross section for the electron excitation for He into either the 23S or 2IS level, i.e., Qam(z3S) Q,pp(2LS),has been measured by Mason and Newell (1987). They produced an atomic beam of ground level He. A very monochromatic electron beam was incident on the atomic He beam. The He atoms were deflected by the electron-excitation collision. The deflected He atoms in either the P S or 2'5 level were detected by a channel-electron multiplier. The electron-beam current was measured using a Faraday cup. The electron beam was modulated so that the excitedatom-velocity distribution could be determined using time-of-flight techniques. This made it possible to discriminate against a signal from uv photons emitted by He atoms excited to radiating levels. Mason and Newell have measured earn@%) Qam(21S)for electron energies from thresholds up to 140 eV. They obtained absolute cross sections by normalizing their results to the earlier results of Borst (1974) at 20-eV incident electron energy. The peak value of Qam(z3S) Qam(2IS)is 6 X 10-l8 cm2 at an electron energy of 25 eV. The measurements of Mason and Newell agree well with earlier measurements and seem to give reliable values for Qam(2%) Qa,(2'S). Earlier measurements of Qaw(Z3S) Qam(21S)have been carried out both by Borst (1974) using Auger electron detection of the metastables and by Lloyd et al. ( 1 972) using channel-electron-multiplier detection of the metastables. Borst ( 1974) obtained absolute cross sections by calibrating at low energy using the trapped-electron method, whereas the measurements of Lloyd et al. yielded only relative cross sections for electron energies of 19.8-200 eV. Overall the results of these experiments are consistent with the more recent measurements of Mason and Newell (1987). In addition to the experiments discussed, there have been a large number of experiments to measure the near-threshold behavior of the total apparent cross section for the production of He atoms in either the 23Sor 2IS metastable levels. As examples we cite experiments by Brunt et al. (1977), Buckman et al. (1983a), Pichanick and Simpson (1968), and Johnston and Burrow (1983) for QaPp(2 ' S ) with emphainvestigating the near-thresholdbehavior of Q,(23S) sis on studying the resonances that occur in these cross sections. Because our primary interest is in the overall cross sections as a function of the energy rather than the near-threshold behavior, we do not discuss these experiments further. Several experiments have been carried out to determine separately QaW(2S) or Qapp(2'S).Cermak (1966) has measured separately the relative values of Qapp(21S) and QaW(2'S)from near threshold to 60 eV. The 23S or 2IS levels were detected using Penning ionization of Ar, and the kinetic energy of the electrons ejected in the Penning ionization was measured to detect separately the 23S or 2's levels. Holt and Krotkov (1966) have measured Qapp(z3S) or QaPp(2'S) from 19.0-23.2 eV. They detected the metastables using Auger electron emis-
+
+
+
+
+
+
ELECTRON EXCITATION OF RARE-GAS ATOMS
15
sion from a metal surface, and used electric-fieldquenching to eliminate the 2ISlevel atoms from an atomic He beam. They have also used an inhomogeneous magnetic field to permit separate measurements for excitation into the m = 1, 0, - 1 states of the 2% level. Dugan et al. (1967) have measured separately relative values of Qa,,(23S) or QaP,(2'S) using an inhomogeneous magnetic field to separate the 2% and 2's atoms. Their measurements were carried out for electron energies of 25- 135 eV. Their results show a narrow peak in QaPp(2%)near 25 eV and a broad maximum in Qapp(21S)near 100 eV. All these measurements appear to be carefully done and offer useful results. Another nonoptical method used to measure electron excitation cross sections is the inelastic scattering of electrons with adequate energy resolution to determine the level in He that is excited. These experiments yielded direct differential cross sections. One must integrate over all scattering angles 8 the differential cross section times sin 8 to obtain a total cross section. Chamberlain et al. (1970) and Trajmar (1973) have used inelastic electron scattering to obtain differential direct cross sections. The sum of total cross sections for 23S and 2 ' s obtained using Trajmar's differential cross sections lies somewhat below the total apparent cross sections obtained by Mason and Newell (1987). This may be due to cascade effects in the experiments of Mason and Newell (1987) that are not present in the measurements of Trajmar (1973). In summary, for the 23S and 2IS metastable levels of He, there are reliable Q,,(21S). The early measurements using nonoptical techniques of Q,,,(23S) optical measurements of Woudenberg and Milatz (1941) appear to give reasonable relative values of Qapp(23S).The LIF measurements of Zetner et al. (1986) near threshold are reliable and of high quality. There are some other nonoptical measurements of Q,pp(23S)and Qa,p(21S)that appear reasonable. The total direct cross sections obtained by integrating the differential cross sections obtained by inelastic electron scattering also appear reasonable. The totai apparent cross sections obtained by Bogdanova and Marusin (1975) using sensitized fluorescence do not agree well with the results of others and appear to suffer from some systematic uncertainty. It would be desirable to have further experiments using optical methods to obtain separately absolute direct cross sections for the 23S and 2 ' s levels over a wide range of electron energies. There have been only a few measurements of the cross sections for the separate levels and they have yielded only apparent cross sections except for the inelastic-electron-scattering measurements. Optical measurements are especially needed to give both the apparent and direct cross sections.
+
B. EXCITATION INTO THE METASTABLE LEVELS OF NEON
The ground level of Ne is 2p6 IS,. The lowest excited configuration is 2p53s which gives rise to four levels designated as Isz, Is,, Is,, and Iss in the Paschen
16
Lin and Anderson
notation. As explained in Section 11, the Is, and Is, levels are accurately described by the LS-coupling 3P, and 3P2 terms, respectively. The Is, and Is, levels are superposition of 'PI and 3P1as expressed in Eqs. (10) with Is, being approximately 93% 3P1plus 7% 'PI and the opposite mixture for Is,. The Is, and Is, levels are metastable because a transition from either level to any lower levels is forbidden from consideration of the J quantum number and parity. Both the Is, and Is, levels decay to the ground level by electric-dipole radiation and the lifetime of the Is, level is about 13 times as long as the lifetime of the Is, level. An early optical experiment for measuring the excitation cross section of the Is, level of Ne was carried out by Milatz and Ornstein (1935). In their experiment the density of 1s5 metastable atoms produced by electron impact was determined by monitoring the absorption of a beam of 640.2-nm light by the Is, atoms through the Is, + 2p9 transition. They reported the energy-dependence of the apparent excitation cross section of the Is, level but without an absolute calibration. They observed a rapid decrease of cross section with increasing energy. They used a low Ne pressure (-0.02 torr) but without the sophisticated detection apparatus available today. Hadeishi (1962) used a similar method and operated in the pressure range of 0.4- 1.7 torr where secondary processes may occur. His measured cross sections for the Is, level depend on the pressure and are more than 10 times the cross sections of others. Aside from the difference in absolute magnitude, the energy-dependence of Hadeishi's cross sections is also quite different from that of others. More recently Mityureva and Penkin (1983) also used absorption to monitor the production of Ne atoms in the Is, and Is, levels and measured the apparent cross sections of the Is, and Is, levels in a manner similar to that used by Milatz and Ornstein (1935). They subtracted the cascade cross sections to obtain direct excitation cross sections to the Is, and Is, levels. They found the peak value of the direct cross section for the Is, level is 3.4 X 10-l8cm2at an electron energy of about 23 eV. A new and innovative method to obtain the electron excitation cross sections for the Ne metastable levels using laser-induced fluorescence has been introduced by Phillips et al. (1981a,b, 1985). A dye laser operating at the appropriate wavelength was used to excite the metastable atoms to a higher level that radiates. The LIF was used to detect the metastables. If the rate of excitation out of the metastable level by the laser is much greater than the total decay rate out of the metastable level by all other mechanisms and if the dye-laser intensity and frequency distribution are such that one can saturate the entire Doppler absorption line width out of the metastable level, then the rate of photons produced by LIF is equal to the rate of formation of metastables by direct electron excitation plus cascades. In this case the absolute apparent electron excitation cross section can be obtained from the absolute LIF decay rate, the electron current, and the target density. For Ne the lowest configuration above the 2p53s is the 2p53p configuration.
ELECTRON EXCITATION OF RARE-GAS ATOMS
17
This configuration gives rise to 10 energy levels denoted as the 2p,, . . . ,2p10 levels in the Paschen notation. The wavelengths of many of the optical transitions between the levels in the 2p53p and the 2p53s configurations are in the visible region and are well suited for LIF experiments using cw dye lasers. In order to make concrete the use of LIF for the measurement of the cross section for electron excitation into a metastable level, we discuss the excitation into the Is,(’P2) level of Ne. The relevant processes involved are shown in Fig. 6. The Is, level was populated both by direct electron excitation and by cascade from higher levels. A cw laser operating with a wavelength of 588.2 nm excited atoms in the Is, metastable level up to the 2p2 level. The 2p2 level has a configuration 2ps3p and an electronic angular momentum of J = 1. The 2p2 level decays by radiation to the four 1s levels and the branching ratios for this radiation are known (Sharpton ef al. 1970). The radiation at 659.9-nm wavelength corresponding to the 2p2 4 Is, transition (LIF) was observed. The 2pz level was populated in two different ways. The first method, by which the 2pz level was populated, was the electron excitation of this level by both direct and cascade processes, and the second method was the laser absorption from the Is, metastable level. These two processes for populating the 2pz level are independent and consequently the 2p2 -9 Is, radiation results from the same two processes. The 2p2+ Isz LIF produced by 2p2 atoms excited from the Is, level by the laser was separated from the 2p2 --.* Is, radiation due to electron excitation
Electron Impact Excltatlon
Ground Level FIG. 6. Diagram showing the electron excitation, laser absorption, and laser-induced fluorescence processes involved in the use of the LIF technique for measuring cross sections of electron excitation out of the ground level into a metastable level of neon.
18
Lin and Anderson
by modulating the laser beam. A collimated electron beam passed through a stainless-steel chamber filled with about one mtorr of Ne exciting some of the Ne atoms in the ground level into the metastable Is, level. A cw dye-laser beam with a wavelength of 588.2 nm passed through the chamber intersecting the electron beam at right angles. The optical emission emitted from the region of the electron beam crossed by the laser beam was observed along an axis that was perpendicular to both the electron-beam axis and the laser-beam axis. The laser beam was on-off modulated by a rotating mechanical chopper. The optical emission was analyzed using a monochromator and then detected using a PMT. The output current of the PMT was the input for a lock-in detector, the output of which was the y-input of an x y recorder. The x-input to the recorder was the voltage used to accelerate the electron beam. The plot obtained is directly proportional to the electron excitation function. The absolute calibration of the electron excitation cross section posed an important problem. If the intensity and frequency distribution of the dye-laser beam were such that the entire Doppler linewidth of the Is, + 2p2 transition was saturated, then the number of photons emitted per second in LIF would be equal to the rate of electron excitation of atoms into the Is, level. In this situation a measurement of the absolute intensity of the LIF coupled together with the solidangle of the detector would enable one to obtain the calibration of the absolute apparent cross section. This calibration scheme has not yet been carried out because of the high laser intensity required. An alternative method for calibrating the apparent electron excitation cross section is as follows. As explained in Section 11, the Is, is a pure triplet level. The direct electron excitation cross section of a pure triplet level is expected to peak at an energy only a little above threshold and then decrease rapidly with increasing energy. Since most of the levels that cascade into Is, (such as 2p2 and 2p4) have a mixed singlet-triplet character, the excitation cross sections of the majority of these cascading levels decrease with increasing energy rather mildly in comparison with the Is, level. Thus, for the Is, level, the direct electron excitation cross section is expected to be very small in comparison with the cascade cross sections at electron energies of 60 eV or higher. The total apparent electron excitation cross section of the Is, level at energies of 60 eV or higher is therefore very nearly equal to the sum of the cascade cross sections, which are the optical emission cross sections for transitions into Is, from the various higher levels and have been measured absolutely by Sharpton et al. (1970). Thus, the absolute calibration of the apparent cross section for the electron excitation of the metastable 1s, level was obtained by setting the Is, cross section equal to the sum of the cascade cross sections at 90-eV incident electron energy. The direct cross section at energies below 90 eV was obtained by subtracting the sum of the cascade cross sections from the apparent cross section. This calibration procedure of course means that the direct cross section was not obtained for energies of 90 eV or higher.
ELECTRON EXCITATION OF RARE-GAS ATOMS
19
In addition to measurements of the cross sections for electron excitation into the Is, level of Ne, measurements of the cross sections for the electron excitation into the Is, ( ,Po) level of Ne have been made using the same method as for the Is, level (Phillips er af., 1981b). By analyzing the polarization of the LIF as a function of the laser polarization, absolute excitation cross sections for separate Zeeman states (m,-states) of the metastable Is, level have been measured at an impact energy of 18.2 eV (Phelps et af., 1983). At this impact energy the cross sections for exciting the various m,-states are found to be equal to one another within experimental error. In addition to the measurements of the electron excitation cross sections over a broad energy range using LIF, there has been an important extension of the LIF to measurements of the electron excitation cross sections of Ne for nearthreshold energies by Zetner er al. (1986). They have used the LIF method to study resonance structure in cross sections for the electron excitation of the Is, and Is, levels, and they also have measured the cross sections for electron excitation into a single m,-state of the ls, level. The results they obtained show resonance structure in both the Is, and Is, cross sections. There are also several nonoptical measurements of the electron excitation cross sections for the metastable Is, and Is, levels of Ne. Mason and Newell (1987) reported measurements on the total cross section for the sum of the apparent cross sections for the electron excitation into the Is, or Is, levels of Ne for electron energies of 16.2-140 eV, i.e., Qapp(ls3)+ QapP(ls5),using the same method described in Section 1II.A. They normalized their cross sections to those of Teubner et at. (1985) at 26 eV for absolute calibration. Earlier measurements of the sum of the apparent excitation cross sections of the two metastable levels include the work of Dorrestein (1942). He used electron bombardment of a Ne gas to form metastables. The metastables drifted to a detector where Auger electron emission was used to detect them. The difficulties of this method include determining the efficiency for the ejection of secondary electrons from the metal surface by an incident metastable atom and separating the contributions to the total secondary electron emission of the metastable atoms and the ultraviolet photons. He used modulation techniques to overcome these problems. Teubner et af. (1985) also have measured the sum of the apparent excitation cross sections of the Is, and Is, levels using a method somewhat similar to that of Dorrestein (1942) except that they used a time-of-flight technique to eliminate the photon signal. Teubner er al. raised the point of whether the Is, and ls, levels can be accurately described as purely triplet levels. We have discussed this issue in Section II. The works of Teubner er al. (1985) and of Mason and Newel1 (1987) seem carefully done and they obtain cross sections that are reasonably consistent with each other. In addition, relative values of Qapp(ls3)+ Qapp(ls,)have been reported by Theuws et al. (1982). Dunning et af. (1975b) have measured the ratio of the population of the Is,
20
Lin and Anderson
and Is, levels produced by electron impact by an optical method. They used a laser to selectively remove either the Is, or Is, atoms from a beam of metastable atoms and detected the metastable atoms by secondary electron emission from a metal surface. Their measurements give the Is,: Is, ratio at electron energies 35, 60, 80, and 100 eV as, respectively, 6.9 ? 1.4, 5.8 0.6, 5.1 2 0.4, and 5.0 2 0.4. Register et al. (1984) have reported differential cross sections for excitation of 16 features in the electron energy-loss spectrum of Ne. By integrating their measured differential cross sections times sin 8 over all scattering angles, they obtain the total cross section for direct excitation. Their integrated results agree reasonably well with those of other researchers. There are a number of papers using nonoptical methods that measure the sum of the electron excitation cross sections for the 1s levels with primary interest in near-threshold resonance structures. Papers reporting such measurements include Pichanick and Simpson (1968), Buckman et al. (1983b), Brunt et al. (1976), Schaper and Scheibner (1969), and Johnston and BUKOW (1981). In summary, optical experiments have been used to measure the individual apparent cross sections into either the Is, or Is, levels, i.e., Qaw(ls3)or Qa,( lsJ, whereas the nonoptical experimentals have been used to measure the sum of cross sections into the ls, or Is, levels, i.e., Qaw(ls3)+ Qa,(ls5). The measurements of Milatz and Ornstein (1935) show a much more rapid decrease in Qa,(ls5) than more recent measurements. The results of Mityureva and Penkin (1983) indicate that above about 30 eV, Q(ls5)is smaller than Q(ls,). This is in sharp disagreement with the results of Phillips et al. (1985) and Dunning et al. (1975b). In particular, Phillips et al. (1985) find that the ratio of the direct excitation cross section for Is, to Is, is equal to the 5 : 1 ratio of the statistical weight of these two levels within experimental uncertainty and that even the ratio of the apparent excitation cross section for Is, and Is, is fairly close to 5 : 1. The results of Phillips et af. (1985) provide individual cross sections measured over a wide energy range and with reasonable values for Qapp(ls3),Qdk(ls3), Qaw(ls5),and Qd,,(ls,). The sum of the cross sections Qapp(ls3)+ Qapp(ls5) from Phillips et al. (1985) can be compared to the results obtained using nonoptical techniques to measure total apparent cross sections. The measurements of Mason and Newell (1987), Teubner et al. (1985), Dorrestein (1942), and Phillips et al. (1981a,b 1985) all place the peak of the sum of the apparent cross sections at an electron energy of about 26 eV and all have a peak value for the sum of the apparent cross sections of about 3.5 X cm2. However, the dependence of Qw(ls3) + Qapp(ls5)on the electron energy at high energies is somewhat different for these papers. The apparent cross sections of Dorrestein and of Phillips et al. agree well and the apparent cross sections of Mason and Newell and of Teubner et al. agree well but the apparent cross sections of Dorrestein and of Phillips et al. decrease from the peak somewhat less steeply at higher
*
ELECTRON EXCITATION OF RARE-GAS ATOMS
21
energies than the apparent cross sections of Mason and Newell and of Teubner et al. The differential cross sections of Register et al. (1984) for the Is, and Is, levels times the sin 8 have been integrated over all scattering angles 8 to provide total direct cross sections. The integrated total direct cross sections obtained by Register et al. (1984) agree well with the direct cross sections obtained by Phillips et al. except for electron energies above 60 eV. The disagreement above 60 eV is expected since Phillips et al. took Qdir(1s3)and Qdir(ls5)to be zero at energies above 90 eV.
c. EXCITATION INTO THE METASTABLE LEVELS OF ARGON The ground-level configuration of Ar is 3p6 and the configuration of the lowest excited levels of Ar is 3p54s. The 3P2 and 3P1levels that arise from the 3p54s configuration are virtually pure LS levels and are metastable. Mityureva and Smirnov (1985, 1986) in a pair of papers have measured the apparent cross sections for the production of the ,P2 or ,Po metastable levels. They produced the Ar metastables using electron bombardment in a low-pressure gas. The number density of metastable atoms is detected using absorption out of the metastable level. They have also measured at 22 eV the cascade contribution to the cross section. They found that the peak values of the apparent cross sections for the 3P2 and ,Po levels are, respectively, 3.4 X 10-l7 cm2 and 0.9 x l O - I 7 cm2. The peak values of both cross sections occur at an incident electron energy of about 20 eV. They also found that the peak values of the direct cross sections for the ,P2 and ,Po levels are 4 X 10-l8 cm2 and 8 X lO-I9 cm2,respectively. The sum of the apparent electron excitation cross sections into the meta3P0), has been measured stable 3P2or 3P0levels of Ar, i.e., Qapp(3Pz) + Qapp( using Auger electron emission to detect the Ar metastables by Borst (1974), who obtained absolute cross sections. The sum of the apparent cross sections Qapp( 3P2) Qapp(3P0) has been measured using a channel electron multiplier to detect the metastables by Mason and Newell (1987), Theuws et al. (1982), and Lloyd et al. (1972). Both Theuws et al. (1982) and Lloyd et al. (1972) obtained only relative cross sections. Mason and Newell (1987) calibrated their measurements at 22 eV to the results of Borst (1974) to obtain absolute cross sections. The sums of the apparent cross sections Qa,(-’P2) + Qa,,> 1, the atom should absorb the so-called threshold number of radiation quanta, K , given by the relation
where {n} denotes the integer part of x . If the condition (3) is fulfilled, i.e., if the atom absorbs the threshold number K of radiation quanta, the energy E = Khw - I < hw is distributed between the positive ion and the electron and, since the latter is much lighter, practically all of it transforms into electron kinetic energy. In the opposite case (i.e., where y 1 for radiation in the near ultraviolet, visible, and near-infrared frequency ranges. The process of multiphoton ionization is called direct if the energy of any number K' < K of absorbed quanta does not coincide with the energy of any transition in the atomic spectrum from the ground to an excited state allowed by selection rules for multiphoton transitions (Delone and Krainov, 1985). If the E, = 5 X lo6 V cm-I, one can strength of the radiation field is E > 1 we come to the case y >
1 and E
I , i.e., the ionization process is always of multiphoton character. In the multiphoton limit ( y >> l ) , the direct process can be clearly isolated only if the field strength is not very high, so that the change in the energy of bound states, 8&(E),owing to the dynamic Stark effect (Delone and Krainov, 1985), is much smaller than the difference between the energies of the neighboring states. It is evident that the field strength E at which the condition
44
Ammosov, Delone, and tvanov
FIG. 5. Typical scheme of an experiment where multiphoton cross sections are measured: I , laser source; 2, attenuator made of two polarizers; 3, neutral optical filters; 4, optical wedge; 5, focusing lens and a lens identical to it; 6, windows of the vacuum chamber; 7, vacuum chamber; 8, system of electrodes of the time-flight mass spectrometer; 9, field-free flight gap; 10, ion detector; 11, vacuum gauge; 12, calorimeter; 13, ion-signal detecting and processing system; 14, photodetector; 15, wideband oscilloscope; 16, microscope; 17 camera.
SE(@ lo) at K = 2 to 6. For this reason, for practical measurements one can neglect the difference between (lg Ni) and (lg Ni), .
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
55
And so, for the rest of the chapter, as main relations for the measured quantities we shall use relations (17)-(19), (24), (25), (31), (32)-(3% and (38).
D. MEASUREMENT OF THE POWER OF NONLINEARITY OF THE
IONIZATION PROCESS
The measurements of the power of nonlinearity K play an important role in the study of multiphoton atom ionization. The agreement between the experimentally observed value of K and its theoretical value {I/hw 1) serves as a basis for identifying the ionization process, indicates that this process is direct, and thereby makes it possible to measure other characteristics of the ionization process. Only that region of intensity values where the observed power of nonlinearity is an integer and agrees with the calculated value can be used for measuring the ionization probability and multiphoton cross section with the absolute method. When one does not know the value of the multiphoton cross section, the power of nonlinearity K is measured by observing the growth of the ion yield with radiation intensity. It is convenient to analyze the dependence of the ion yield on radiation intensity by using a plot with a double logarithmic scale, where this dependence has the form of a straight line with the slope K:
+
For each laser pulse we have a corresponding point on this diagram. By looking at the set of experimental points obtained for different values of intensity one is able to select the range of intensity and ion-signal values where the dependence In Ni on In Fo is a straight line, and find the value of K from the slope of this dependence. The broader the range of the values of radiation intensity and ion signals within which the straight-line dependence is observed, the more accurately K can be measured. The value of K is obtained by approximating the data in this region by a straight line. This is a necessary procedure since the experimental points cluster around a straight line with a certain dispersion. The latter depends on the apparatus error with which one measures Ni and F,,, on the specific procedure used to measure the intensity, and on the mode composition of the laser radiation used in the measurements. The resulting error with which the value of K is measured is determined by the dispersion of experimental points and by the range of intensity values within which the measurements are made and the number of points therein. An evident requirement to the error with which one measures K is the condition 6 K S 0.5. In this case there can be no mistake about what integer number the measured value corresponds to. However, more accurate measurements are also important since they indicate that the direct pro-
56
Ammosov, Delone, and Ivanov
cess of ionization takes place. A favorable circumstance for making high-accuracy measurements is the fact that the measured value of K is independent of the absolutization of the ion-signal and radiation-intensity values. The value of the nonlinearity power K can be found from the experimental values x, = lg F, and y, = lg(Ni), (where index m = 1, 2 , . . . , M labels the pairs of measured values of Ni and F , ) by using the least-square method. The specific formulae used to estimate K within this method may be different and depend on which one of the two errors (Sx or 6 y ) determines the spread of experimental points. If the latter is mainly due to the fluctuations of intensity (K SX >> Sy, the case where single-frequency laser radiation is used, to be discussed shortly), the value of K is estimated by minimizing the sum Ern[x, - ( y , - b)/K]' (by regression of x to y). Then, one has
If however the ion-signal fluctuations are dominating (Sy >> KSx), K is estimated by minimizing the sum Ern( y , - Kx, - b)* (by regression of y to x), and
T K =
XmYm
c
-
(T
xi -
x m ) ( C
Yrn)/M -
(C
=
Kmin.
(41)
Xrn)'lM
Though the values K,, and Kminare close to each other, they do not coincide. The difference between K,,, and Km,"should be taken into account when the error with which K is found is less then K,, - K,. (to be discussed shortly). The error with which K is estimated within the least-square method depends upon the scatter of experimental points about a straight line, and is calculated by using the correlation coefficient of experimental data:
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
57
The value of R characterizes the quality of the experimental data obtained. The smaller the spread of points, the closer R is to unity and the smaller the error of K estimation is:
Moreover, K,, - K d n = K m ( l - R 2 ) . As one can see, at a small number of experimental points, M - 2 < 1 / ( 1 - R 2 ) . the values K,,, and Kmincoincide with each other within the error 6 K . However, in a situation more likely for experiments on multiphoton ionization of atoms by laser radiation, where M - 2 > 1/(1 - R 2 ) , the difference between K,,, and Kminexceeds the error SK, and one has to select a correct expression for K . The correlation coefficient can be predicted basing the data on relative errors with which the radiation intensity and ion signals are measured: 6xllg e = 6F/F and 6y/lg e = 6Ni/Ni,respectively: 6x2
6Y2
t
In the case where the experimental points are evenly distributed throughout the and Ay = y,,, - ymin, in Eq. (44) one can replace the range Ax = x,,, - xmin sums by their approximate values and obtain 1 - R 2 = 12
I (E)’ (z)’I +
L
-1
=
(AY)2
[K26x2+ 6 ~ 2 1 ,
(45)
where we have taken into account that Ax = AylK. Equation (45) enables one not only to make prognoses about the error in the measurement of K in a future experiment, but also to check how realistic are the estimates for 6x and Sy after the measurements have been done. Furthermore, Eq. (45) explicitly contains the contributions of 6x and Sy to the resulting scatter of experimental points. Let us now turn to the estimation of the error in the measured value of K with Eqs. (43) and (45). As was already mentioned, the result of data approximation according to Eqs. (40) and (41) is independent of the units one uses to measure the intensity and the ion signals, i.e., it does not depend upon the error in absolutization of F o and Ni (as well as of n o ) . Thus, Sx and 6y in Eq. (44) are determined only by the errors in Fo and Ni in relative units. The latter include the random errors of the measuring apparatus and the errors that are due to fluctuations of the measured quantities. For 6x we have (Sx/lg e)’ =
($)2
+
(?)*+ if)’+ (F)‘,
(46a)
Ammosov, Delone, and lvanov
58
whereas in 6y2, in addition to the ion-signal-measurement error given by Eq. (21), one should also take into account the errors in n o , V, , and TK: (6y/lg e)2 =
(F)2+ (2)'+ (2)2 + (?)*.
(4%)
If the measurement procedure is such that for each laser pulse one measures the radiation energy Q and the parameters of the space-time distribution of S and T, and, on their basis, calculates Fo (the dependence N,(F,,)), 6 x will contain only the apparatus errors of Q,S, and T. For such a measurement procedure one can expect 6xAg e "- 0.05 to 0.1. If, however, in the process of measurement one controls the radiation energy only (the dependence N,(Q)), 6 x will contain the errors 6S/S and STIT due to technical fluctuations of these quantities. In this case, one can expect Sxllg e = 0.1 to 0.2. The error 6y essentially depends upon what sort of a laser (single-frequency or multifrequency one) is used in the measurements. In the case of a singlefrequency laser, Sy is mainly due to the apparatus error 6N,IN,, to which the technical fluctations 6noln0, 6vK/vK, and 6T,/T, are added. Under such conditions, one can expect 6yllg e = 0.1. For a multifrequency laser 6N,lN, also contains fluctuations associated with nonreproducibility of the fine temporal structure of radiation. Depending upon the number of generated modes N and on the power of nonlinearity of the ionization process, the multifrequency fluctuations 6N,lN, fall within the range of 0.01 (N = lo3, K = 2) to 2 (N = 10, K = 10) (Masalov and Todirashku, 1980) and in realistic experimental situations are dominating. In the case of multifrequency lasers, one should expect 6yllg e = 0.1 to 2. Both the range of variation of the ion signals, Ay = l g ( N y / N y ) , and the associated range of intensities, Ax -- AylK, acceptable for measuring the power of nonlinearity are limited. From below they are limited by the sensitivity of the ion detector; when making estimates we shall suppose that the sensitivity is maximum and N y = 1. From the preceding, the range in question is limited by saturation of the ionization process: N y = noVK= noh31K04( 0 is the angle of radiation focusing into the interaction volume.) For an atom beam no cm and 8 = 1120, we get N Y = 1 O l o ~ r n - ~and , taking h = 0.5 X = lo3 and Ay = In lo3 = 3. When the ionization of atoms occurs in a gas, the atom concentration can be much higher, and the range of measurements can be made wider. In this case, however, one encounters difficulties that are due to the fact that the apparatus that records the ion signals has a limited dynamic range. = lo3. For this reason, in our further estimations we shall take N-INLet us now present the three resulting estimates:
(1) When one studies the dependence N,(Fo) with a single-freqency laser and 6xllg e = 0.05 to 0.1, 6yl lg e = 0.1, K26xZ> 6y2, and K = K,,,, one has
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
I - R 2 K 2 (0.0005 + 0.002) 6K -- K2 (0.02 t 0.04)lFM.
59 (474 (47b)
The accuracy 6K < 0.5 can be reached with several tens of laser pulses up to K = 10. (2) When one studies the dependence Ni(Q)with a single-frequency laser and 6xllg e = 0.1 to 0.2, 6yllg e -0.1, K 2 6 x 2>> 6y2, and K = K,,, ,one has 1 - RZ = Kz (0.002 + 0.008) 6 K = Kz (0.04 f 0.09).
(47c) (474
Here, for a “reasonable” number of laser pulses, M S 100, the accuracy SK 6 0.5 is guaranteed to be reached within the range up to K = 10. (3) When one uses the radiation of a multifrequency laser, the values of K26x2 and 6y2 can be comparable to each other, and the estimates for 1 - R 2 and 6K strongly depend upon the value of K and the number of modes N. For instance, at K = 5 and N = lo’, and Syllg e = 0.23 < KGxIlg e 0.5, we have
-
I - R 2 0.05 6 K = l.l/-, which ensures the necessary accuracy after several tens of laser pulses. On the other hand, at K = 7, N = 100, Sy/lg e = 0.9, and KGxllg e = 0.7 we have 1 -
R 2 0.26 21
6K = 3.5/<M, which ensures the necessary accuracy at M = 100. We note that the measurement results for K obtained with multifrequency lasers are independent of the magnitude of the statistical factor of radiation. One should note two circumstances that are common to both types of measurements just considered, but are more important for those made with multifrequency lasers. First of all, in cases where the ion-yield fluctuations are high (6Ni/Ni = I), the correct measurement procedure is to average the logarithms of the ion signals rather than the signals themselves. The analysis shows that in such cases the ionsignal distribution is far from normal. For this reason, estimates of the average value are quite sensitive to the loss of signals that happened to be outside the dynamic range of the recording apparatus, and these estimates converge to the true average with increase of the number of measurements slower than l / e . On the contrary, distribution of the ion-signal logarithms is closer to normal, and estimates of the average values are only weakly sensitive to the loss of extreme values and converge to the true value in the “correct” manner. The second circumstance concerns the estimates of the power of nonlinearity based on the experimental data in those cases where K 2 W is comparable to
60
Ammosov, Delone, and lvanov
6y2 (as it is in the case we just considered, where K = 7 and N = 100, so K6x/lg e = 0.7 while 6y/lg e = 0.9). In this case neither Eq. (40) for K,,, nor Eq. (41) for Kmingive correct (nondisplaced) estimates for K. Moreover, K,, - Kminmay be >1 (in our example, K,, - Kmin= 1.8), and then the possibility of reaching high accuracy in K measurements is not used. The account of the comparable contributions of errors 6x and 6y to the spread of experimental points leads to the following formula for K:
This formula is valid if Eq. (44)holds. Without the analysis of the contributions of 6 x and 6y to 1 - R2it is impossible to specify the value of K more precisely inside the interval between Kminand K,,, . To summarize the preceding considerations concerning the measurement of the power of nonlinearity of the ionization proces, we can assert that the existing pulse lasers enable one to make measurements within the range K = 2 to 10 with an error no higher than 6 K G 0.5 for a moderate number of laser pulses, M = 30 to 100. The best accuracy can be reached with single-frequency lasers, when the intensity of radiation is measured (in relative units) for each pulse. Such measurements require rather complicated apparatus, so the intensitymeasurement procedure is justified in those cases where it is necessary to measure K with very high accuracy, or where the high-radiation-energy and lowradiation-energy regimes of the laser are noticeably different (Lompre er al., 1982). On the other hand, if one does not measure the radiation intensity for each laser pulse, the experiment becomes much simpler, while the loss in accuracy is insignificant. By comparing the measurement results obtained with single-frequency and multifrequency lasers, one can see that the former provide higher accuracy. For single-frequency laser radiation, the measurement error for K is mainly determined by the relative error with which one measures the radiation intensity and by the range of observed ion signals. In this case the error 6K is proportional to K2 and at small K (G5) may reach one or two tenths. When one uses multifrequency laser radiation, the fluctuations of the ion yield become greater. This, in its turn, increases the measurement error and makes the data-processing procedure more complicated: it becomes necessary to estimate the contributions of the errors in intensity and in the ion yield to the spread of experimental points, and to select a formula for the nondisplaced estimate of K. Let us note that by increasing the number of measurements M (the number of laser pulses) one cannot increase the accuracy of measurements to any noticeable extent, since owing to
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
61
the square-root-like dependence, the variation of M has little effect on 6 K . The most efficient way to increase the accuracy of measurements made with multifrequency lasers is to increase the number of generated modes: with increase of the number of modes the multifrequency fluctations of ion signals become smaller. And though multifrequency lasers, when used for studying the processes of multiphoton atom ionization, give a worse resulting accuracy than do the singlefrequency lasers, their application is justified by the simplicity of the laserresonator construction and, consequently, by reliability of operation, and, also, by the stability of radiation energy and of the space-time distribution parameters in a series of pulses.
E. MEASUREMENT OF THE IONIZATION PROBABILITY The ionization probability per unit time, w = aKFK, characterizes the rate of K-photon ionization of an atom in a field with fixed intensity. The value of the ionization probability is used in the calculation of ion-concentration dynamics n,(f):& ( f ) = (no - n,)w. Under variation of radiation intensity the ionization probability changes as w' = w(F'/F)K.
(49)
The observation of ion signals in the process of multiphoton atom ionization by pulsed laser radiation enables us to measure the ionization probability. The value obtained will correspond to one of the intensity values that have been realized within the interaction volume during the laser pulse. Since the greatest contribution to the ion yield comes from the intensity at the maximum of the space-time distribution, the experimental measurements are usually referred to that value of intensity. The measurements are based upon the following relation between the integral (over the pulse) ion signal Ni and the characteristics of radiation: Ni = aKFi$oTKVK= nOw(F,)TKVK. For the ionization probability per unit time this gives
If the measurements are made with multifrequency lasers, it is necessary to take into account the statistical factor of radiation, which makes allowance for the time fluctuations of intensity within a pulse:
Ammosov, Delone, and Ivanov
62
In practice one should use the latter formula, since this is what corresponds to the experimental situation, where with observation of ionization in a series of laser pulses the logarithms of the ion signals are averaged. Equation (51) is presented to give a complete picture, and it can only be used in those cases where the number of modes generated by the multifrequency laser is sufficiently large and g, and lWKcoincide within the ion-yield measurement error. It should be emphasized that Eqs. (50)-(52) are correct only far away from saturation, where the ion yield is proportional to the Kth power of radiation intensity. This means that the measurements of the ionization probability should be preceded by the measurements of the ion yield versus radiation intensity dependence. Any point on the straight-line dependence of In Ni on In Fo with the slope K is good for measuring the ionization probability w. It is then evident that in the case of multifrequency lasers the averaging procedure gives (lg Ni), so one should use Eq. (52) and not (51). As was already mentioned, the ionization probability w calculated according to one of the Eqs. (50)-(52) corresponds to the value of intensity at the maximum of the space-time distribution: Fo = Q/(huST). The error with which one can measure w with the method described previously is determined not only by the random error of all the quantities that enter Eqs. (50)-(52) but also by the errors of absolutization of the neutral atom concentration no and ion yield N i values: (!?)2
=
+
(%)2
+
($)2
(T)’ + p$;bs + (?)*
(!3)2 +
(53)
abr
In the case of multifrequency lasers, where the fluctuations 6 N i / N , ,due to the nonreproducibility of the fine temporal structure of pulses, are large, the contribution of multifrequency fluctuations of the ion yield to the resulting error can be suppressed compared to (6NilNi), by averaging over one or two tens of laser pulses. One should keep in mind that the preceding estimate of the error in w expressed through the errors of no, Ni , T, , and V, absolutization is not a complete characteristic of the measured value of w:there is also a “hidden” error due to the error in absolutization of intensity values for which w was measured. This hidden error becomes apparent when it becomes necessary to find the value w ’ of the ionization probability at another value of intensity F’ basing on the measured value w :w ’ = w(FA/F0),. The error in w’ obtained this way now contains the error in intensity 6 F o . Thus, in order to give a complete characteristic to the result of measuring the probability of multiphoton atom ionization, one should know not only w and 6w but also Fo and SFo .
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
63
Owing to the dependence on radiation intensity, the ionization probability w is not a universal characteristic of the process of multiphoton ionization. Such a universal characteristic is the K-photon cross section a K For . this reason, we do not consider here the methods of measuring the multiphoton ionization probability in more detail. In conclusion, let us mention one important case where it is necessary to know the ionization probability: the case where one compares the processes that have different powers of nonlinearity (i.e., that result from absorption of different numbers of photons). Let us recall that, since the dimension of multiphoton cross sections depends upon the power of nonlinearity of the process, one cannot compare these cross sections for those processes that have different Ks. In this case one should compare the probability of the process measured at fixed radiation intensity.
F. MEASUREMENT OF MULTIPHOTON CROSS SECTIONS OF DIRECT ATOMIONIZATION The cross section of K-photon ionization of an atom is the proportionality coefficient between the ionization probability per unit time and the Kth power of radiation intensity: LYK
= W/FK.
(54)
The magnitude of aKdepends upon the frequency of laser radiation and upon its polarization, but does not depend upon the radiation intensity. Measuring these cross sections is one of the most important aims of experimental studies of the process of multiphoton atom ionization in a light field. By comparing the experimental values of aKwith its theoretical values, one is able to check the correctness of the used theoretical models of an atom and of its interaction with radiation. When describing the methods used to measure the ionization cross sections, we shall pay principal attention to the resulting accuracy of these measurements. Let us mention in advance that the modern methods enable one to measure aK with an accuracy of the order of aK = loA'", where 6A = 1. This accuracy may prove to be insufficient to enable one to establish a difference between the rough estimates of aKand the results of detailed calculations. This is why it is important to improve the accuracy of the experimental measurements of aK. There are two substantially different methods of measuring the atom ionization cross sections: the absolute and the relative. They differ in the set of quantities measured in the experiment and, consequently, in the collection of the measuring apparatus, in the methods of data processing, and, in the end, in the resulting
64
Ammosov, Delone, and Ivanov
accuracy of measuring a K .For instance, among other quantities, within the absolute method one has to measure the ion yield Ni and the concentration of neutral atoms no in absolute units, whereas within the relative method it is sufficient to measure Ni in arbitrary relative units, while no does not have to be measured at all. (This is what gives these methods their names.) For this reason, as concerns the accuracy of measurements, the potentialities of the relative method are greater.
G . THEABSOLUTE METHODOF MEASURING MULTIPHOTON CROSS SECTIONS The absolute method of measuring the cross sections of K-photon atom ionization is based upon the interrelation between the ion yield and the radiation intensity far from saturation: CW, =
Ni/n,FgTKVK.
(55)
Equation (55) can be used to measure the ionization cross sections with radiation of single-frequency lasers. When one uses multifrequency lasers, in the formula for the cross section one should take into account the statistical properties of radiation: (YK
= (N,)/(no&'#o)KTKVK)
or
(56)
CXK
= 10'" "l'-'~/(no(Fo)'TKvK).
(57)
Since the absolute method of measuring the cross sections is applicable when the ion yield observed is far from saturation, the measurements of the cross section according to the absolute method should be preceded by the measurements of the ion yield versus radiation intensity dependence. Any point of this dependence from the range where N, F t is fitted for calculating the cross section according to Eqs. (55)-(57). Since the experimental data on the dependence of N, on Fo are analyzed on a diagram with a double logarithmic scale, in the case of measurements made with multifrequency lasers one should calculate aKaccording to Eq.(57). The error in measuring aKwith the absolute method is mainly due to the errors of absolutization of intensity, of the number of ions, and of the neutral atom concentration:
-
In those cases where multifrequency lasers are used, in this relation one should also take into account the error in the value of the statistical factor. The quantity
65
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
66 takes into account the error that is due to technical and random apparatus fluctuations 6Fo and 6 N , : (6Mg e)2 = (6Fo/Fo)2+ (6Ni/Ni)2.These fluctuations manifest themselves in the spread of experimental points on the dependence of lg N ion lg Fa. For this reason, the error 66 can be directly estimated from the spread of experimental points x, = (Ig F,,), and y, = (lg Ni), (rn = 1,2, . . . , M))about the line y = Kx + 6:
The error 66 can be made substantially smaller if a, is calculated, not from a single experimental point from the region of the nonsaturated Ig Ni versus lg Fa dependence, but by using all the experimental values of x, and y, corresponding to the straight-line part of the dependence of Ig N , on lg Fo. In this case, for single-frequency laser radiation we have A
5
Ig
=
[6 Ig e
-
lg(noT~V~)],
(60)
while for multifrequency radiation A = 6 lg e - Ig(n,T,V,)
- H,.
(61)
The value of b is obtained here by averaging over all the experimental values: 6 = (l/M)(Zm y, - K Emx,), and then the error IS^)^ in (59) becomes M times smaller and amounts to = ( I / M ( M - I)) X Z, (y, - Kx, - 6)2. The described procedure of experimental data processing makes the contribution of 66 to 6 A practically zero. Then, the maximum accuracy with which one can measure the multiphoton ionization cross section with the absolute method is limited by the errors in absolutization of radiation intensity, of the number of ions, and of the neutral atom concentration, and the error in intensity is multiplied by the factor K. Taking (6FolFo),, = (IjNi/Ni)abs = (6q,/nO)& = 0.3, we obtain the following estimate for the maximum accuracy of measuring the cross sections: 6A = 0.4343
X
0 . 3 q m =0.13V‘m.
(62)
According to this estimate, the accuracy with which one can measure aK with the absolute method can be within one decimal order up to K = 6 or 7. However, the practical results prove to be somewhat worse. (See Section IV.)
H. THERELATIVE METHODOF MEASURING MULTIPHOTON CROSS SECTIONS The relative method of measuring the cross sections of K-photon ionization of atoms makes use of the dependence of the ion yield upon radiation intensity in the region of saturation, where the product aKFtTK is close to unity, and the ion yield becomes comparable to the number of atoms in the effective volume:
66
Ammosov, Delone, and Ivanov
Ni = n,V, . The fact that in the saturation region the quantity (YKF~TK has definite values is used in the relative method and allows one not to measure the neutral atom concentration, while the values of the ion yield can be measured in arbitrary relative units. This makes the experimental measurements much simpler, and the accuracy with which one can measure the cross sections in this case is limited mainly by the error in absolutization of the radiation intensity value. This is why, as of 1990, the multiphoton cross sections are measured, as a rule, with the relative method. The relative method was introduced into practice by Delone et af. (197 1) and has at least three modifications that differ in the way the experimental data are processed.
1. The Two-Term Series Method In this variant of the relative method one analyzes the experimental data on the lg Ni versus Ig Fodependence in the region of low saturation, where the deviation of experimental points from a straight-line dependence becomes noticeable. In this region, the approximate analytical dependence of lg N, on lg Fo has the simple form of a two-term series Ig Ni = lg N'f - (Y~F:T,(V~~/~VK)I~ e,
(63)
which makes the experimental data processing much simpler. (Here Ig Nf is the straight line extrapolated from the nonsaturated region into the region of saturation. See Fig. 10.) The two-term series method can be used to analyze the experimental data on the lg N , versus lg Fo dependence both in the nonsaturated region and in the region where the saturation begins. The experimental points from the nonsaturated region are approximated by a straight line that is then extrapolated into the 1 h Ni
,Wi"
CROSS SECTIONS OF DIRECT MULTlPHOTON IONIZATION OF ATOMS
67
region of high intensities in order to find the value of lg N T . After that, one selects the experimental point with the smallest possible intensity whose deviation from lg NT is within two or three ion-signal measurement errors. From the deviation of this point from the nonsaturated dependence A = lg NT - lg Ni one calculates the value of the cross section:
With experimental points that deviate from Ig NT by less than two or three errors in 6 N , / N i one gets lower accuracy, whereas the use of points with greater deviation leads to a systematic error in the value of the cross section since such points lie outside the region where the two-term series is valid. When one uses the radiation of a multifrequency laser, the cross section is calculated according to the formula
In this case the statistical factor that corresponds to the ion-signal logarithm averaging procedure is g, and not HK . The error with which one can measure the multiphoton cross section with a two-term series method is mainly determined by the error in absolutization of radiation intensity values:
The term (lg e/A)(6iVi/Ni)is the relative error with which the quantity A = lg NT - Ig N , is measured, in which only the spread SN, is taken into account, while the spread 6NT is considered to be insignificant since the values of lg NT are found from a large number of experimental points. The quantity (lg e / A ) ( G N i / N i can ) be regarded as the error with which the experimental points are approximated by the theoretical curve. According to the approximation recipe proposed, it amounts to (lg e/A)(6Ni/Ni) 1/(2 + 3). One can see that the contribution of this term is not small. The low accuracy of data approximation by a two-term series is one of the weak sides of this method. Its advantages are the following: 2-
(1) It is enough to have a minimum amount of data for the beginning of saturation. (2) Data processing is simple.
68
Ammosov, Delone, and Ivanov
(3) It can be used in the case of an arbitrary spatial distribution of radiation intensity. (4) It can be used in the case where the measurements are made with radiation of multifrequency lasers.
Thus, in order to find the values of multiphoton cross sections with a two-term series one has to measure the following quantities: N , (in relative units), F, TK , and V,/VK. In the case of multifrequency lasers, one should also measure g K . The range of intensity values where the ionization process is to be observed should cover the nonsaturated region and the beginning of the saturation region. The accuracy with which the cross sections are measured is limited by the error of radiation intensity absolutization increased K times, and by the error in approximation of (lg e/A)(GNilNi). 2 . The Intersection-Point Method This method makes use of the experimental data on the ion yield in the region of deep saturation, where the In Ni versus In Fo dependence is a straight line with the slope 3/2. The straight line with slope Kin the nonsaturated region intersects the straight line with slope 3/2 in the region of deep saturation at the point where the product CUKF~TK has the value aKthat depends on K and on the shape of the spatial distribution of radiation intensity at the focus of the lens (Fig. 11). Having found the intersection point and the value of intensity F8 at this point from the experimental data on the ion yield lg N i as a function of lg Fo , one can easily calculate the cross section according to the formula
'
K-
1g Ni
~.-~3/2
Ni-F,K
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
69
If the radiation of a multifrequency laser is used, one should calculate the cross section from the relation
The intersection-pointmethod can be used only in the case where the experimental data on the lg N i versus lg F,,dependence exhibit the slope K in the region without saturation and the slope 3/2 in the deep-saturation region. For this reason, before using this method, one should measure the degrees of nonlinearity of the ionization process both in the region of small intensities and in the region of deep saturation. If from these experimental data one can find intensity ranges where the required slopes are observed, the intersection point x* = Ig F$ can be found by approximating the data within these ranges by straight lines according to the least-square method: x* = b2 - b ,
K - (3/2)'
where b, is the parameter of the line y = Kx + b, far from saturation, b, = ( l / M I )X, ( y , - Kx,), M I is the number of corresponding experimental points, b2 is the parameter of the line y = 3x12 + b, in the region of deep saturation, and b2 = ( 1 / M 2 )C, ( y ; - 3xA/2) for M 2 experimental points. Within this approximation, it is not necessary to use the experimental values of the slopes. It is enough to take the correct integer K and 312. The error with which the intersection point is found can be expressed through errors Sb, and 6b2:
We have presented here the formulae for those errors Sb, and Sb, that come from concrete experimental data. When making prognoses about the error with which ionization cross sections can be measured in an experiment that is being planned, the errors 6b, and Sb, can be estimated by using the following expressions: Sb: = (Syg Sb: = (Sy:,,
+ K2Sx2)/M1,
+ (3/2)'6x2)/M2,
(70d) (70e)
70
Ammosov, Delone, and Ivanov
where Sxllg e = SFo/Fois the expected measurement error for Fo (due to technical and random apparatus fluctuations) in relative units, 6yKllg e = SN,/Ni is the expected ion-signal measurement error in the region without saturation (also in relative units), and 6yW,/lge = 6 N i / N ,is the same error in the region of deep saturation. When the experimental data are such that in each range one has several tens of experimental points, one can expect that the measurement errors 6b, and Sb2 will be at the level of several hundredths, which gives a measurement error for intensity at the intersection point SF%/F8 smaller than the error of Fo absolutization. The resulting accuracy with which the multiphoton cross section can be measured with the intersection-point method amounts to
Thus, KSF$/F%plays the role of the error with which the experimental points are approximated by the theoretical curve and, together with the error of intensity absolutization, limits the accuracy of the cross section measurements. A shortcoming of the intersection-point method is the fact that the values of uKhave been calculated only for the light beam with a Gaussian profile and one does not know the sensitivity of the values of aK to the variation of the spatial distribution of radiation. The advantages of this method are: (1) the relative simplicity of experimental data processing and (2) its applicability in the case where the measurements are made with radiation of multifrequency lasers. Thus, in order to measure the multiphoton cross sections with the intersectionpoint method one should have a laser with a Gaussian spatial distribution and the recording apparatus should be able to measure N , (in relative units), F, , and TK (and also HKin the case of multifrequency lasers). The range of radiation intensities where the ionization process is to be observed should cover both the region without saturation and the deep-saturation region. The accuracy with which the cross section can be measured is limited by the error of radiation intensity absolutization and by the error in intensity at the intersection point increased K times.
3. The Method of Complete Comparison This method is based on comparing the experimental data on the ion yield in the region of saturation with the dependence N , ( F , ) calculated according to Eq. (14) for a concrete spatial distribution of laser radiation intensity. Within this method no preference is given to the data from one or another saturation region. Instead,
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
71
one makes the comparison by using the whole set of experimental points. The method of complete comparison has been introduced into practice by Cervenan and Isenor (1974) and Cervenan et al. (1975). These papers have stimulated the numerical calculations of the Ni( F O )dependence for radiation with a Gaussian intensity profile in the focusing region (Boulassier, 1976). The ideological side of comparison of the calculation results with experimental data and of finding the value of the cross section presents no problem. Since in the calculations, the is used as an independent variable, a comparison with the quantity CYKF~TK experimental data enables one to find the value of intensity Fc for which CYKFETK= 1 . (We have taken unity only to be definite.) The cross section of the ionization process is calculated in this case according to the formula
It is difficult to use the method of complete comparison when one uses the radiation of multifrequency lasers, since the theoretical dependencies N i ( ( F ) ) with averaging over the multifrequency fluctuations are not known. Only when the number of modes generated by the laser is sufficiently high, so that the multifrequency fluctuations are suppressed, the dependence N i ( ( F ) )is the same as it . this is for single-frequency radiation if only one replaces F t by 1 0 H ~ ( F ) K(In limit gK and 10". coincide.) Here the method of complete comparison can be used, and the cross section of the ionization process is calculated according to the formula
It is convenient to compare the experimental data with the calculated dependence on a diagram with a double logarithmic scale: x = lg F0 for the experimental data, x' = (l/K)lg(a,F{Z',) for the calculation data. Fc is found as the value of intensity that coincides with the value x' = 0 given by the theoretical dependence at optimal fitting of data. When the theoretical curve y = f ( x ) is fitted to the experimental points by using the least-square method, one encounters the same problems as when the power of nonlinearity of the ionization process is measured. One should decide which of the two deviations (in y or in x ) should be minimized by comparing the concrete errors 6 x and 6y. When the radiation of single-frequency lasers is used, and the ion-yield fluctuations are relatively small, one should expect that the error Sx, due to the technical and random apparatus fluctuations of radiation intensity, will be dominating: K26x2 > 6yz in the region without saturation in the deep-saturation region. In this case, in order and (3/2)26x2> to find Fc within the least-square method, one should minimize the sum
Ammosov, Delone, and Ivanov
72
Z, (x, - f- ( Y,))~,where f-,( y,) is the dependence inverse to f ( x ) . The error with which the scales along the x axis are compared is given by the relation 6x6 =
(2
lg e ) , = 6 x 2 ( z
+
F)I(K
- (3/2))2.
(74)
Here, when estimating ax,, the experimental points have been divided into two groups: M I is the number of points with x, < xc and M2 is the number of points with x, > x,. It is assumed that for the first group of points the slope of the theoretical curve is close to K, while for the second group of points it is close to 3/2. One can see that at a given number of laser pulses, the best accuracy is reached when the numbers of points in the first and second groups are proportional to the corresponding slopes: M I:M , = K : 312. In this case the measurement error for intensity Fc falls off with increase of the number of pulses as U C M : (6F,/F,) = ( 6 x l f l ) ( K 3/2)/(K - 3/2)lg e. At M I + M 2 = 100, one can expect that the comparison error will be noticeably less than the error of intensity absolutization. If, in the case where the error 6x is dominating, one finds Fc by minimizing the sum Z, (y, - f(x,)),, the resulting value xc will be smaller by the amount =26x2/Ax. (Ax is the range of intensity variation in decimal orders.) Though smaller than 6x, this displacement may nonetheless be greater than the comparison error axc and thus is undesirable. The opposite case, where the dominating error in the spread of experimental data is 6y, will not be considered here. Such a case could be realized if one uses the radiation of a multifrequency laser with a small number of generated modes, when the multifrequency ion-yield fluctuations are high: 6y/lg e = 6Ni/Ni 1. However, as was already mentioned, the method of complete comparison cannot be applied in this case owing to the absence of theoretical dependencies Ni((F)). The resulting accuracy with which one is able to measure the multiphoton cross section by complete comparison is
+
The dominating role here i s played by the error in intensity absolutization. The absence of theoretical calculations that could describe the ion yield versus radiation intensity dependence for an arbitrary spatial intensity distribution and that could take into account the multifrequency fluctuations of the ion signals, is a shortcoming of this method. Its advantages are: ( 1 ) its applicability to the analysis of data in any part of the saturation region and (2) the potentially high accuracy of fitting the calculated curve to the experimental data.
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
73
Thus, in order to measure the multiphoton cross sections with the method of complete comparison, it is necessary to have the radiation of a single-frequency laser (or, at least, of a multifrequency laser with a sufficiently large number of generated modes) with a Gaussian profile of the spatial distribution. The measuring apparatus should enable one to record N,(in relative units), Fo, and TK. The range of intensity variation should include some part of the saturation region. And the accuracy with which the cross sections can be measured is limited by the error in radiation intensity absolutization increased K times. The choice of the method for measuring the multiphoton cross sections is mainly determined by the characteristics of the laser radiation used to observe the ionization process:
(1) by the profile of the spatial distribution of its intensity in the focusing region, (2) by the composition of longitudinal modes, and (3) by the highest intensity at the maximum of the space-time distribution. For instance, it is difficult to use the intersection-point method and the method of complete comparison in those cases where the profile of the spatial distribution of laser radiation intensity is other than Gaussian, since there are no theoretical dependencies of the ion yield on intensity. So in such cases the two-term series method is more preferable. The method of complete comparison cannot be used if one has to work with a multifrequency laser that generates a small number of longitudinal modes since here, also, there are no theoretical dependencies. Finally, the intersection-point method cannot be used in those cases where the intensity of the laser radiation in the focusing region is not high enough to enable one to observe deep saturation of the ion signals.
III. The Procedure for Measuring the Quantities Needed to Find the Multiphoton Cross Sections and Its Accuracy The schematic diagram of the setup for measuring the cross sections of direct multiphoton atom ionization is depicted in Fig. 5 . The radiation produced by a pulse laser is focused into the vacuum chamber filled with the gas of examined atoms under low pressure. As a target, one may also use an atom beam. For producing direct multiphoton atom ionization, the lasers with pulse duration of the order of nanoseconds or shorter are necessary to have high intensity of radiation: from lo2’ photons/cm2 sec for atoms of alkali metals to photons/ cm2 sec for noble-gas atoms (i.e., of the order of lo8 W/cm2 and lOI4 W/cm2, respectively, for the energy of a photon hw = 1.2 eV). The ions produced in the region of laser radiation focusing are extracted by a constant (or pulsed) electric
74
Ammosov, Delone, and Ivanov
field into the flight gap, where they are separated according to their masses and come into the detector. The measured quantities are the laser radiation intensity F in the region of interaction with atoms; the initial atom density no; and the number of ion-produced N,.(See Eq. (II).) The quantities no and N ican be measured directly, but there is no direct method of measuring the intensity F of the laser radiation used for multiphoton ionization. Measuring the laser radiation intensity in absolute units is one of the most important procedures in experiments where multiphoton cross sections are measured. First of all, the error of intensity absolutization gives the largest contribution to the resulting error in the value of the multiphoton cross section since intensity has to be raised to the Kth power. Second, for pulse lasers (and only lasers of this kind can be used for observing multiphoton ionization) the measuring of intensity is a special case, since the space-time distribution of F is essentially nonhomogeneous. Within a laser pulse, at different points of the interaction volume one has different values of intensity. So the measurement of intensity comes to measuring a large number of values corresponding to different moments of time within a pulse and to different points within the interaction volume: F ( t , x , y, z). For measuring the multiphoton cross sections one has to use the radiation of lasers that in the regime of a single lowest transversal TEMm mode generate one or many longitudinal modes (i.e., single- or multifrequency radiation, respectively). Only in this case can one separate the temporal ( t ) and spatial ( x , y, z ) variables, and the intensity distribution F ( t , x , y, z) near the focal plane of the focusing lens can be measured with required accuracy. If the laser generates many transversal modes, the spatial and time variables cannot be separated. As the radiation pulse develops and the resonator quality grows, the modes with lower transversal indices are radiated first, while those with higher indices are radiated later on. When many transversal modes are generated, the intensity distribution along the front of the wave is no longer smooth: it has local regions with very different intensity. For radiation of a single-frequency laser, the space-time intensity distribution in the interaction region can be presented using Eq. (13):
where F , is the value of intensity at the maximum of F(t, x , y, z); Q is the total energy in the pulse; T = J +(r) dt is the pulse duration; S = J $ ( x , y, z,) dx dy is the area of the cross section in the plane of the maximum intensity value F,,; and + ( t ) and $ ( x , y, z ) are dimensionless distributions whose form is smooth and close to Gaussian, and which are normalized to their maximum value. In this case the measured quantities are the energy Q and the distributions + ( t ) and $(x, y. z). Having measured these quantities experimentally, one is then able to
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
75
calculate all the parameters that enter Eqs. (55), (64), (67), and (72) for multiphoton cross sections: T, S, F o ( F 8 , F C ) , T , = $ + " ( f ) dr, and V , = $ $"(x, y, Z) dx dy dz, VXIVK. A single-frequency laser is the most convenient one for measuring the multiphoton cross sections in the respect that the time distribution of its radiation, +(r), has no small-scale oscillations and can be measured reliably by standard methods (as is discussed shortly). In practice, however, one often uses multifrequency lasers (in the regime of the lowest transversal mode generation). There are several reasons why these lasers are used, such as the higher energy in a pulse, the simplicity of the resonator construction, and the higher stability. A shortcoming of these lasers is the existence of small-scale oscillations (of the order of lO-"s) in the time distribution, which are difficult to record without distortion by standard methods. In the case of multifrequency radiation (i.e., generation of many longitudinal modes of the lowest transversal index) we have
Here F( r) describes the fast oscillations inside the pulse envelope +( r). One should keep in mind that, owing to the inertialess nature of direct multiphoton ionization, the atom reacts to the instantaneous value of intensity. This is why in the case of multifrequency lasers, the ion yield fluctuates from pulse to pulse owing to the nonreproducibility of their fine temporal structure. Owing to the high peak values of radiation intensity in the fluctuations, the average ion . . pulse energies: yield is higher than it is for single-frequency radiation at equal (Ni) ( F K ) = g K ( F ) , . In this case the measured quantities are the same (Q,+ ( I ) , and $(x, y, z)), while to the calculated quantities there is added either g K or H K , and F , is replaced by (F). (See Section 1I.C.) The majority of measurements of the multiphoton cross sections have been made with nanosecond radiation, so we shall not consider the problems that arise when one measures multiphoton cross sections with pico- or femtosecond lasers. Thus, in order to find the multiphoton cross sections (Y, one has to measure Q , + ( f ) , $(x, y, z), n o , and N i and then calculate the values of T,S, F, V,, TK, and g K and substitute them into the expressions for a, presented in Sections 1I.G and H. When one observes the process of multiphoton atom ionization, a single laser pulse gives practically no information about the value of multiphoton cross section. In order to measure the latter, one needs a series of pulses both to obtain the data on N i ( F ) dependence and to get reliable data about the average values of the ion signals. The latter is especially important for multifrequency lasers. The need to use a series of laser pulses imposes certain requirements upon the stability of the laser radiation and atom target parameters. The stability of the radiation produced by modern pulse lasers is limited by technical fluctuations:
-
76
Ammosov, Delone, and Ivanov
by the instability of the voltage feeding the pumping lamps, by the fluctuations of the lamp discharge, by the instability of the temperature and of the thermal lens of the active element, and so on. One may also consider the fundamental fluctuations due to the quantum nature of light and the discreteness of atoms, but such fluctuations are not dominating in modem setups. A. MEASURING THE ENERGY OF A LASERPULSE
The measurement of multiphoton cross sections imposes the following requirements upon the device that measures the energy of the laser radiation. First of all, it should measure the energy values ranging from 10-* to several J’s. Second, the absolute gauging of this device must be as accurate as possible. For this reason, of all the ways to measure the laser radiation energy, one selects the calorimeter method. In standard calorimeters, the laser radiation is absorbed by a blackened metal cone whose temperature is recorded by a battery of thermocouples. The stability of these devices is sufficiently high (= 1% for the measurement time one hour), but the systematic error in absolutization of the measured energy value (i.e., the absolutization error) is around 10%. A higher accuracy can be reached with double-chamber calorimeters, in which one measures the temperature difference between the irradiated surface and an identical surface not exposed to radiation. The systematic error in devices of this type is about 3 or 4% (Demtroder, 1982; Rabek, 1982). The calorimeter is installed behind the interaction chamber (see Fig. 5 ) , so in the measurements one should take into account the energy losses due to reflection from the surface of the optical elements, including the chamber windows and lenses, placed between the interaction object and the calorimeter (about 4% at each surface). The energy losses on ionization can be neglected. Let us note that the full power of laser radiation coming into the calorimeter may lead to destruction of the blackened layer upon the receiving element and, consequently, to an uncontrollable increase in the systematic error. The calorimeter can be positioned in another way, where it receives only a small portion of the radiation energy that is directed upon it by a light-dividing wedge. In this case one has to gauge the light divider, and the error in its reflection coefficient will be added to the total measurement error. A weak point of calorimeters is their large inertia-of the order of seconds. No such problem arises with photoelectric detectors of radiation energy, but their response to radiation intensity is not standardized, so in order to use them one has to perform independent absolute gauging of the corresponding devices. Let us now consider the measurement error of radiation energy. The total error with which one measures the energy is made up by apparatus errors (both random and systematic, their typical values were presented previously) and by the
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
77
error associated with technical energy fluctuations from pulse to pulse. When measuring the multiphoton cross sections, a number of measures are taken to stabilize the operation of the laser: the stabilization of the feeding source, thermostabilization of the active elements, sustaining equal time intervals between laser pulses, etc. It is evident that the degree of stabilization should be the greater the higher is the power of nonlinearity of the studied process. Practice shows that for processes with the degree of nonlinearity K S 5 or 6, for which most of the experimental data on multiphoton cross sections have been obtained, it is necessary to stabilize the energy in a series of laser pulses to within 2 or 3%. The variation of the laser radiation parameters from pulse to pulse makes one consider the problem of controlling these parameters in each laser pulse. The measurement of Q , +(r), and +(x, y , z) in each pulse requires complicated and expensive apparatus that should include a computer for accumulating and processing the data. So this control is justified only in those cases where the multiphoton cross sections have to be measured with an accuracy higher than usual. Furthermore, one should take into account the range of variation of the radiation parameters in different regimes of laser operation. Compared with singlefrequency lasers, the multifrequency lasers have better stability of energy and of the spatial characteristics, but the fluctuations of the ion signals due to the multifrequency nature of radiation are high. For this reason, it is very unlikely that a pulse-by-pulse control of radiation parameters in this case would increase the accuracy of measurements. In the case of single-frequency lasers, such a pulseby-pulse control can lead to a noticeable increase of the measurement accuracy and, in addition, would enable one to eliminate those cases where two or more longitudinal modes are generated.
B. MEASURING THE TIMEDISTRIBUTION OF INTENSITY IN A LASERPULSE The standard procedure for measuring the laser pulse duration in the nanosecond region comes to the following. (See Fig. 5 . ) Part of the laser radiation is directed by a light-dividing wedge upon a photodetector. If additional weakening of the signal is needed, one can use neutral optical filters. As a photodetector one uses coaxial photoelements or fast-acting photodiodes that transform the light pulse into an electric signal. The latter is fed to a wideband oscilloscope, and the distribution on its screen is recorded on a film. Such a procedure can give a time resolution of to 1O-Io s. The most fast-acting photodetectors are photodiodes with a time resolution of lo-" s. They can be used together with wideband oscilloscope with a band up to 10 GHz. Wider possibilities are offered by pulse-periodic lasers that enable one to make measurements by using stroboscopic oscilloscopes with the time resolution = l o - " s (Demotroder, 1982; Rabek, 1982).
78
Ammosov, Delone, and Ivanov
By using the apparatus for fast signal processing (such as a transient recorder, a fast transient digitizer, or a boxcar integrater), followed by computer processing of data, enables one to exclude the photographic procedure. For measuring the picosecond pulse durations one uses streak camera with a time resolution below lo-’* sec. The femtosecond durations can be measured by nonlinear-optics methods (Demtroder, 1982; Rabek, 1982). Let us now consider the accuracy with which one can measure +(r). There are two sources of errors: (i) the fluctuations of + ( t ) from pulse to pulse and (ii) the apparatus errors in the +(t)measurements. The apparatus-measurementerror consists of a systematic error and a random error. The first one, 6Tab,,is determined by the time resolution and by the absolutization error. The time resolution of different devices already has been presented. As a rule, the absolutization error in the +(r) measurements is no less than 10% (as it is, for instance, in proper calibrators of oscilloscopes). In order to increase the accuracy of absolutization, different time tags are used. In certain cases one can bring the absolutization error down to several percent. The random apparatus error, 6T,,, is associated with the errors of reading from measuring devices (that are due to the finite width of the line on the oscilloscope screen, of the pointer, and of the scale marks, etc.). Usually, this error is small and does not exceed 1 or 2%. The fluctuations of +(t) are mainly due to the technical fluctuations in the laser-operation regime. They can be made smaller by stabilization. Their typical value is ST, = 1 to 3%. The methods of measuring + ( t ) in the case of single-frequency lasers and in the case of multifrequency lasers are essentially different. Whereas in the former case one should make the detector time resolution as high as possible, in the second case there exists an optimal value of the detector time response 7.On the one hand, the detector should smooth out the small-scale oscillations F( t) (7 > lo-’’ s); on the other hand, it must not distort the envelope +(t) (7 < 10-9 s). Thus, for the limiting error with which one can measure the time resolution, one can write down
+
(6T)2 = (6Tabs)24- (6Trm)2 (6Tn)*.
The preceding analysis clearly shows that the main contribution is given by the systematic error of absolutization, ST,,, , and by the error 6T, due to fluctuations of +(r) from pulse to pulse. The minimum error in +(r) (about several percent) is obtained when one uses single-frequency radiation, minimizes the technical fluctuations of + ( t ) , and takes special measures to gauge the measuring apparatus. The typical values are 6TflIT = 2 to 4%, 6Tab,/T= lo%, and 6TIT = 10 to 15%. By controlling +( r) in each laser pulse, when measuring the multiphoton cross
79
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
sections, one is able to eliminate the contribution 6Tn in the total error 6T. In those cases where a pulse-by-pulse control is impossible, one should measure &t) for a large number M of radiation pulses. The error in the average value will then be given by the formula 6Tn = (6T,,),,,/-.
c. MEASURING THE SPATIAL DISTRIBUTION OF LASERRADIATION INTENSITY Measuring the spatial distribution $(x, y , z) of the laser radiation intensity is a difficult problem since it is necessary to measure a three-dimensional distribution. This problem can be divided into two subproblems. The first one is to measure the distribution $(x, y , zo) in the focal plane of the focusing lens. The second subproblem is to measure the distribution $(x, y , z ) along the z axis of propagation of the laser radiation near the focal plane. The procedure of measuring the distribution of the laser radiation intensity in the focal plane of the lens comes to the following (Fig. 5). A small portion of radiation is directed by an optical wedge 4 upon a lens identical to the one that focuses the main part of the laser beam upon the atom target in the vacuum chamber. The optical paths from the laser to each of the two lenses should be made equal. A microscope with 10- to 100-times magnification reflects the focal plane of the auxiliary lens upon a photographic film or a photographic plate. (The typical values of the area into which the laser radiation is focused are S=3 X to 3 X lo-* cm2.) The intensity of radiation in the auxiliary beam is selected, by using neutral filters, to be within the range where the blackening of the photoemulsion is linear. The intensity distribution $(x, y, zo) is obtained by processing the photographic images with a microphotometer. In more detail, this photometric procedure is presented in Rabek (1982). The error of measuring $(x, y, z,,) is made up by the systematic and the random apparatus errors (6S,,, and as,) and by the error SS, due to fluctuations of $(x, y , z) in a series of laser pulses. The typical values of the errors are: 6Sab,/S= 10% and 6S,/S 1 to 2%. The value of 6Sn/S depends upon the stability of laser operation and upon the mode composition of its radiation. This problem has not been examined in detail in the literature. By controlling $(x, y, zo) at each laser pulse, one is able to eliminate the contribution 6S, in the total error 6s. In those cases where a pulse-by-pulse control is impossible, one should measure $(x, y, zo) for a large number M of laser pulses. The error in the average value will be 6S,,= (tSSIf),Jfl. A typical value of the total error is ~1
(6S/S)2 = (6S,,,)2
+ (6Sn)2 + (6Sm,)2 = 15%.
(77)
Lompre et al. (1982) have proposed a method of measuring S by computer processing of the video image. The laser radiation was recorded by a silicon mosaic vidicon with the diode matrix 225 X 225 elements. The video signal
80
Ammosov, Delone, and lvanov
went through the image amplifiers and to a high-speed digital encoder. From there the information went to the graphic PERICOLOR terminal coupled to a computer. In contrast to the long procedure of photometering, in this setup the value of S is calculated with an error SS/S = 4% in 5 seconds. This method enables one to make a pulse-by-pulse control of the spatial distribution cross section. The spatial distribution $(x, y, z ) along the z axis of the optical system can be measured in two ways. The first way is to record successively the different cross sections of the focal region in a series of pulses. This can be done by both photographing and processing video images. The other way is the method of a mirror wedge, which enables one to record different cross sections of the focusing region in one radiation pulse. The successive reflection from two mirrors produces a series of spatially separated images of the focusing region. By selecting the proper angle between the mirrors and the proper incidence angle of radiation, one can get the desired number of images with necessary spacings between them. A similar method was used by Lompre et al. (1982). There, instead of a mirror wedge, the radiation was reflected by two parallel glass plates, so that, simultaneously, one recorded four different cross sections of the focusing region. Thus, having measured the intensity distribution cross sections at different z , one can obtain the dependence of intensity on the zo coordinate. In those cases where only estimates of the multiphoton cross sections are needed, and there is no need to measure the cross sections with the smallest error possible, the procedure of measuring $(x, y , z ) can be simplified substantially if one makes use of the fact that the intensity distribution in a laser beam is close to Gaussian. As is known, the intensity distribution of a focused Gaussian beam has the form 1 2 yz $(x, y , z ) = 1 ( z A / 2 ~ r 2exp( ) ~ -ri (zA/2~r,)~
+
+
+
where A is the radiation wave length and ro is the beam radius at the focal waist, which is taken to be the distance from the beam axis, over which the intensity lowers e times. As one can see from this expression, the wave length and the waist radius completely fix the intensity distribution of a Gaussian beam. The radius ro is related to the beam radius at the lens entrance R and to the distance L between the lens and the waist:
ro = ALI~TR.
(79)
If the lens diameter is large enough: D > 3R, the intensity distribution in the vicinity of the waist is monotonic and has no local maxima and minima typical of those cases where a plane wave is focused by a lens of finite size (Born and Wolf, 1964). The fundamental difference between the two cases is that in a Gaussian beam the intensity falls off smoothly from the axis to the periphery.
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
81
Thus, for a Gaussian beam, the measurement of the spatial intensity distribution comes to measuring the cross section of the beam in the plane of the focusing lens and the distance between the lens and the waist. In conclusion, let us note that the Gaussian distribution gives a good description to the radiation of single-frequency lasers and of multifrequency lasers with a large number of longitudinal modes in the transversal TEM, mode. However, the intensity distribution at the focus of the lens can be distorted substantially by its aberration. When focusing laser radiation, one can practically always get rid of aberration of slanting beams by thorough adjustment of the system. It is however impossible to eliminate spherical aberration for a single lens. The following inequality gives the condition under which the distortion of the wave front of the beam will be negligible:
CR4 3
-€-A.
f’
4
C = 1.3 for a plano-convex lens (Born and Wolf, 1964).
D. MEASURING THE NUMBER OF IONS
PRODUCED
The ions produced in the region of laser radiation focusing are extracted by an electric field into a detector (Fig. 5 ) . The ion detectors used are either secondary electron multipliers of different types or Faraday cylinders. The operation of a multiplier is based on the effect of an avalanche-like secondary electron emission. In usual multipliers, the proliferation of electrons takes place during the flight through a system of electrodes. In channel multipliers this happens on the walls of a single spiral-shaped channel about 1 mm in diameter and about 100 mm long. In a microchannel multiplier, the electrons are multiplied in a large number of parallel microchannels each about 0.05 mm in diameter and about 5 mm long. The amplification coefficient for the usual multipliers ranges from lo5to lo6; for channel multipliers they are lo6to lo8.In the measurements, the signal from the multiplier goes through a preamplifier matched to the cable and onto the low-impedance input of a wideband oscilloscope or a high-speed ADC. If necessary, the measuring circuit can be supplemented by a wideband amplifier. As a source feeding the multipliers one uses stabilized high-voltage rectifiers. The random error of the measurements done with multipliers is small: around several percent. It is determined by the stability of the source feeding the multipliers, by the fluctuations of the amplification coefficient, by the fluctuations of the dark current, and by the random apparatus error of the oscilloscope. In the relative measurements of the number of ions, N,, the random error determines the total measurement error.
82
Ammosov, Delone, and Ivanov
In the absolute measurements of the number of ions produced, the error is determined by the systematic error of absolutization. As a rule, the latter is quite high: 20 to 100%. The reason is that the secondary emission coefficient (and, consequently, the amplification coefficient) depends upon the type of ions, their charge, and their kinetic energy. So it is necessary to make absolute gauging of the multiplier by using a precision source of the studied ions. A shortcoming of the multipliers is the small dynamic range where the amplification of the signal is linear: from two to four orders of magnitude for linear amplification. It can be increased by changing the feeding voltage, but then one has to make additional gauging. Let us note that in the ion-counting regime, where the multiplier detects individual ions (owing to the large amplification coefficient), the measurement error is minimum and is determined only by accidental triggerings of the detector. In order to realize this regime, it is necessary that within a laser pulse there be produced only a small number of ions, i.e., that the ions come to the detector with intervals greater than the proper time of the multiplier (1-30 ns). This regime can be used to measure the multiphoton cross sections with the absolute method. (See Section 1I.G.) The absolute measurements of the number of ions are also made with the Faraday cylinder by measuring the charge of the ions that have accumulated upon the cylinder walls. The error of such measurements is around 25%. In order to isolate the useful signal from the signal of residual gas ions, the ions are separated according to their mass in the flight gap preceding the detector. The scheme of the time-of-flight mass spectrometer with two field-filled gaps and one 50-cm-long field-free flight gap gives the mass resolution m/Am = 100, which is quite sufficient for the experiments discussed. In order to gather all the ions produced into the detector, one has to sustain sufficiently low pressure in the interaction chamber and in the flight gap to avoid collisions of the ions with atoms. The length of the flight gap has to be smaller For the ion-atom collision than the free path length of ions: X < hi = cm2 and X = 100 cm, the density of the gas should cross section crai = torr.) (For pressure this means p S be no S 10" Furthermore, to gather all the ions, one has to use a sufficiently large diaphragm upon the extracting electrode, with diameter d , > e (see Fig. 5), where t? is the effective length of the region where laser radiation interacts with atoms. We mention this trivial requirement because for radiation from the visible frequency region the length e may reach several millimeters. Sogard (1988) has shown that the ion-yield dependence can be substantially distorted if the ions from the periphery of the interaction region are not recorded. This is especially important for processes with a small number of photons, where the length of the effective interaction volume is close to, or may even exceed, the Rayleigh length 2,. (Z, = 27rr3h >> ro is the distance counted off the waist of the Gaussian
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
83
beam, over which the intensity at the beam axis decreases two times.) For a Gaussian beam, the effective length of the interaction region is t!K = VK/SK Z R / f l , where V K is the effective volume of the K-photon process given by Ehq. (87) of Section F and SK = .rrri/K is the effective area of the waist. To gather all the ions produced, one also has to make sure that the extracting potential difference is sufficiently high and that saturation of the ion signal occurs with increase of the extracting voltage. The value of the minimum extracting potential difference that ensures the gathering of all the ions into the detector is determined by Coulomb repulsion of ions (Ammosov et af., 1989):
-
where 5 is the degree of ionization. Let us mention that for [n, the value of Urninmay reach several kilovolts.
3
10'O
E. MEASURING THE DENSITY OF NEUTRAL ATOMS In those cases where laser radiation is focused in a gas of examined atoms, the atom density is measured with vacuum ionization gauges, provided that the density of residual gases is much lower. As a rule, the density used ranges from to lo* to lo'* cm-3. (The corresponding pressure range is torr.) The error is determined by the gauging of the apparatus and is under 10%. When an atomic beam is used, the atom density in the beam can be measured in several ways. In the measurements of multiphoton cross sections one often uses the method of surface ionization or the method of depositing the atoms upon a piezoelectric cell. The device that measures the surface ionization is an emitter made of tungsten in the form of a strip or a filament, upon which the surface ionization of beam atoms takes place. The electrons or ions produced with ionization are collected by a constant electric field upon a collector. The latter is shaped either as a plate (for a strip-shaped emitter) or as a cylinder coaxial to the emitter (when the latter is a filament). For each kind of atoms one selects a certain material for the emitter. For instance, for alkaline and alkaline-earth atoms one uses either pure tungsten (for Li, K, Rb, Cs, and Ba) or oxidized tungsten (for Na, Mg, Ca, and Sr). The measurement error is determined by gauging and amounts to about 20%. In the deposition method, one uses a piezocrystal coupled to a resonance circuit. The atoms are condensed upon the crystal surface, and as the layer of atoms becomes thicker, the resonance frequency of the crystal changes. The measurement error is also determined by gauging and is around 25%. One can also use the ionization method and the method where one observes resonance fluorescence. In the first case the atoms of the beam are ionized by an electron beam or by ultraviolet radiation. The ions produced are recorded by a
84
Ammosov, Delone. and lvanov
detector. The error with which the density is determined is around 30%. In the second case the atoms of the beam are excited by a dye laser (with radiation frequency equal to the resonance frequency of the atoms) and the light they emit is detected by a photoelectron multiplier. The error is of the same order as in the previous case. In both cases it is determined by the apparatus gauging. The most accurate way to measure the atom density is to use the vapor of examined atoms instead of an atom beam. In this case, into the interaction chamber one places an open ampule containing the metal whose vapor is then ionized. The whole chamber is placed into a thermostat and is kept at high temperature, so that the metal vapor fills the chamber. After that, the measurement of density comes to measuring the chamber temperature and to determining from this temperature the saturation vapor pressure. In this case the error is determined by the quality of the thermostat where the chamber is kept and amounts to several percent. F. CALCULATING THE LASERPULSE INTENSITY, THE EFFECTIVE PULSE DURATION, AND THE EFFECTIVE INTERACTION VOLUME Having measured the distributions +( t) and $(x, y , z) and the energy Q in a pulse of laser radiation, one can calculate the value of intensity Fo at the maximum of the space-time distribution: Fo
=
1 Q - [photons/(cm2 x s)],
ho ST
where S = J $(x, y , zo) dr dy and T = J + ( t ) dt. $(x, y, z) and + ( t ) are integrated graphically. This produces an additional error, which is usually small and does not exceed 1 or 2%. When calculating the absolute values of S and T, one should pay attention to the normalization of the $(x, y, z) and + ( t ) distributions. For calculating the multiphoton cross sections (in Eq. (82) for intensity and in any other expression for c y K ; see Section II) both distributions have to be normalized to the maximum value, i.e., the value at the point where the intensity is maximum is taken to be unity. This choice of normalization is related to the nonlinear character of interaction. Some authors have used other normalizations, such as the normalization with respect to the level lle, or l/e2, or 1/2. If the values of S and T have been calculated in one of these normalizations, in order to obtain the maximum intensity value in Eq. (82) one has to use conversion coefficients. They are easily calculated in the case of a Gaussian distribution (with respect to the space and time coordinates):
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
== 1.43 F ( l / e )=z 4.00
F(lIe2)
85
0.83F(1,2).
Considering what has been said previously, when $(x, y, z) is measured along the radius of the laser beam cross section on the focusing lens by using Eqs. (78) and (79) for a Gaussian beam, the value of S is calculated as follows:
S =ma
(84)
where r, is the beam radius at the intensity level lie. The error with which the intensity is measured is made up by the errors in the measured quantities:
($j2 =
($?)2
+
(!E)2+ (!E)2.
As already mentioned, the main contribution to the latter errors comes from the absolutization error. When one uses the radiation of a stable single-frequency laser, measures Q,T, and S for each laser pulse, and takes measures to lower the absolutization error (measures the energy with a precision calorimeter and uses time markers to measure the pulse duration), one is able to reach an error 6FlF =z 10%. For instance, by using the method of Lompre et al. (1982), Normand et al. (1989) have measured the intensity with a 6% accuracy. Such measurements, however, are either very labor-consuming or require very complicated apparatus. As a rule, even when one uses a multifrequency laser, T and S are measured in a separate experiment, and, when measuring the cross section, only the energy is controlled in each laser pulse. The typical values of 6FIF in this case are,25-30%. Finally, let us mention Perry and Landen’s (1988) proposal of a “direct” method of measuring the field strength based on the changes in the ionization potential Z of the atom under the action of an alternating field, by measuring the kinetic energy E , , ~ ” = Khw - Z(E) of the electron produced. However, in reality, it is still difficult to give a strict substantiation to this method owing to the existence of above-threshold absorption (Delone and Fedorov, 1989a,b), and, also, in view of the limits on the conditions under which one is able to distinguish clearly an elementary ionization event from pondermotive effects in the region where the laser radiation is focused (Delone and Fedorov, 1989a,b; Goreslavsky et al., 1989). As it follows from the principal relations presented in Section 11, from the measured quantities 4(t)and $(A y, z) one should also calculate the time T K and the volume V, of interaction that are effective for a K-photon process. Let us recall that the differences between T K and T and between V, and V, and thereby
86
Arnrnosov, Delone, and Ivanov
the need to calculate TK and V,, are due to the nonuniformity of the space-time distribution of laser radiation. In principle, these quantities can be measured directly by using multiphoton detectors. For example, it is possible to measure T2 by recording the time distribution of the second harmonic of laser radiation, excited in a nonlinear crystal. However, the measurements with multiphoton detectors are difficult to make correctly (Delone and Masalov, 1980), so this method is practically never used. The value of T K = J 4,(?)dt is found by raising the distribution 4(t)to the Kth power and integrating it. The error with which T K is found is somewhat higher than the error in T since (6TK)ran= K6T,, (and similarly for (8TK),), while the systematic (absolutization)errors are the same: (ST,),, = 6T&,,. If the distribution 4(t)is close to Gaussian, we have T K = T / a , and the total error S T G , where STG is associated with the deviation of +(?) from is 6TK = ST the Gaussian form. The typical values are GTKITK -- 10 to 15%. The value of V, = J JIK(x,y, z ) Cix dy dz is calculated as follows. The distributions I,!@, y, z , ) measured at different values of zi are normalized, raised to the Kth power, and summed up: V , = Az Xi S K I . Here Az is the discrete step in z with which the cross sections of $ ( x , y, z i ) are selected; SKI = J $,(x, y, z , ) ak dy is the effective area of the ith cross section. The error SV, is given by the following expression:
+
Here (SS,/S,) is the error with which the effective area of the cross section is found; (SS,), = KISS, is the error given by the fluctuations of $(x, y , z ) from pulse to pulse; ( S S , ) , = KSS, is the random apparatus measurement error; = 6Sinl is the integration error; (SV,), is the error associated with discrete selection of the cross sections; and e K = V,/S, is the effective interaction length. If the distribution $(x, y . z ) is close to Gaussian, then having measured the beam radius R at the entrance to the lens, the distance L from the lens to the waist, and thereby the waist radius ro (see Eq. (79)), one can calculate V , according to the formula:
1
for K = 2 = [ r ( K - (7/4))]-1’2 for K
(2K - 4)!!
3.
87
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
In this case the error SV, has the form
where 6L is the error in L and ( 6 V K ) ,is the error associated with the departure from the Gaussian form. The typical values of SVK/VK are 15-20%. G . CALCULATING THE IONIZATION PROBABILITY PER UNITTIME AND THE MULTIPHOTON CROSSSECTION The process of multiphoton ionization is characterized by two fundamental parameters: by the K-photon cross section and by the probability per unit time w at a given value of intensity. The relation between these two quantities is given by Eq. (6): w = aKFK. However, the observed quantity is not w , but the total probability W per laser pulse. Let us write down the relations that connect a K , w , and W to the measured quantities we have already discussed. For a singlefrequency laser we have W
= w/TK = ni/(noTK) = Ni/(VKTKno)
ffK =
N,/( iloTKVKFK) ,
(894
for multifrequency radiation we have W
= N,/(noTKFKgK),
LYK
= Ni/(noTKVKFKgK).
(89b)
The error in the probability 6wlw is made up by the errors in the measured quantities:
($)’
=
(F)’+ (2)’+ (z)2 + (2)’.
(90)
As was previously shown, the typical values of the errors are
and, consequently, 6wlw = 40%. The error with which one finds the multiphoton cross section is substantially higher, since it includes K times the intensity-measurement error. Since B a K I a K 3 1, the measurement error can be presented in the form 6A = S(lg aK), where aK = 10A’SAand (SA)2 = (lg e)’
[(T)’+ (?)*+ r2)’+ r?)’+
Kz($)‘].
88
Ammosov, Delone, and Ivanov
The typical value is 6FIF c= 30%.For this reason, for example, for three-photon ionization we have 6a3/a3= 100% and 6A = 0.3. H. THEGENERAL PROCEDURE FOR MEASURING MULTIPHOTON CROSSSECTIONS
Before measuring the multiphoton cross section, in a separate experiment one measures the degree of nonlinearity K of the ionization process (Section 1I.D). This is dictated by the need to check whether the ionization mechanism is direct (i.e., that there are no resonances) and to select the range of laser radiation intensity where the measurements should be made. One records the ion-yield dependence on the laser radiation intensity (Figs. 6 and 7). For calculating the cross sections according to the absolute method, one has to make measurements in the nonsaturated region, where the ion yield N i is proportional to F K . The different modifications of the relative method require the measurements to be made both in the nonsaturated region and in the region of saturation (Section 1I.H). The advantage of the relative method is that there is no need to measure the neutral ion concentration no, and one only has to make relative measurements of the ion yield N i and of the effective interaction volume V,. As a shortcoming of the relative method one may consider the need to obtain the experimental data in a wide range of radiation intensities, including the regions where the dependencies N i ( F ) are saturated, which requires high laser radiation intensities. A very important problem for any experiment where the multiphoton cross sections are measured is the selection of the operation regime for the laser. As was already mentioned, it is not recommended to use a laser that generates many transversal modes. The choice between the two possibilities-a laser that generates many longitudinal modes of the lowest transversal index and a laser that generates single-frequencyradiation-determines the experimental methods and, in the end, the resulting accuracy of measurements. When one uses multifrequency lasers (which, compared to single-frequency lasers, have more stable Q,S,and T in a series of successive pulses), the spatial and time distributions of radiation intensity can be measured in a preliminary experiment. Then, the cross section measurements come to measuring the ion yield as it depends upon the laser pulse energy. The method of measuring the cross section in this case is the following. First, one measures the c#~(t)and I,!J(x, y, z) distributions. Then, one measures the power of nonlinearity of the process, i.e., the dependence N i ( Q ) . After that, in the required region of the N i ( Q ) dependence, one generates a series of laser pulses with constant energy, so it is enough to measure the energy in a single pulse of the series. The ion yield, measured for each pulse, is averaged. Let us emphasize that for multifre-
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
89
quency lasers, owing to fluctuations, the peak value of intensity can be two or three times higher than the maximum value of the measured average intensity. For this reason, in this case the dispersion of the ion yield is very large, and for measuring the multiphoton cross sections one must use a larger number of laser pulses than in the case of single-frequency lasers. When measuring the N , ( Q ) dependence, the energy of a laser pulse should be varied in such a way so that the 4(?)and +(x, y , z) distributions would remain the same. There are several ways of doing this: (1) By using neutral filters. (2) By using the polarization method for completely polarized laser radiation. The energy of the laser radiation that has gone through two polarizers is proportional to cos28, where 0 is the angle between the axes of the polarizers. (3) By weakening the radiation by reflecting it from a surface. The laser radiation is let through two identical plane-parallel transparent plates, one of which makes an angle 8' with the optical axis of the system, the other being inclined at the angle n - 8'. By changing the angle of the first plate one can change the transmission coefficient, while the second plate restores the direction of the beam along the system axis. The second method is the most convenient one. The first method does not allow one to vary the energy continuously and, owing to the nonlinear effects on glass filters, it is difficult to use for laser radiation of high intensity. The third method yields to the second one only in the relative complexity of energy-variation calibration. Often, one uses a combination of the first method with the second or third one. One cannot recommend varying the radiation energy in a pulse by changing the operation regime of the pumping lamps for active elements of the generator and/or of the amplifier. As a rule, this changes the space-time distribution of the radiation, which was shown by controlling it in each laser pulse (Lompre et al., 1982). As concerns the measurement of multiphoton cross sections, a shortcoming of multifrequency laser radiation is that one has to measure the statistical factor g, = J F K ( t )d?/(F)K(Delone et al., 1980) (or the pseudostatistical factor H , ) . The direct measurement of F( r) requires apparatus with very high time resolution (no worse than lo-" s). As a rule, one uses the values of g, calculated for Gaussian radiation with a different number N of longitudinal modes (Masalov and Todirashku, 1980). (For N + m, g K + K!.)The number of modes in the laser radiation can be found by measuring its spectral width. In this case, when calculating the cross section, one has to take into account the error associated with the deviation of the envelope 4(t)from the Gaussian form. When one uses single-frequency lasers (which are less stable than the multifrequency ones), the experimental technique is different in that it becomes necessary to measure the radiation energy in each pulse. In addition, one has to
90
Ammosov, Delone, and Ivanov
control the time distribution +(r) in each laser pulse to exclude the cases where two or more modes have been generated. An advantage of using single-frequency radiation is the smaller dispersion of the ion signals at constant radiation energy and no need to measure g K . One can give the following recommendation for measuring multiphoton cross sections with minimum error. First of all, one should have a stable singlefrequency laser. The intensity distribution should be measured in each pulse, which enables one to eliminate the error associated with fluctuations of the energy Q,duration T, and cross section area S. The energy should be measured by a precision two-chamber calorimeter; T should be measured with time tags. This enables one to reduce the main error-the error of intensity value absolutization. S and V, sliould be measured by combining the method of a mirror wedge with computer processing of the video image. By measuring the intensity this way one is able to bring the error down to 6FIF = 10%. (Normand er al. (1989) have reached 6%.) Furthermore, the measurements should be made according to the relative method (see Section ILH), where there is no need to measure the neutral atom concentration and one only has to make a relative measurement of N , and V,. In this case the errors S N , / N , and SVKlV, are substantially lower-around several percent. The values a, have not been measured this way before, evidently, because such measurements require much work. However, this method enables one to expect a minimum error: GaK/a,= KSF/F -- K X 10%.
IV. The Results of Measuring the Multiphoton Cross Sections of Direct Atom Ionization In Table I1 we present practically all the data on absolute values of the cross sections aK measured with laser radiation that have been published up to this time (1990). Most of the data have been obtained by using linearly polarized laser radiation, and only a small portion of them have been obtained with circularly polarized radiation. For all the studied atoms, the measured values of the cross section refer to multiphoton ionization from the ground state; only for He have the values of a, been measured for ionization from the ground state and from the first excited states 2s3S, and 2s1S0. Both the absolute and the relative methods have been used for measuring a,. For frequencies presented in Table 11, the fact that the process of multiphoton ionization was direct has been checked by measuring the dependence of the ion yield Ni on the laser radiation intensity F and thereby determining the power of nonlinearity Kexp= d(lg Ni)/d(lgF ) . In all the cases in Table 11, the measured value of the nonlinearity power was equal, within the experimental error, to the
TABLE I1 EXPERIMENTAL VALUES OF MULTIPHOTON CROSS SECTIONS oLK OF DIRECTIONIZATION PROCESS ~~~~
'0
Measurement method
K
Atom
2 2 2 2 2 2 2 2 2
K K cs Ca Ca He(2slS0) He(2s'S I ) Xe Xe
3 3 3 3
Na Na* Na
2 2
K
2
3 3
K*
3
Rb Rb*
2 2 2 2 2 2
3 3 3 3 3 3 3
K cs cs* cs
1
I I 2 2 2 2 -
-
1
cs
1 1
Ca Ca
2 2
~~~
w,
cm-'
IgU;
lg U K
gK
References
18.870 18,870 19,940 28,185 29,700 28,797 28,797 51,400 5 1,700
-47.2 ? 0.8 -48.8 ? 0.8 -49.2 ? 0.1 -46.3 t 0.5 -46.0 2 0.5 - 48.6 Z: - 48.8"; -49.4 - 52.0
2 2 2 2 2 2 2 2 2
-47.5 2 0.8 -49.1 ? 0.8 -49.5 2 0.1 -46.6 f 0.5 -46.3 2 0.5 - 48.9 ?: -49.1 T f i -49.7 -52.3
Held et a/., 1972 Delone et a/., 1973 Normand and Morellec, 1980 Alimov et a/., 1986a Alimov er a / ., 1986a Lompre ef a/.. 1980 Lompre et a / ., 1980 McCown et a/., 1982 Bischel et a / . , 1979
14,400 14,400 18,870 14,400
-76.3 +- 0.1 -75.9 t 0.1 -79.6 ? 1 . 1 -78.5 t 0.1
I I 6 I
- 76.3 ?
14,400 16,537 14,400 14,400 14,400 14,400 14,410 14,410 18,790 18.928
-78.0 -78.2 -76.8 -76.5 -76.7 -76.4 -76.0 -77.0 -78.7 - 77.6
1
- 78.0 5
Cervenan ei a/., 1975 Cervenan er a / ., 1975 Delone e t a / . , 1973 Cervenan et a/.. 1975 Cervenan and Isenor, 1975 Cervenan et a / ., I975 Alimov et a / ., 1980 Cervenan ef a/., 1975 Cervenan ei a / . 1975 Cervenan et a/., 1975 Cervenan et a/., 1975 Evans and Thonemann, 1972 Fox e r n / . . 1971, Koganetd.. 1971 Akramova e t a / . . 1984c Bondar and Suran. 1986 Bondar et a / ., I985
t 0.1 k 1.1 ? 0.1 t 0.1 t 0.2
0.2 t 1.5 t 1.5 k 0.5 k
6 1 1
1 1
6 6 6 6
0. I -75.9 + 0.1 -80.4 ? 1 . 1 -78.5 2 0.1 0. I -79.0 ? 1.1 - 76.8 ? 0. I -76.5 2 0.1 -76.7 ? 0.2 -76.4 ? 0.2 -77.8 2 1.5 -77.8 + 1.5 -79.5 ? 0.5 - 78.4
.
(continues)
TABLE I1 (continued)
References
N rg
3 3 3 3 3 3 3 3 3 3 3
Ca Sr Sr Sr Ba Ba Ba He(2s I S,,) He(2s1S,,) He(2s3S,) Xe
2 2 2 2 2 2 2
4 4
K K
1
4
cs
1
1
2 2 1
I
cs
1
Mg Mg Ca Kr
2 2 2 I
16,785 18,790 18,828 17,655 18,790 18,875 16,092 14,415 14,400 14,400 33,860
6 6 6 6 6 6 6 6 6 6 1
9,435 9,435 9,435
- 104.2 f 1.2 -107.8 f 0.2 - 107.0 f 0.2
24 24 24
9,470 18,790 16,790 15,527 33,860
- 108.1 5 0.2 - 104.1 f 0.4
24 24 24 24
-111.1:::
-112.5::; - 115.1 f 0.6
1
- 82.8 -+0U6s -78.8 -77.6':.: - 77.0 -78.6 ? 0.4 -75.3:: - 78.4fO-3 4 5 -79.1:;: - 80.3 - 81.3 f0.2 -0 6 -82.7 f 0.1
Bondar and Suran, 1988 Feldman et a/., 1984 Bondar et a/., 1983 Bondar and Suran, 1988 Akramova et al., 1984b Bondar e t a / . . 1984 Bondar and Suran, 1988 Bakos et a/., 1976 Lompre et al., 1980 Lompre et a/., 1980 Peny and Landen, 1988
- 105.6 2 1.2 - 109.2 2 0.8
Held et a/.. 1972 Delone er a / ., 1973 Delone et a / ., 1973 Arslanbekov er a/., 1975 Normand and Morellec, 1980 Alimov et a/., 1986b Bondar and S u m , 1990 Bondar and S u m . 1988 Perry and Landen, I988
- 108.4 2 0.2 -109.5 f 0.2 -105.5 f 0.6 - 112.5::; - 113.9::; - 1 1 5 . 1 ? 0.6
I 2 2 2 2 2 2
9,435 9,435 9,450 9,395 9,435 9,395 9,395
- 138.0 f 1.7 - 136.9 f 0.5 - 140.6'1; - 141.6 - 141.61bA - 138.9 f 1.0 - 138.7'(,:
I 20
- 1 4 0 . 1 2 1.7
1
5 5 5
Na Na Sr Sr Sr Ba Ba
120 120 120 I20 I20
- 136.9 ? 0.5 - 142.7';; - 143.7 - 143.71;; - 140.0 2 1.0 - 140.81bi
6 6 6
Ca Ca Ca
2 2 2
9,450 9,395 9,455
-
170.2 - 170.3 f 1.2 - 169.4::;
720 720 720
- 173.1'i: -173.2 2 1.2 - 172.311:
6 6 6 6 7 7 8
Hg Kr Xe Xe Xe Kr Ar
I 1 2 -
-
14,410 18,870 18,870 17,000 14,410 17,000 17,000
- 171.5 f 2.3 - 170.2 f 0.6 - 169 - 173.7 -203.7 f 2.8 -205.5 - 237.5
720 720 I 1 5040 I
-174.4 2 2.3 -173.1 & 0.6 - 169 - 173.7 -207.4 ? 2.8 -205.5 -237.5
Aleksakhin et a / . , 1979 Akramova er al., 1984a.b Bondar et al., 1985 Bondar and Suran, 1986 Chin et al., 1969 Agostini et al., 1970 L'Huillier et al.. 1983a Perry et a / . , 1988 Chin et al., 1969 Perry er al., 1988 Perry et al., 1988
11 I1
Xe Xe
2 2
9,435 9,435
-336 ? 2.0 -338.2
I
-336 f 2.0 -338.2
Arslanbekov and Delone, 1976 L'Huillier er al., 1983b
22
He
2
9,435
-731.9;::
1
- 731.91::
Lompre et al., 1976
5
5 5 5
a
W
1
1 1
Delone er a/., 1973 Arslanbekov et 01.. 1975 Aleksakhin et a/., 1979 Feldman er a/., 1984 Bondar et a / ., 1983 Akramova er a/., 1984b Bondar et a / ., 1984
K is the nonlinearity power of the ionization process; o is the laser radiation frequency; ah is the measured value of the cross section; gK is the statistical factor of radiation; a, is the value of the cross section with account of the statistical factor: Ig aK = Ig a; - Ig g K . In the "Measurement method" column, I and 2 refer to the ?bibsolute and relative methods of measuring a,, respectively. All other notations are explained in the text. * indicates the values of a, measured for circularly polarized radiation; all other data have been obtained for linear polarization.
94
Ammosov, Delone, and Ivanov
minimum number of photons that should be absorbed to ionize the atom, i.e., one had Kexp= K. One should note however that the accuracies with which the measured dependencies d(lg N,)ld(lg F) could be approximated by a straight line with the slope Kexp= K were substantially different. This accuracy was determined not only by the type of the laser used and by its stability but also by the type of the ionized atoms. For instance, as a rule, for alkali atoms Kexp= K with high accuracy, and with increase of radiation intensity the powerlike dependence with Kexp= K clearly transformed into the dependence Kexp= 3/2, corresponding to ion-yield saturation. On the contrary, for ionization of alkaline-earth atoms, the dependence Kexp= K , as a rule, is very inaccurate. Sometimes one can qualitatively see that a single powerlike dependence with Kexp= K is not realized (Bondar et al., 1988a). The detailed experimental studies, in which besides ion detection one also measured the energy spectra of produced electrons, have shown that in the case of alkaline-earth atoms, owing to absorption of more than the threshold number K of photons, the single-charge ions are produced not only in the ground state but also in excited states (Delone and Fedorov, 1989a). This process explains the deviation of the observed dependencies d(lg N,)ld(lg F ) from the powerlike ones with Kexp= K. The production of ions in the ground A + and in excited (A+)* states shows that in the standard experiments, where one records only the yield of singlecharge ions, one actually measures a summary effect produced by realization of different channels in which the ions are formed in different final states. Accordingly, all the characteristics extracted from the measured total yield of ions in different final states are also of integral character. This evidently applies both to the ionization probability and to the multiphoton cross section. To a certain extent, the notion of an integral ionization probability as of the probability of production of single-charge ions per unit time (in any final state) is justified, though, as one can see from the principal relations presented in Section I, w is expressed through quantities that depend on K. Numerically, the inaccuracy in w due to the K-dependence of T K is small since the dependence of T K = f ( K ) itself is not very sharp (see Section 1II.F) and the spread in the value of K for processes with production of ions in different final states is also not very large (Delone and Fedorov, 1989a). Both in principle and qualitatively, the situation is worse with cu,, which is determined from the integral yield of single-charge ions produced in different states. From the principal point of view, aKis a function of K and even the dimensions of aKare different at different K. From the quantitative point of view, the strong powerlike dependence of aK = w/FK on F produces a significant inaccuracy in the value of aKdetermined from the integral ion yield. Let us now summarize the situation with the values of the multiphoton cross sections presented in Table 11. On the whole, the data can be clearly divided into
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
95
two groups. The first group is the data on alkali atoms. The experimental results on production of double-charge ions have been obtained only by He et al. ( 1 985) for ionization of alkali atoms by ultraviolet radiation. Double-charge ions have been observed only when the radiation intensity was greater than 200 MW/cmZ (E = 5 X lo5 V/cm). The power of nonlinearity of this process is K = 10. In the visible and near-infrared regions we are interested in, the power of nonlinearity is two to three times higher, so the radiation intensity necessary to produce double-charge ions should be much higher than the value presented previously. However, the measurements of the cross sections of direct alkali atom ionization by radiation from the visible and near-infrared frequency ranges (K =Z 5 ) have been made at E =Z lo5 V/cm. Thus, the values of the multiphoton cross sections presented in Table I1 for alkali atoms have been measured under such conditions (at such frequency w and intensity F of radiation) where both the above-threshold ionization and production of excited ions can be neglected. Thus, the values of multiphoton cross sections for alkali atoms presented in Table I1 are well grounded and, with good accuracy, characterize the threshold ionization channel A
+ Khw-A'
+ e-,
(92)
where the atom absorbs the threshold number K of photons and the ion is produced in the ground state. It is evident that, in principle, these data should be described in the framework of the one-electron approximation by nonstationary perturbation theory in the first nonvanishing (Kth) order. The second group is the data for alkaline-earth atoms. .The experimental data of many studies (see, for example, Suran and Zapesochnyi, 1975; DiMauro et al., 1988; Haugen and Othonos, 1988; and Camus et al., 1989) show that the measurements (whose results are summarized in Table 11) have been made under such conditions (for such w and F) where the ions are produced not only in the ground state (A+) but also in excited (A+)* states. (See, e.g., Figs. 3 and 4.) Thus, the values of multiphoton cross sections for alkaline-earth atoms presented in the table, as a rule, are summary characteristicsfor several ionization channels S of quanta absorbed, in the final state of the ion, differing in the number K and in the electron kinetic energy:
+
Thus, the index K that labels the quantities aKcharacterizes only the threshold (with respect to the number of photons absorbed) channel of ionization. It is evident that any rigorous theoretical description of the ionization process in such cases should be made outside the framework of the one-electron approximation and with account of the structure of concrete channels, of their relative weights, and of the dependence of the latter on radiation intensity. In many cases, the measurements of multiphoton cross sections have been
96
Ammosov,Delone, and Ivanov
made with multifrequency lasers, and in a number of cases with the radiation of the second harmonic generated by such lasers. In all these cases, the calculation of multiphoton cross sections from the experimental results requires that one take into account the statistical factors of radiation, g, or HK . However, the values of these factors are not always known. The situation is the simplest in the case where the measurements are made with the fundamental radiation of multifrequency lasers. Here the statistical factor depends upon the number of generated longitudinal modes N = Av * 2L,. (Av is the width of the laser radiation spectrum in cm-I and L, is the length of the laser resonator.) The value of g , varies from 1 for single-frequencyradiation to K!for radiation with a large number of longitudinal modes. According to Delone et al. (1980), the number of modes for which the asymptotic limit g , = K! is realized depends upon K and varies from N = 10 for K = 2 to N = lo3 for K = 11. The analysis of the experimental conditions under which the results in Table IJ have been obtained shows that in all those cases where multifrequency radiation was used, the value of g , was close to the asymptotic limit K!.For this reason, for the fundamental radiation of multifrequency lasers it was taken that g , = 10". = K !. The situation is more complicated with statistical factors for the second harmonic of multifrequency radiation: here g? depends not only upon the number of modes generated but also upon the efficiency of conversion into the second harmonic. In the limit of low conversion efficiency, where the intensity of the second harmonic is proportional to the fundamental radiation intensity squared, gimcan be expressed through the statistical factors g , and g , of the fundamental radiation: g;to = (FL)/(F2JK= (FZK)/(F;)" = g l K / g f . When the number of modes is large, we have gi- = (2K)!/2K,which is greater than K!.In the limit of high (almost 100%) conversion efficiency, where F, = F,, the statistical factors g? and g , coincide. Thus, for a large number of modes, the statistical factor of the second harmonic of a multifrequency laser, gp, varies between K ! and (2K)!/2K.At K = 2 this is the range from 2 to 6, at K = 3 it is 6 to 90, etc. The absence of any data on the efficiency with which the radiation of a multifrequency laser is converted into the second harmonic, as well as the absence of calculations that would describe the dependence of gi- on this efficiency, increase the uncertainty with which the multiphoton cross sections are found. For this reason, the data in Table I1 on the cross sections are presented in two forms. In the fifth column we present experimental data processing with no account for the statistical factors. These values (denoted as lg ah) can be regarded as multiphoton cross sections only in those cases where the measurements were made with single-frequency lasers. In the case of multifrequency lasers, the quantities lg ah characterize the initial experimental data. In the seventh column we present the data on multiphoton cross sections where the statistical factors, presented in the sixth column of the table, have been taken into account: lg a, = lg a; lg g K .
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
97
As one can see from Table 11, there are several cases where the values of effective cross sections for the same atom under the same conditions (same frequencies and same polarization of radiation) have been measured in several studies. By comparing these data we see that, within the experimental error, they coincide. The exception is the values of the effective cross section of four-photon ionization of the K atom at the frequency of a neodymium laser measured by Held et af. (1972) and by Delone et af. (1973) and of six-photon ionization of the Xe atom at the same frequency measured by Alimov and Delone (1976), by Arslanbekov and Delone (1976), and by L'Huillier et al. (1983a). As of late 1990, there were no additional data that would enable one to tell which result is more correct. It should be noted that the errors presented in Table 11, with which the values of effective cross sections cu, for large K (K = 11, 22) have been found according to the relative method, can hardly be considered reliable. The point is that the studies by Alimov and Delone (1976) and Arslanbekov and Delone (1976) were done a long time ago, and the laser radiation used was not very stable in a series of pulses. Accordingly, the fluctuations in radiation intensity and in the ion yield in a series of pulses were high, but the authors did not always make a correct account of the large deviations of the ion signal from the average value, so the measurement errors presented are probably underestimated. Let us now discuss the results that concern the dependence of effective multiphoton atom ionization cross sections on the radiation polarization. Table I1 presented data on the absolute values of effective cross sections aK measured for circular polarization. Besides this, in a number of studies the authors have measured the ratio of the ion yield N: for circularly polarized radiation to the ion yield N f for linear polarization at fixed F: R = N f / N f . Under the conditions discussed previously, the ratio of the ion yields, R, will coincide with the ratio of multiphoton ionization cross sections for circular and linear polarizations: R = Nf/Nf = ak/ag. It should be noted that since for measuring R one does not have to make absolute measurements of the laser radiation intensity, the value of R can be found with a much higher accuracy than the absolute values of the cross sections. The experimental results for Rcxpobtained for fixed radiation frequencies are presented in Table 111. There also we present the values of R calculated according to the well-known factorial formula (Klarsfeld and Maquet, 1972): R
=
(2K
+
l)!!/K!.
(94)
This relation holds in the one-electron approximation for ionization from an arbitrary initial state with any main (n)and orbital (e)quantum numbers (Krainov and Melikishvili, 1988) in those cases where the power of nonlinearity is not very high (K s 5 ) and the ionization occurs on a frequency for which the compound matrix element Eq. (7) is dominated by the virtual transitions that obey the Bethe rule (n + n + 1; t' + e + 1). The constraint on K comes from
TABLE I11 EXPERIMENTAL A N D THEORETICAL VALUES OF RATIOOF MULTIPHOTON IONIZATION CROSS SECTIONS FOR CIRCULAR A N D LINEARPOLARIZATION OF RADIATION 0,
K
Atom
K
cm-I
R,,,
Rex*
References
2 2 2
cs cs
18,870 14,410 18,870
1.5 1.5 1.5
1.2 -C 0.4 1.28 2 0.19 0.9 -C 0.18
Delone, 1975 Fox e t a l . . 1971; Kogan etal.. 1971 Klewer et a / ., 1977
3 3 3 3 3 3 3 3 3 3 3 3 3
Na Na K Rb cs cs Ca Ca Ca Ca Ca Ca Ca
18,870 14,400 14,400 14,400 14,400 14,410 18,245 18,483 18,504 18,591 18,607 18,790 19,816
2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5
0.42 +. 0.08 2.30 2 0.09 2.66 ? 0.11 2.16 2 0.13 2.24 ? 0.11 2.15 2 0.4 0.89 +. 0. I 1 1.01 4 0.14 0.73 2 0.12 0.65 +. 0.10 0.60 2 0.08 I .o 0.03 +. 0.005
3 3
Sr Sr
17,216 17,298
2.5 2.5
2.73 2 0.38 1.47 +. 0.18
Delone, 1975 Cervenan et a / ., 1975 Cervenan et a/., 1975 Cervenan et a/., 1975 Cervenan et a / ., 1975 Foxeta/.. 1971;Koganeta/.. 1971 Bondar et a / ., I988b Bondar et a/., 1988b Bondar et a / . 1988b Bondar et a / . . 1988b Bondar er a/., I988b Akramova et a/. 1984c Bondar et a / ., I984 Bondar and Suran, 1986 Bondar et a / ., 1988b Bondar et a / . 1988b
.
.
.
3 3 3 3 3 3 3 3 3 3 3 3
Sr Sr Sr Sr Sr Ba Ba Ba Ba Ba Ba He(2s3S,,)
18,832 16,642 16,808 16,903 17,106 17,696 I 7.762 17,784 17,894
4 4 4
K K
18,870 14,415
2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.5
2.5 ? 0.4 2.38 f 0.46 2.85 2 0.35 1.25 ? 0.23 0.79 f 0.11 1.29 f 0.23 1.42 f 0.24 0.91 2 0.18 1.19 ? 0.20 I .o 0.7 2 0.2 2.9 f 1.2
Bondar et a / ., 1983 Bondar e t a / . , 1986, 1987 Bondar e t a / . , 1986, 1987 Bondar e t a / . . 1986, 1987 Bondar ct a / . , 1986, 1987 Bondar er a / . . I988b Bondar et a / ., 1988b Bondar er a / ., 1988b Bondar et a / . . 1988b Akramova er a/. I984c Bondar er a / . , 1984 Bakos ei a / ., 1976
cs
9,435 9,395 9,435
4.4 4.4 4.4
1.67 f 0.51 3.9 2 0.8 4.0
Delone, 1975 Akramova e t a / . . 1983 Delone, 1975
5 5 5 5
Na Sr Ba Ba
9,435 9.435 9,434 9,395
8 8 8 8
0.5 5.5 2 0.090 0.330
6 6
Ca Ca
9.395 9,435
14 14
0.72 f 0.25 1.4 2 0.3
18,790
* 0.1 1.8 2 0.009 f 0.006
.
Delone, 1975 Bondar et a / ., I983 Bondar et a / ., 1984 Akramova et a/.. I984c Akramova er a / ., I984c Bondar et a / ., 1985 Bondar and Suran, 1986
100
Ammosov, Delone, and Ivanov
neglecting the C-dependence of the matrix element that describes the transition into the continuous spectrum. This neglect is justified if the values of C for the final states in the case of linear (e = 1) and in the case of circular (C = K ) polarizations are not very different. With lowering of the contribution to the compound matrix element from the channel where the Bethe rule holds, the value of R also becomes smaller compared to Eq. (94), but still remains above unity: R > 1. The exception is the relatively narrow frequency intervals where there are intermediate resonances allowed in the field with linear polarization and forbidden in the field with circular polarization, so that ak = 0. Thus, one can summarize the situation by saying that under the conditions just specified, the direct multiphoton ionization proceeds more efficiently in the field with circular polarization. The special choice of the frequency that distinguishesthe transitions satisfying the Bethe rule has not been made in any of the studies except Bondar et al. (1988b) and Alimov et al. (1987), where the data for several alkaline-earth atoms have been obtained. Our analysis of the conditions under which the experiments have been made has shown that the necessary criteria are also satisfied by the data obtained by Fox et al. (1971), Cervenan er al. (1975), Delone (1975), Bakos et al. (1976), and Kogan et al. (1971). If of all these data we consider those that can be compared with the factorial relation (94) with sufficient reason, we may conclude that this relation holds for ionization of alkali atoms and breaks down for ionization of alkaline-earth atoms. One should suppose that the discrepancy between the experimental data and Eq. (94) in the case of alkalineearth atoms is associated with the fact that it is impossible to describe the interaction of laser radiation with these atoms in the one-electron approximation, which is also indicated by other data (Delone and Fedorov, 1989a). In a number of studies (Agostini and Lecompte, 1976; Samson, 1982; Bondar et al., 1983, 1984; Akramova et al., 1984c; Bondar and Suran, 1986) the authors have made a detailed examination of how the ion yield N , depends upon the ellipticity parameter 8 of laser radiation (upon the ratio of the polarization ellipsoid half-axes). Under the conditions similar to those that are necessary for aKto hold, the dependencies N,(8) will be the proportionality relation N , proportional to the 8-dependenciesof the ionization cross section a K ,i.e., N,(O) a,(@).In all the studied cases, the dependenciesN,(8) were monotonic curves. With enough justification, the dependencies N,(B) can be compared at present only with the calculations made in the one-electron approximation for the threshold ionization channel (92) (Delone and Krainov, 1985;Smith and Leuchs, 1984). Let us once again turn to the data presented in Table 11. From all these data it follows that at fixed values of the nonlinearity power K,the values of the effective cross section of direct multiphoton atom ionization coincide within the difference between the values of aK at minimum and maximum detuning of the intermediate resonances,
-
-
CROSS SECTIONS OF DIRECT MULTIPHOTON IONIZATION OF ATOMS
101
Let us note in conclusion that despite the relatively large body of experimental data obtained through 1990, there are several reasons why the program of measuring the effective cross sections of direct multiphoton atom ionization cannot be considered complete. In the first place, up to now the detailed experimental studies of multiphoton cross section have been made only for individual groups of atoms (alkalis and alkaline earths). Evidently, one should make similar measurements for other groups of atoms. In the second place, the measurements should be made under such conditions where one knows that neither the above-threshold ionization nor the production of ions in excited states takes place. Accordingly, in addition to detecting the ions, one should also measure the electron-energy spectrum and the luminescence due to relaxation of excited states. The method of recording the energies of produced electrons is well developed and is widely used in the studies of above-threshold ionization (Delone and Fedorov, 1989a, b) and of multicharge ion production (Delone and Fedorov, 1989a). The study of Haugen and Othonos (1989) convinces one that additional information can be obtained by measuring the luminescence. Only when a single ionization channel is dominating can the experimental data be correctly characterized by the value of the multiphoton cross section and the latter can be calculated theoretically in the framework of the corresponding model. When realizing such a program, one should keep in mind that in all the cases one should have constraints upon the intensity of the laser radiation used that arise from the fact that in a strong electric field, there appear dynamic multiphoton resonances where the Stark shift is ~ E ( E2) AE,,"?,. Let us recall that the constraints upon the field strength are actually constraints upon the nonlinearity power of the process that can be recorded.
V. Analytical Expression for Estimating the Multiphoton Cross Sections of Direct Atom Ionization As is known (Delone and Krainov, 1985), in principle, the rigorous calculations of any quantity can be made only for the hydrogen atom, since only there one knows the exact wave function. However, even in this case there arise practical limits in numerical calculations made by a computer. A sufficiently complete account of the results of such calculations made according to the perturbation theory is presented by Karule (1978, 1985, 1988). When one turns to more complicated atoms, even with only one valent electron, there arises a need to use different approximation methods in the framework of the perturbation theory, since neither the exact potential of the frame nor the exact wave functions are known. In practice, the simplest are the methods of the quantum defect and of the model potential (Delone and Krainov, 1985; Faisal,
102
Ammosov, Delone, and Ivanov
1987; Morellec er al., 1982; Manakov et al., 1986). The calculations of the ionization cross sections aK for alkali atoms made with these methods are presented in a number of papers for several values of K and for narrow frequency ranges. However, in order to make such calculations, one has to design special programs for each kind of atom and frequency interval. As was already mentioned, both in the case of hydrogen and in the case of other atoms with a single electron on the external shell (alkali atoms), at sufficiently high field strength, the process of above-threshold ionization begins to play an essential role. For this reason, there always exist constraints upon the field strength more stringent than the requirement E E l , i.e., if qI> 0. Note that the rate of inelastic collision yl, will be zero for the optical transitions but it could be significant for nearby levels. Note further that y,, can be made nonzero by pumping the level li) from the level l j ) by using a broadband field. The decay of the coherences (off-diagonal elements of the density matrix) is given by Ts. These also include contributions from phasechanging collisions, i.e.,
The model (7) is based on the impact approximation for collisions. The collisions occur instantaneously with respect to various other time scales in the problem. Note that the model (7) assumes transitions within the states of the system. If the population leaks out of the system, then (7) can be generalized to include terms of the form (-iL&Il, = - r : , p , +
PA,.
(9)
Here p I is the rate at which the level li) is being pumped. Such a pumping is essential to maintain the equilibrium if r:, # 0. The pumping terms determine pc0).We shall mostly work with the model (7). The nonlinear-response theory can be developed by expanding p in various powers in H,. We would be considering the response of the system to external electric fields and since we would only be interested in dipolar response, we write H,(t) as
Here da is ath component of the dipole moment operator. Let us introduce the Liouville operator Lf
L,W
= [H,(t), Ilh.
(11)
118
G . S. Agarwal
The quantum mechanical time-dependent perturbation theory then shows that the density matrix pen) to nth order in applied fields is
x L,(w2) .
(&I
- w,,)-'Lf(w,,)pfo).
On combining Eqs. (lo)-( 12), we can calculate the response of the dynamical variable Q to external fields. The nth-order contribution to (Q(t))will be
dQ = [da,
1.
(14)
Here Sym stands for the symmetrization of the nth-rank tensor in front of it in the indices ( a l w l ) (, a 2 w 2 ) ,. . . , (~,Jo,,).Note that if Q is chosen to be the dipolemoment operator, then Slff)gives the usual nonlinear susceptibilities
x$il
an(wl
. . . 0,)
= n S j ~ ,ajw1
. . . wn)(p=dm,
(15)
where n is the density of atoms/molecules. Equation (15) assumes independent particle approximation. In what follows we shall use both S'")and x'"'. The fourwave-mixing and saturated-absorptionexperiments will be described by appropriate third-order nonlinear susceptibilities. For the purpose of calculating the laser-induced fluorescence we need to know, say, the population in the excited state. We shall see that collision-induced coherences can be seen in a variety of experiments involving four-wave mixing, saturated absorption, and laser-induced fluorescence. In order to obtain explicit expressions for the nonlinear response, we need the eigenfunctions of the Liouville operator Lo. From the structure (7) it is clear that Lo(i)(j(= A,lli)(j(,
A,
= 0,) -
iT,],
i # j.
(16)
In addition, Lo has eigenfunctions that depend on the inelastic part of the relaxation matrix. Let us diagonalize the matrix R defined by
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
119
It turns out that "04k
=
iAk4kr
4 k
=
2e
sf'k[e)(el.
(18)
The complete set of eigenfunctions La are given by (16) and (18). These depend on the collisional parameters. Note that the action of a function of Lo on any operator can be obtained from f(Lo)Q
=
x
k#e
Q M ~ ( L o ) I ~ )+ ( ~ IC, Q , k f ( L d ) k ) ( k j
The expression for first-order response is rather simple:
In deriving (20) we have used the property pi? = pi!'&. The first-order response is determined only by the relaxation of the coherences. Note that while the width of the resonance is dependent on the collisional dephasing, the collisions themselves do not lead to any new resonances.
III. Second-Order Optical Response and Collision-Induced Coherences We next consider the calculation of the second-order response. Here the model of relaxation can make substantial difference. We show that the collision-induced coherences appear in the second-order response of appropriate variables. Using (16) and (14) we can show that
120
G. S. Aganval
To simplify (21) we use (17) and (18):
Be,(W)
=
(W
- iAp)-lSqp(S-')pe.
P AP#O
Note that if one had used the decay model (9) (with yii = 0), then one would get the much simpler result (W
- Lo)-qe)(el
=
(W
+ iTee)-llt)(t1,
(23)
which amounts to using Beq(W)
+ See)-'.
= 6,(w
(24)
In view of (24) it is useful to introduce the auxiliary function Cqe(W) = B q d W ) - 6eq(W
+
iree)-l,
(25)
which identically vanishes for the model (9). Using (22) and (25) in (21) we get (Agarwal, 1986)
up =
0,
+
-
w2.
This is the most general expression for the second-order response of a system. Note that the third term (with (YO, ow2) and fourth term in the first square bracket can be combined to yield
s,
=
d$d$Qkj[Ajk
(Afi -
-
Wp
Wp)(Aik
+
i(rjk
-
rji
-
rik)]
- WJCAji - WI)
(27)
121
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
It is obvious from (27) that a resonance in Sjtdlpat w, = A,, can be seen if the decay of coherences is such that
r,, - r,, - rltz 0.
(28)
In addition, Q, terms show that one can have a resonance at w,
= 0 provided that
2r,t # -A,
(29)
where A, is one of the eigenvalues associated with the relaxation matrix defined by (17). The condition (29) depends on both phase changing and inelastic collisions. Thus, for a two-level atom with radiative relaxation and collisional dephasing, the dephasing-induced resonance w, = 0 will have a width 27, i.e., just the radiative width. We have thus shown that the second-order response of appropriately defined physical quantities can have two types of collision-induced resonances: w,k = w, and w, = 0. In the next section we discuss an explicit example of such resonances. The pressure-induced coherence between two ground levels (Bloembergen et al., 1985; Gong and Zou, 1988) can also be seen from (26). Consider a specific situation with two ground states Igl) and (g2)and say one excited state le) (Fig. 2). Let Q represent the operator (gl)(g21.We combine the first two terms in the first square bracket in (26). This leads to
S m
-
=
-
d:&fgl
2(A,,,, -
Pi:\z(Aegl - W)
[
-
-
~ ~ z e ~ ~ g l ~ i ~ ’ , l [u~ pg z ~ il( r e g l
+
rgze
wp)(Ag2e - wl)(Aegl
-
02)
-
]
- w,)
pk:bl(Ag2e
(Ag2e- w N L g 1 -
UP)
2(Ag2gl -
+
rg,gJI
4
(30) (31)
if p(Bq)gl= p60:,,. Expression (31) clearly shows that there is no ground-state Hanle resonance in the absence of collisions and in the absence of population imbalance between is negligible and the two ground levels Isl)and Is2). For Na ground state rg,g2 the Hanle resonance has a width essentially determined by the residual Doppler
FIG. 2. Schematic illustration of a three-level model used in the text for ground-state coherences.
122
G. S. Agarwal
14) (91
19> rz3.The broad resonance occurs only if rI2 T13- rZ3 - rlI# 0. Thus, the existence of the broad resonance requires the presence of collisions. The ratio of the intensities of the two resonances depends on pressure. The theoretical prediction was verified in experiments (Stahlberg et al., 1985) on the 2P4 ( J = 2)-3& ( J = I) transition in Ne. It is interesting to note that the resonance w13 = 0 with a width rI2 arises from the Doppler averaging of a term such as 1 (78) (o,?+ S,, - O , b ) ( w l 3 - ir,, - O , + kv) and, thus, the broad resonance can arise from the overlap of two Bennett holes. Finally, we mention that several collision-induced nonlinear dispersive effects have been studied by Zou and Gong (1990). They have for example reported collision-induced self-defocusing.
+
+ r,3
+
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
145
VII. Relation between Collision-Induced Coherence in Second-Order and Third-Order Responses We have seen in the previous sections how the pressure-induced resonances can appear in fluorescence spectroscopy and in third-order response which describes processes such as four-wave mixing and saturated absorption. The character of these resonances is the same in both second-order and third-order nonlinear response. One might therefore expect a relation between the two; i.e., one might look upon four-wave mixing and saturated absorption as ways to observe the resonances that are already present in the second-order response. To see this explicitly, we return to the basic Eq. (14) which we write for second- and thirdorder responses: ( - 1)3 s&2013(w1 0 2 1 w3) = -sym Tr[Q(wl + w2 + w3 - iLO)-lda1 3! (79) (w2 + w3 - i ~ ~ ) - ~ d ~-z iLo)-Ici"3p(o)], (w~ t
Suppose we want to calculate the induced polarization in third order. Then typically Q will be li)(jl, with Ii) and lj) connected by dipole transition. Note further that
=
(R - A,j)-l(jlB[i).
Thus, (79) can be written as s!&ap,(wl
9
w2. w3)
+ w2 + w3 - Aj,)-I
= (-1)3 (a1
3!
(82)
x sym Tr{14(jl[da1, GI), where G = (w2
+ o3- iLo)-Ici"2(w3- iLO)-'dap(O).
Equation (82) on simplification reduces to
(83)
146
G. S . Aganval
Note that the quantities that appear in (84) are basically the second-order responses given by (80). Consider now the first (second) term in (84). (See Fig. 21 .) It requires that dipole transitions are allowed between li) and l j ) and between Ij) and lk) (Ii) and Ik)) and, thus, the first (second) term requires the secondorder response between level li) and Ik) (Ik) and Ij)) which are not dipoleconnected. Thus, in the context of Zeeman levels, the second term in (84) would require the second-order response of the coherence between two Zeeman levels. The four-wave-mixing signal at 2w, - w2 can be obtained by choosing w, = w3 = w,, w2 = - w2 after the symmetrization has been carried out. It is thus clear that the third-order response, say four-wave-mixing signal, would exhibit any collision-induced resonances that are exhibited by the second-order response of coherences between two levels not connected by dipole transition as is illustrated in the diagram. For example, in the context of the three-level system discussed in Section 111, the states Ij) and Ik) would be, say, (1) and 12) and the second-order response p$)(wl - w2) exhibits a resonance whenever w, - w2 is equal to Zeeman splitting between 11) and 12). The analysis of this section would in turn imply that the four-wave-mixing signal at 2w, - w2 will exhibit a collision-induced resonance at 0 , - w2 = ? 2s. This in fact is the resonance first predicted by Grynberg (1981a). For a pump-probe absorption experiment we can choose 0 , = - w3. Then the induced polarization on li) t* 1j ) transition will depend on the coherence pik where Ii) and Ik) are, say, two ground levels. Thus, the collision-induced coherence in pik (cf. Eq. (31)) will also be seen in pump-probe absorption experiments as indeed was reported by Gong and Zou (1988).
FIG. 21. Schematic illustration of connection between coherences in second-order and third-order responses.
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
147
VIII. Collision-Induced and -Enhanced Resonances in Fifth-Order Nonlinearities It is clear from the discussion in the previous sections that the collision-induced resonances occur quite generally. We have already discussed a host of nonlinear processes that exhibit such resonances. One would expect such resonances in higher-order nonlinear processes as well. For example, six-wave-mixing processes might show these. The higher-order susceptibilitiesbecome very complex and just contain too many terms. However, in specific cases of two-level and three-level systems, the higher-order response can be calculated in closed form. We present a simple example. Consider six-wave mixing in a two-level system (Agarwal and Nayak, 1986). We e5amine the coherent signal produced at 3w1 - 2w2 in the direction, say, 3k1 - 2k2. Calculations using optical Bloch equations show that the induced polarization at 3w, - 2w2 is given by
x
(I- iAl -
T2
T2
x where A, nances at
(-!--
= w,, -
T2
-1
iAl - 2i6)
+ iAl - 2iS
(i
-
i6)(:
-
3i6),
w l , 6 = w1 - w2. This fifth-order response exhibits reso-
It may be recalled that 6 = 0 resonance in four-wave mixing is induced by collisions. Let us examine the character of 6 = 0 resonance in six-wave mixing to see if collisions are needed. Let us assume that A l >> l / T l , 1/T2, 6 so that (85) simplifies to
G . S . Aganval
148
where In the absence of collisions r = 0 and then
The signal shows an inverted resonance at x = 0. It is clear from the structure (87) that for large pressure, the structure of the resonance at 6 = 0 changes. This is shown in Fig. 22. The collisions enhance the resonance w1 = w2 in sixwave mixing. Trebino and Rahn ( 1987) reported collision-enhanced ground-state hyperfine resonances in sodium-seeded hydrogen air flame (Fig. 23). They observed at high intensities additional resonances at & w,/2 in nondegenerate four-wave mixing. Here ohis the ground-state hyperfine splitting. They also performed higherorder wave-mixing experiments in six-wave and eight-wave-mixing geometries. 2 Oh149 and They demonstrated the existence of subharmonics -+ whf21 ? ? wh/5. These subharmonics can be understood in terms of the nonlinear susceptibilities of higher order. For example, the subharmonics ? wh/2( 2 wh/3) can be
X"' NONLINEAR I T Y
-2
-1
0
1
2
X
FIG. 22. Six-wave-mixing signal (87) for different values of the collisional parameter r. The curves (a), (b), and (c) are for r = 0, 1 , and 10, respectively. The maximum values in (b) and (c) are 1.6 X 10' and 1.46 x 104, respectively. The minimum value in curve (a) is unity.
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
149
I
6WM Geometry
8WM Geometry
-0.00
-004
0 00 0.04 (ern-')
0 00
wI-wZ
FIG. 23. (a) Spectrum of collision-enhanced resonances in sodium in a flame obtained using a
6WM geometry (3w, - 202, in lowest order). Note the strong Zeernan resonance at zero frequency difference, the sharp four-photon subharmonic resonances at 20.03 cm-I (labeled 2), and the weak two-photon hyperfine resonances at *O.M crn-l (labeled I). Also observe the small dips at 20.02 cm-1 (labeled 3). which are probably due to six-photon resonances at * o , / 3 . The small constant background is due to scattered light and is not a nonlinear-optical effect. (b) Spectrum obtained using an 8WM geometry (40, - 3 0 ~ ) .Observe the subharmonics at 2 112, t 113, 2 1/4, and ?I15 of the hyperfine splitting (labeled 2, 3, 4, and 5, respectively). The subharmonics at tow,/5are due to at least 14WM (Trebino and Rahn, 1987).
understood in terms of (~('9. As indicated in Fig. 23, the explanation of & w,/5 requires at least 14 wave-mixing processes. Trebino (1988) has further shown that dephasing-induced nonlinear optical effects exist in all orders of the pertubation theory. To show this he rewrote susceptibility as a sum of terms all of which are proportional to the combination rU+ r,, - of the dephasing rates.
ck
M. Effect of Cross-Relaxationon Collision-Induced Resonances Our discussion of collision-induced resonances has been based on the assumption that the different lines do not overlap. If the levels are close by and if the pressure
150
G. S. Aganval
w1 13)
1
FIG. 24. Energy diagram of the model system with various relaxation rates. 27 is the radiative relaxation rate of each transition and (+ is the strength of collisional coupling between the two components.
is large, then line-mixing effects become important. In such a case the relaxation model (7) is to be changed. In order to see the effect of cross-relaxation on collision-induced coherences, we have examined a simple four-level model (Kumar and Agarwal, 1987) which is schematically shown in Fig. 24. The figure shows various allowed transitions and the collisional rates. Let 5 be the crossrelaxation parameter. The density-matrix equations for this model are given by u I p l 4 + 5pZ3- W 4 ( p U - p I 1 ) , 4 + Y + uIp23 + (PI4 - W3(p33 - p 2 A pI1= -(u + 2y)pll + w22 - i ( H ; 4 ~4 1p I 4 H L ) , PI4 p23
= -ti(wo + = -[i(oo -
$1 + Y +
(u + 2 7 ) -~ i ~( H ~S 3 ~ 3-2 P23%2),
P22
=
P33
= 2 ~ ~ 2-2up33
~ P II
pa = 2yplI
(90)
+ + up33 -
up4 upU
+ +
i ( H h 2 - P23%2)? i(H14~41 - P14H:I).
Here H ' ( t ) is the atom-field interaction in the dipole approximation. The coherences are coupled to coherences due to the cross-relaxation. These densitymatrix equations are easily solved to obtain the linear and nonlinear susceptibilities. The linear susceptibility x ( I ) ( o , )is found to consist of, in general, two lines X"'(WI)
= -
where
As the cross-relaxation 5 increases beyond s, one of the two lines in the absorp-
15 1
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
tion spectrum becomes narrower and stronger while the other line becomes u/s broader and weaker. Finally, only the narrower line survives. For {/s >> 1 one has
-
Hence, at very high pressure, one observes a single narrow line with width approaching the natural linewidth. This is the well-known line-narrowing phenomena resulting from the mixing of the lines. The third-order susceptibility responsible for four-wave mixing is found to be
x(3)(wI, 0, , - ~
2
)
A+ + A , + ih, +
-6 x[-(
A,
A+ ih+
+
+ A,
c
-
1
+ 2i(y + u) - 6 + A , + ih,
6
A,
+
-
A,
ih,
(6
+ A,
ih-
I (94)
ih-
-6
+ A, +
ih-
+ ih-
1 +
+ 6 + A,
- iX,
A+A-
-
+ A, +
+ A, + ih-
A+
'(6
A-
-6
- ih+
-
6
+
where 6 = w, - w 2, A, = wo - w , , and A, and A , are defined by (92). Let us now examine the effect of cross-relaxation on the collision-induced resonance. Let the fields w, and w2 be detuned far from the atomic resonance, i.e., A , , A2 >> y , (T,etc. Then (94) leads to an approximate expression I X ' Y ~ I w,, 1
-w*))2
a
1
+
2 i ( u - 5) 6 2iy
+
(95)
Therefore, the collision-induced resonance at w, = w2 becomes less prominent
152
G. S. Aganval
as cross-relaxation increases. Note that the width of the resonance is still determined by the radiative relaxation. The numerical results for the behavior of the four-wave mixing signals for a range of relaxation parameters can be found in Kumar and Agarwal (1987). Note that simplified theories of line-mixing effects are given by Hall et al. (1980) for four-wave mixing and by Kothari and Agarwal (1990) for six-wave mixing.
X. Dipole-Dipole Interaction-InducedResonances We have so far considered the resonance induced by collisions with atoms of the buffer gas. In the absence of the buffer gas but' at high densities the dipoledipole interaction becomes important and this can lead to extra resonances (Leite and De Araujo, 1980; Rand and Lam, 1987; Varada and Agarwal, 1991). For example, consider a system of two two-level atoms with frequencies R, and R2. Let this system interact with an external field of frequency w , . We can now study the probability of simultaneous excitation of both atoms by absorbing two photons from the field. If the atoms are noninteracting, then the probability p12of simultaneous excitation is the product of the probabilities of finding each atom in the excited state. Thus, in the absence of direct interaction between two atoms pI2will exhibit resonances at w , = R, ,a2. However, if the dipole-dipole interaction (d-d interaction for brevity) between atoms is significant, then p l Zexhibits an additional resonance (two-photon resonance) at
a, + R2 = 2w,.
(97)
The d-d interaction-inducedresonance (97) can be easily understood in terms of the perturbative expression for the transition probability for two-photon absorption. Let us label the states for the two atoms as shown in Fig. 25. The states Ij, a)and li, p) are connected by d-d interaction. Thus, the states after the d-d interaction part has been diagonalized will have the structure
FIG. 25. Energy-level diagram for two two-level atoms and the two-photon transition between the energy levels of the total system.
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
153
where S is a 2 x 2 diagonalization matrix. The matrix element rn for two-photon transition from the state (i, a)to Ij, 0) will be m =
I I%>
FIG. 27. Dipole-dipole interaction-induced excitation of two atoms by the absorption of one photon of frequency w,
-o
~ ~ .
which in terms of the collective operators can be written as H = ~ o s 2+ V(S+S- - 1 - F).
(105)
The eigenstates of (105) are the triplet and singlet states. The triplet states have energies oatV, and - oo. Thus, the spontaneous-emission spectrum will consist of lines at oo f: V. For the case of atoms driven by the coherent field, the spectrum of fluorescence can be calculated from (102). The calculations are rather complex and have to be performed numerically. The actual form (Varada, 1990) depends on the relative magnitudes of the dipole-dipole interaction and the strength of the external field. We can consider other types of situations where new resonances can appear due to d-d interaction. Consider the case of transitions shown in Fig. 27. Clearly, the pairwise excitation of the two atoms to the state Ij, p) is possible even to second order in the field. The dipole-dipole interaction leads to resonance at o, = w,, wpn. This clearly involves matrix elements of the form (pjl Vlyi)(yilH,lai). On the other hand, single-photon excitation involves a matrix element of the form (yilH,lai). Such problems can also be handled in terms of the general expression (26) which gives the fluorescence. In (26) the states V and d is to be replaced by are to be replaced by the eigenstates of Ha d, d l . The probability of pairwise excitation can be obtained by choosing Q = 1j , p)( j , PI. Note that such a problem can also be solved exactly (Rand and Lam, 1987).
+
+
+
XI. Collision-Induced Resonances in Spontaneous Processes In the introduction we discussed how the collisions can lead to redistribution of radiation. The radiation-field spectrum is determined by the two-timecorrelation function of the dipole-moment operators. Such correlation functions can be calculated for any system (Kumar and Agarwal, 1991) by following a procedure similar to that used in the calculation of the nonlinear polarization.
G . S. Aganval
156
For example, such correlation functions to second order in the applied field will show the collision-induced resonances in the radiation emitted by a quantum mechanical system. In order to understand the collision-induced resonances in emission, let us consider the Eq. ( 5 ) for a two-level atom (Fig. 1):
aP -- - - y1 z , ( S + S - p
-
2S-pS'
2
at
1
- -y12(S-S+p - 2S'pS2
+ pS+S-) + ps-S+)
Here y2,(yI2) is the rate of decay of the excited (ground) state to the ground (excited) state and r ( p ) is the collisional dephasing parameter. In the previous sections we calculated the nonlinear response to applied fields, i.e., we evaluated quantities such as (S-(t)), (Pet)). We now need to calculate correlations such as (S+(tl)S-(t2)).Such correlations can for example be calculated by writing Langevin equations equivalent to (106). One can, for example, show that the dipole-moment operator is given by
+ F - ( r ) - -hi [ S - , H,(t)] where F ( - ) ( t )is the noise term such that ( F - ( t ) ) = 0,
( F * ( t ) F z ( t ' ) ) = 20'TEj(t -
t),
etc.
(108)
The diffusion coefficients can be obtained in the standard way and one, for example, finds that 20'-
+ yI2
= T(p)(S+S-)
( 109)
which is equal to TqS'S-) as generally y12= 0. The diffusion coefficient D +has the very interesting feature that it vanishes if the collisional dephasing rate is zero. If r ( p ) # 0, then it depends on the strength of the applied field as (S'S-) depends on the external field. Thus, the fluctuations in the dipole-moment opera-
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
157
tor depend on the collisional parameter P). This can be seen from an explicit calculation of the two-time correlation functions. Assuming that the applied field is monochromatic (frequency o1) and weak, the calculation shows that ((S+(O)S-(7))
- (S-)(S+))e-= d r
A = w0 - w,,
(g( =
l$l.
Hence, the scattered spectrum will exhibit a collision-induced coherence with a width l/Tz. For the multilevel case a similar result is obtained. In place of (107) we have (10
(A) = -(r,, -
iw,)li)(jl
-
ni [IWI, H,(~)I+F,,(~),
i
zj .
(110a)
The correlation function of the noise is found to be (Fi,(f)F,k(t’)) =
2DJNf -
f’),
(1 11)
i # j # k.
(112)
where 2Dt,.,k =
Kk-
r,, - r,k>(pt,),
We again see that the noise correlations such as (1 11) are nonzero only if collisional dephasing exists. The noise correlation is at least of order two in the applied field amplitude. Equation (1 12) shows very generally the existence of pressure-induced contributions to the quantum properties of the system. FROM TRAPPED STATES A. COLLISION-INDUCED RADIATION
We next discuss another example in which collisions play a dominant role. Consider the Raman scattering (Agarwal and Jha, 1979) in presence of strong fields. The model system is shown in Fig. 2. We assume that the inelastic collisions can occur between the levels Is,)= 12) and lg2) = 13). The detailed calculations show that if the fields w1 and w2 are tuned so that A, = A? = A, then in the steady state such a system does not radiate provided that the inelastic collisions between 12) and 13) are absent. In presence of collisions the spectrum of the
158
G. S. Agarwal
(U-tlz
I /y,
FIG. 28. The spontaneous Stokes spectrum in the case of exact resonance and strong fields. The spectrum is identically zero if collisional parameters u , and u2 are set to zero (Aganval and Jha, 1979).
radiation emitted near the frequency w,, consists of spectral peaks at (Fig. 28)
where G I = dl3 * El/fz, G, = d,, * E2/fz.One can interpret the resonances as due to the collision-induced coherence at frequencies given by (1 13). The frequencies (1 13) correspond to the transitions among dressed states of the system. The dressed states and their energies in the special case A, = A2 = 0 are given by
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
159
The density-matrix elements in the dressed state basis obey equations similar to (7): bij
=
-iStjPi,
-
ciJP,J;
=
-2 7J,Pi,+ C ytJPJJ. J
(115)
J
All the relaxation parameters depend on the strength of the applied fields. These are given by
Figure 29 shows the various allowed transitions in the dressed-state picture. The figure also shows the transitions if we let v -+0, i.e., if we do not allow any inelastic collisions between the two ground levels. In such a case all the populaas can be seen from the explicit result tion is trapped in the dressed state)~)1 #6;;(") = [as;, a =
+ as, + ( 1
v,IG114 +
-
2a)6;l),
v2lG2I4
~ ~ ( +1 2 1~ ~1 1 +~~ )v 2 ( I ~ , l +4 2 1 ~ 1 ~+ )~ G ~ ( Y , I G ~ I ~+ ~ ~ 1 ~ ~ 1 ~ '
( 1 17)
FIG. 29. Schematic illustration of the transition among the dressed states in both absence and presence of collisions ( v , , v2 f 0).
G. S. Agarwal
160
Thus, in the absence of collisions, no radiation is possible. This is like no radiation from a system in ground state. Collisions in the presence of strong fields which then radiate as the state 11) is a populate the dressed states I+*)and superposition of the dressed states I&) and
Finally we note that the instantaneous intensity produced by such a system shows an interesting off-on behavior (Agarwal et al., 1988) (quantum jumps) particularly if y, is several orders of magnitude bigger than y 2 v, # 0. The off state corresponds to the dressed state 12) - (G2/G,)13)whereas on states correspond to the dressed states 11) 2 (G,/Go)(2)? (Gl/Go)13).It should be remembered that this off-on behavior occurs only because collisions lead to a nonzero decay rate of the level 12) to 13).
-
xu[.
Narrowing and Enhancement of Signals Due to Velocity-Changing Collisions
We have so far ignored the effect of the velocity-changing collisions (VCC) on the generation of different types of signals. The formulation of Section II can be extended to include such collisions. For simplicity we consider here only the case of two-level optical transitions. The basic density-matrix equation ( 5 ) is modified by the addition of (Berman, 1986)
(1 19)
T,,(v)=
I
W,,(v+ v’) dv‘.
(120)
The quantity r,,(v)gives the rate at which atoms in the state (i) and with velocity v either decay to other states or change their velocity as a result of VCC. W,,(v’ + v) brings population from other velocity subclasses into the subclass v. The v) ( i # j ) are analogous quantities for the coherence quantities T,,(v),W,,( v’ p,,( v). We have made the following simplifying assumptions in addition to those given in Section 11: Changes in velocity occur instantaneously, i.e., the time scale for such changes is much smaller than all the other relevant time scales in the problem.
COLLISION-INDUCED COHERENCES IN OPTICAL PHYSICS
161
The value of v following a jump depends at most on the values v’ before the jump (Markov approximation). The VCC and the phase-interruptingcollisions are assumed to be independent. The dynamic equations for the two-level atom in an external field could now be solved using the model ( 1 19). A reasonable model for the collisional kernel is given by Keilson and Storer (1952). However, the analysis simplifies considerably in the limit of strong collisions. In this model collisions result in rapid thermalization of the velocity distribution of active atoms. The model is especially suited if the mass of the perturber atoms is much bigger than the mass of the active atoms. We thus take
W J ~+ ’ v)
=
r,(v)M(v),
(121)
where M(v)is the Maxwellian velocity distribution given by (122)
~ ( v = ) (~‘Gv,~)-le-v+~fh.
The four-wave-mixing signal can be (Singh and Agarwal 1988b, 1990; Gorlicki et af., 1988; Lam et at., 1982) calculated depending on the geometry. The results in the Doppler limit and in the limit of no saturation are found to be:
( I ) Phase-ConjugationGeometry: Xgi(WI9
0 1 9
-%)
+2 A1
=
-
wI,
f i [ 2 r + ~ 7(r22 ~ ~ - rll)y - i6
6
=
(wI
-
(rI,+ rZ2)w1
~ 2 ) .
In this configuration, the signal shows two resonances at 6 = 0, and 2Al. For no buffer gas both resonances have a halfwidth 2y. The resonance at 2Al broadens with increase in the buffer-gas pressure. The resonance 6 = 0, however, narrows and gets enhanced with increase in pressure. Its width now is determined of the ground state. Generally one expects Tzz by the velocity-changing rate rZ2 100 LHD). Another possibility is to reaccelerate the ap+-ion, which has been considered in some detail by Kulsrud (1989), and it requires an essentially matter-free space, which is, however, inconsistent with the assumption that ap+-ionsare generated by fusion processes in a hydrogen target. C. RECAPTURE In addition to stripping, there is a possibility of capturing continuum muons. The final-state Coulomb interaction of the negative muon with the outgoing a-particle results in an enhancement of the muon phase-space distribution in the vicinity of the a-particle (Miiller et al., 1989a; Shin and Rafelski, 1990a, b). Some of these so called convoy muons can recombine with the a-particle by external Auger or radiative capture. The fraction of muons escaping with a velocity somewhat greater than that of the a-particle are decelerated more rapidly due to their smaller mass, enabling the a-particle to catch up with these muons and capture them into a loosely bound state. Environmental conditions, such as target density or the presence of magnetic fields, may influence the fraction of muons available for capture. Computations (Miiller et al., 1989a) show that the fraction of convoy muons can be as great as 0.8%, compared with less than 0.1% when the final-state a-p interaction is neglected. Convoy muons contribute in principle an additional term to the sticking probability, which could be strongly density-dependent due to the competition between the stripping of captured muons from outer shells and the density-independent radiative transitions to the tightly bound inner shells. Clearly, the convoy effect adds to the sticking fraction presented in Section 1V.A. Therefore, if the convoy effect is indeed a cause for the density-dependence of sticking, the intrinsic sticking must be much smaller than expected as of 1990, the value of which is already smaller by a factor of two than the current experiments. (See Section 1V.E.) It should be emphasized at this point that as of 1990, the theoretical calculations show that muons convoy the fusion a-particle. Any suggestions relating to large muon capture rates are still conjectures.
D. MUONICX-RAYSAFTER FUSION Should the muon be bound at any time to the a-particle, there will be some muonic x-ray transitions. These arise either from the initial population of the excited states or from the excitation of the ground state during the a-particle
205
MUON-CATALYZED FUSION
slowdown process. In either case the radiative transitions occur in competition with the other density-driven Coulombic processes and, hence, their observed intensity provides key supplementary information about sticking. The K-series x-ray yield per muon fusion, is obtained from the population probability Pj(t) of the (ap)+states, given the transition rates Arad:
A detailed investigation (Rafelski et al., 1989) of the dependence of the K X yield on the diverse phenomena controlling the history of the (ap)+-ion has shown that significant differences are only brought about through the choice of the Stark mixing in the L-shell. Turning off the mixing reduces the K , x-ray yields by 30%. On the other hand, significant modifications of the stopping power could impact the KX-yields. In particular, a reduction of the stopping power, which leads to greater reactivation (and hence smaller final sticking), increases the yield of muonic x-rays emitted after fusion. A qualitative expression for P, (Rafelski et al., 1989) shows that the K, x-ray yield due to excitations into the L-shell is
where the ratio of the K-L excitation cross section to the stopping power enters. Consequently, a (density-dependent) reduction of the stopping power enhances the KXm yield. This observation precludes any ad hoc manipulation of the stopping power of the (ap)+-ion with the goal of reducing the final sticking. This leads to a very similar proportional enhancement of the theoretical X-ray yield. Taking the present canonical sticking value of 0.92% to renormalize prior theoretical results on x-ray yields in accordance with Eq. (25), we arrive at the theoretical values shown in Table IX. It seems that the most sophisticated theoTABLE IX OF X-RAYSPER 100 d-f FUSIONS: XY, NUMBER
Experiment: Hartmann er a / . (1990) Nagamine (1990) Theory: Cohen (1988) Markushin (1988) Takahashi (1988) Rafelski et al. (1989) Stodden et al. (1990)
4,- 1 0.19 0.05 0.20 k 0.06 4, = 1.2 0.26 0.26 0.25 0.31 0.31
*
X
100 FOR DENSITIES 4, = p/po
qJ = 0.1 0.31 0.31
-
0.39 0.36
Note: Theoretical results are renormalized to 00 = 0.915%
4,=1 0.08
S
4,-1 0.02 ? 0.013
-
-
4, = 1.2 0.12 0.12 0.18 0.07 0.082
4, = 1.2 0.019 0.018 0.024 0.012 0.021
206
J . Rafelski and H . Rafelski
retical results for XY,, of Rafelski et al. (1989) and Stodden et al. (1990) are a factor of 1.5 too large compared with the experimental values (Hartmann et al., 1990; Nagamine, 1990). As this difference is most pronounced (up to errors in measurement) in the absolute yield, it may be taken as a further suggestion that the initial sticking is indeed smaller than expected. (See Section 1V.E.) E. EXPERIMENTAL VALUES FOR THE STICKING PROBABILITY After the experiment by Jones et al. (1986) many laboratories, including LAMPF, LNPI (Leningrad Nuclear Physics Institute), PSI (Paul Scherrer Institute, formerly SIN), Triumph, KEK (Japan), and RAL, pursued measurements of the total initial sticking fraction in either dd or dt fusion. Most of the USLAMPF &-fusion sticking fraction results were obtained considering the cycling rate of muons and measuring the rate of muon loss. Similar experiments have been performed with a somewhat different analysis of kinetics of the neutronemission and cycle dynamics at the PSI. Furthermore, the PSI experiment has been measuring the x-ray emissions from the MuCF cycle (Hartmann et al., 1990) from which the sticking probability can be extracted as well through a theoretical analysis. In a tritium-rich environment this is a very difficult experiment, as the natural triton decay generates a background just in the energy window of interest. Therefore, the PSI experiment was carried out at a very small tritium concentration, C, = 4 X This issue was taken up at the KEK laboratory at C, = 0.3, where a strong magnetic field was applied to limit the range of the @-electrons,and an intense pulsed muon beam allowed an enhancement of the signal-to-noise ratio. A summary of the experimental measurements of &' is given in Table X. While there is only a slight density-dependence arising from the regeneration phenomenon in the theoretical calculations using pure three-body Coulomb theory, there is a pronounced trend (within the error bars) for a decrease of sticking with density in the LAMPF data, which are just barely in agreement with the PSI neutron-based sticking data. However, the density-dependence in the LAMPF data could also be related to the fact that the high-density points were obtained by changing the temperature near or in the liquid-DT phase. (See Section 1V.E.) The interesting point is that these results are nearly half as small as the theoretical expectations. There is no doubt, in qualitative terms, about the experimental sticking being significantly smaller than the theoretical expectation, considering the enormous neutron yield reported per muon, which can be as high as 150 neutrons (Jones, 1986). Interestingly, the PSI and KEK x-raybased sticking is in much better agreement with the LAMPF neutron-based data than is the neutron-based PSI data. The result shown in Table X was extracted using the theoretical yield of K, x-rays per stuck muon: K J o , = 0.53. This agreement between neutron and x-ray data is significant, as the sticking probability measured by observing neutron yield suffers from the need to apply
207
MUON-CATALYZED FUSJON TABLE X THE( u p ) +STICKING FRACTION FOR DIFFERENT d-r TARGET DENSITIES 4 = p/pu
4 = 1.2 Experiment (from neutron detection): Jones efal. ( 1 986) Breunlich er al. (1987) Nagamine (1987) Bossy er al. Experiment (from X-raydetection K J o , = 0.53): Hartman et at. (1990) Nagamine (1990) Theory using o? = 0.915: (I-R) 04 (R: see Table VIII)
*
4
= 0.1
0.07 0.35 0.45 2 0.05 0.42 2 0.07 0.42 ? 0.14
1.1 2 0.5 0.50 f 0.10
0.36 0.09 0.38 f 0.11
*
-
0.59
0.65
-
corrections that account for the muon-loss probability per catalytic cycle from any cause. In particular, in a D-T environment we expect a certain small number of dd and dt fusions, which all have a significantly greater intrinsic sticking, simply because these reactions are less energetic, and the double charge reaction product is less energetic as compared with the dr case. Another important phenomenon is the capture or transfer of the muon to helium impurities, arising primarily from tritium decay. In either of these phenomena there can be a complex temperature- and density-dependence, and many of the involved branching ratios are still ill understood. Thus, most of the corrections are based on information and systematic behavior derived from differing experimental conditions and therefore there can be some difference about the magnitude of the individual effects and corrections to be applied. In particular, the improperly understood value and density-dependence of qIr(Section 1II.B) is a major source of some of the uncertainty, as once the muon reaches the ground state of dp, there is an appreciable probability that it will enter the dd-fusion cycle. All the available experimental data (as of 1990) for o,from LAMPF, PSI, and KEK are given in Fig. 2 along with theoretical curves. A weak theoretical-density-dependence arises from the regeneration factor. All experiments so far have yielded a sticking significantly below expectations, though we anticipated finding a greater-than-expected value (allowing for a yet unknown loss mechanism, etc.). This stimulated a new series of experiments that are presently in progress. LAMPF, in collaboration with RAL, has developed an apparatus in which coincidence measurements of a neutron in conjunction with an a + +or an (ap)+can be performed. Aside from the directness of the measurement, the other advantage of this project is that only a minor correction is required to account for stripping of the muon from an (ap) during the passage through the gas target and mylar window, as the muonic ion loses only a fraction of its energy. (Consider Eq. (20) for Ef= Eo.) LNPI has previously developed a direct method to measure sticking in the d-d reaction in a +
J . Rafelski and H . Rafelski
208
Final sticking fraction
1.0
I
I
1
I
4
I
t
I
1
1
,
-
-- q-3-e-:
0.8 -
-
1(
h
K
3"
-
- --
o.6gJ T
f-
1 1:
0.4
x
0.2 -
-
0 PSI 4
0.0
LAMPF
I
'
KEK Theory
'
" ' '
'
I
I
'
wire-ionization chamber. This group joined forces with PSI to carry out a direct measurement of sticking in a much-refined wire-ionization chamber. At Los Alamos, the intention is to study high-density, low-temperature sticking with the help of a novel target. A Berkeley-Livermore-PSI group is considering hightemperature, high-pressure targets.
F. DENSITY-DEPENDENCE OF STICKING There is a clear disagreement between predicted values of w, and the experimentally observed values at high density with which unusually high fusion yields (more than 150 neutrons per muon (Jones, 1986)) are associated. The source of this discrepancy could lie with any of the possibly misunderstood aspects of the catalyzed fusion cycle, which is density-dependent. But any density-dependence (beyond the trivial linear dependence of all two-body rates) requires either the competition between a two-body rate and a density-independent rate, typically a radiative transition or the competitive presence of three-body processes. A third alternative is to understand the density-dependence as a temperature-dependence in the range of 10 to 40K (see Section III.I), since the high-density points were obtained changing temperature and density at the same time. If the density-
209
MUON-CATALYZED FUSION
dependence of sticking is to be explained, its understanding must be related to the five known density-dependent processes in the MuCF cycle: 1. The probability to reach the ground state of dp (the so-called q,sprobability, Section 1II.B). Here radiative K,, K, transitions in dp compete with transfer rates dp+ tp. 2. The muon-recapture probability (the convoy effect). (See Section 1V.C.) 3. The probability of muon regeneration, where a small density-dependence arises from the competition of 2p- Is transition with excitation into higher orbits. (See Section 1V.B.) 4. The nonlinear density-dependence in electromuomolecule resonance formation. (See Section 1II.F.) 5 . The possibility of a small sticking side cycle of MuCF which dominates at high density and/or at low temperature. (See Section 111.1.) Other effects have been explored, e.g., a density-dependent stopping power (Rafelski er al., 1989). While a density-dependent reduction of the stopping power could explain the density-dependence of the effective sticking, the associated modification of the x-ray spectrum from transitions within the (Hep) -ion increases the already present discrepancy between theoretical predictions and experimental data. (See Section 1V.D.) +
V. Conclusion Of all the hydrogen-hydrogen MuCF reactions that have been studied since 1980, the d-t reaction has attracted by far the most theoretical and experimental attention. However, after 15 years of modern MuCF studies, central questions are still being debated, and there is today a clear perception that we are far from understanding muon-catalyzed processes, let alone deciding if MuCF will ever lead to any significant practical applications. One of the reasons for the recent surge of interest in MuCF has been the recognition of the important role diverse resonant reaction may play in speeding up the catalytic cycle. Until this was understood, the key bottleneck in the chain of reactions was the slowness of the in-flight (Auger) formation of the muomolecule which was believed to be a necessary step prior to a fusion reaction. Resonant molecular formation occurs when a neutral tp muo-atom enters a hydrogen molecule and binds extremely weakly to a deuteron, the muomolecular binding energy being picked up by the (second) vibrational band of the host electromolecule. The rate of formation of such resonances is highly temperaturedependent. After the initial experiments established temperature-dependence of the dtp cycle rate, much effort in the world has been devoted to establishing a reliable
210
J. Rafelski and H.Rafelski
value for the sticking. Most of LAMPF dt-fusion sticking fraction results were obtained by considering the cycling rate of muons and measuring the rate of muon loss. A different analysis of the kinetics of the neutron emission and cycle dynamics was performed at PSI, where the x-ray muo-atomic transitions from the MuCF cycle have also been measured. X-ray emissions have also been studied at KEK, in the presence of an intense magnetic field. The punch line of these experiments is the very low sticking of at LHD, with values of 0.35% (LAMPF) and 0.45% (PSI and KEK)-the results are within error bars of each other. The conventional theoretical high-density value is about 0.6%,which is about 1.5 times the experimental value. Furthermore, in theory there is only a slight density-dependence arising from the stripping, while in the experiment there is a pronounced trend (within the error bars) for an increase of sticking with decreasing density in the LAMPF data, a result that may be seen to disagree with the PSI data. A further disagreement at densities below LHD between LAMPF and PSI may be traced to differences in corrections applied to the raw sticking, which are required in the D-T environment due to a small number of dd and rt fusions which however have a significantly greater intrinsic sticking than dt fusion. From the point of view of applications, what matters is the uncorrected muon loss per cycle (excluding muon decay) W,, and on this quantity the groups nearly agree: muon loss per cycle is significantly below theoretical expectations. Should the muon become bound to the a-particle, there will be some muonic x-ray transitions (Section 1V.D). The theoretical x-ray yield is found to be 1.5 times greater than in the experiment, while the sticking, as we saw, is also a factor 1.5 times larger than in the experiment. Thus, there are two discrepancies between experiment and theory of profound importance and together they suggest that the initial sticking is smaller than expected, at least at high densities. But how small is the sticking really? All the surprising findings have stimulated a new series of experiments in progress as of 1990 that are to measure the initial sticking (LAMPF and RAL) and perform a direct measurement of sticking in a wire ionization chamber (LNPI and RAL). Extreme target conditions are also to be studied. The discussion of recent progress in understanding o,was one of the points emphasized in this survey. We have shown that the issue arises at least in part from a possible misunderstanding of the reaction chain leading to fusion, and in part from the theoretical oversimplification of the complex reaction processes. In our opinion the present data suggest the existence of a small sticking side cycle of d-t MuCF which becomes particularly active at high density and/or at low temperature. However, there are a number of other important results connected primarily with dt fusion that are not yet satisfactorily understood and for which a deeper understanding is required if the practicability of MuCF is ever to be properly assessed: 1. The observed nonlinear density dependence in the dtp molecular formation rate.
MUON-CATALYZED FUSION
21 1
2. The observed high probability qIsof the muon reaching the long-lived and inactive state (dp)],. 3. The low-temperature behavior of the cycling rate and, in particular, the molecular formation rate in D2 and DT molecules. 4. The observed temperature and molecular-hydrogen composition-dependence of the pdp fusion. Experiments as of 1990 arrive at the molecular rates during the process of complex analysis of the kinetics of the reaction-rate data and fusion yield obtained at varying conditions of density, isotope composition, molecular composition, and temperature with data taken as a function of time. The difficultiesjust noted may simply imply that our global understanding of the MuCF cycle of reactions is indeed incomplete, in the sense that parts of the MuCF cycle proceed via different chains of reactions than we allow for in the analysis. After an intensive search for such alternative paths we have reason to believe (Harley and Rafelski, 1990a, b) that direct in-flight reactions compete with the conventional molecular reaction picture. Many of the MuCF puzzles and foremost the sticking problem may have simply arisen from the ignorance of a possible important direct-reaction channel. A similar study could also help clarify peculiarities noted in the d and p d cycles. Should this explanation be correct, we must revisit in detail two areas of MuCF that have been rather quickly dismissed: 2 > 1 fusion and the tepid plasma environment (Section 1V.E). Even if dr-MuCF-based cost effective power production can be achieved, the dt reaction poses severe health and technological problems and requires a significant inventory of tritium in order to stop muons in a DT target. Ironically, the aneutronic and nontritium pd-reaction is suppressed in its rate by the relatively high Q-value, which suppresses the muon-conversion probability. Disregarding these concerns, which are similar to those voiced in other approaches to fusion, we note that in principle the required progress appears to be small in comparision to more conventional fusion approaches, such as plasma fusion or inertial confinement fusion. However, as highlighted in this chapter, it is generally believed that MuCF is already close to the limits posed by a combination of practical approaches with fundamental laws of physics, and hence further improvement is exceedingly difficult. But in principle a much greater number of fusions per muon will be possible than the present record of about 150, once all the atomic and molecular processes in the MuCF cycle are properly understood.
Acknowledgments Between 1982 and 1990, most of the progress in MuCF research has arisen directly from a coherent program of research sponsored by the Advanced Energy Projects Program of the U.S. Dept. of Energy under the leadership of its former
212
J . Rafelski and H . Rafelski
director, R. Gajewski, to whom this work is dedicated. This work was in part supported by the U.S. Dept. of Energy, BES/AdvancedEnergy Projects.
REFERENCES Adarnczak, A., and Melezhik, V. S . (1988). Muon Cardyzed Fusion 2, 131-136. Ad’yasevich, B. P., Antonenko, V. G. and Bragin, V. N. (1981). Sov. J. Nucf. Phys. 33,619-662. Aissing, G . , Bakalov, D. D., and Monkhorst, H. J. (1990). Phys. Rev. A42, 116-126. Alexander, S. A., and Monkhorst, H. 3. (1988). Phys. Rev. A38,26-32. Alvarez, L. W., Bradner, H., Crawford, F. S. Jr., Crawford, J. A,, Falk-Vairant, P., Good, M. L., Gow, J. D., Rosenfeld, A. H., Solmitz, F.,Stevenson, M. L., Ticho, H. K., and Tripp, R. D. (1957). Phys. Rev. 105, 1127-1128. Anderson, A. N. (1990). In “AIP Conf. Roc. 181: Muon-Catalyzed Fusion” (S. E. Jones, J. Rafelski, and H. J. Monkhorst, eds.), pp. 57-67. American Institute of Physics, New York. Aniol, K. A., Noble, A. J., Stanislaus, S., Virtue, C. J., Measday, D. F., Horvath, D., Jones, S . E., Robertson, B. C., and Salomon, M. (1988). Muon CuralyzedFusion 3,63-72. Aniol, K. A., Noble, A. J., Horvath, D., Robertson, B. C., Salomon, M., and Jones, S . E. (1989). In “AIP Conf. Proc. 181: Muon-Catalyzed Fusion” (S. E. Jones, J. Rafelski, and H.J. Monkhorst, eds.), pp. 68-78. American Institute of Physics, New York. Bakalov, D. (1988). Muon CurufyzedFusion 3,321-334. Balin, D. V., Vorobyov. A. A., Vorobyov, A. A., Zalite, Y. K., Maev, E. M., Medvedev, V. I., Semenchuk, G. G., and Smirenin, Y. V. (1984a). Phys. Letr. 141B, 173-176. Balin, D. V., Vorobyov, A. A., Vorobyov, A. A., Zalite, Y.K., Maev, E. M., Medvedev, V. I., Semenchuk, G. G., and Smirenin, Y. V. (1984b). JETP Leu. 40, 1112- 1114. Balin, D. V., Baturin, V. N., Chesmov, Y. A., Ilyin, A. I., Maev, E. M., Petrov, G. E., Petrov, L. B., Semenchuk, G. G., Smirenin, Y. V., Vorobyov, A. A., Vorobyov, A. A., and Voropaev, N. I. (1990). Investigationof temperadependence of the pCF-parameters in pure deuterium. pCF ’90International Conference on Muon Catalyzed Fusion, Vienna. Bogdanova, L. N., Markusbin, V. E., Melezhik, V. S., and Ponomarev, L. I. (1982a). Phys. Letr. 115B, 171-173. Bogdanova, L. N., Markushin, V. E., Melezhik, V. S., and Ponomarev, L. I. (1982b). Sov. Phys. JETP 56,931-935. Bogdanova, L. N., Markushin, V. E., Melezhik, V. S . , and Ponomarev, L. I. (1986). Erratum. Phys. Len. 1678,485. Bogdanova, L. N., Kuperin, Y. A., Kvitsinsky, A. A,, Markushin, V. E., Merkuriev. S . P., and Ponomarev, L. I. (1988). Muon Catalyzed Fusion 2,377-387. Bossy, H., Daniel, H., Hartmann, F. J., Neumann, W., Plendl, H. S . , Schmidt, G., von Egidy, T., Breunlich, W. H., Cargnelli, M., Kammel, P.. Marton, J., Nagele, N., Scrinzi, A., Werner, J., Zmeskal, J., and Petitjean, C. (1987). Phys. Rev. Lerr. 59, 2864-2867. Bracci, L., and Fiorentini, G. (1981). Nucl. Phys. A364, 383-407. Bracci, L., Chiccoli, C., Pasini, P., Fiorentini, G., Melezhik, V. S., and Wozniak, J. (1989). Phys. Lerr. A134,435-439. Breunlich, W. H., Cargnelli, M., Kammel, P., Marton, N., Nagele, N., Pawlek, P., Scrinzi, A,, Werner, J., Zmeskal, J., Bistirlich, J., Cowe, K. M., Justice, M., Kurck, J., Petitjean, C., Sherman, R. H., Bossy, H., Daniel, H., Hartmann, F. J., Neumann, W., and Schmidt, G. (1987). Phys. Rev. Leu. 58,329-332. Breunlich, W.R . , Kammel, P., Cohen, J. S . , and Leon, M. (1989). Ann. Rev. Purr. Sci. 39.3 11-355. Bystritskii, V. M., Dzhelepov, V. P., Petrukhin, V. I., Rudenko, A. I., Suvorov, V. M., Fil’chenkov, V. V., Chemnitz, G. M., Khovanskii, N. N., and Khomenko, B. A. (1976). Sov. Phys. JETP 44, 881-885.
213
MUON-CATALYZED FUSION
Bystritskii, V. M., Dzhelepov, V. P., Ershova, Z. V., Zinov, V. G., Kapyshev, V. K., MukhametGaleeva, S. S., Nadezhdin, V. S . , Rivkis, L. A., Rudenko, A. I., Satarov, V. I., Sergeeva, N. V., Somov, L. N., Stolupin, V. A., and Fil’chenkov, V. V. (1981). Sov. Phys. JETP 53, 877-884. Ceperley, D., and Alder, B. I. (1985). Phys. Rev. A31, 1999-2004. Chupka, W. A., Russel, M. E., and Rafeaev, K. (1968). J. Chem. Phys. 48, 1518-1527. Cohen, J. S. (1986). Phys. Rev. A33, 1437-1439. Cohen, J. S. (1988). Muon CaralyzedFusion 3,421-437. Cohen. J. S . , and Leon, M. (1986). Phys. Rev. A33, 1437. Cohen. J. S., Leon, M. and Padial, N. T. (1989). In “AIPConf. Roc. 181: Muon-Catalyzed Fusion” ( S . E. Jones, J. Rafelski, and H. J. Monkhorst, eds.), pp. 111-123. American Institute of Physics, New York. Danos, M., Miiller, B., and Rafelski, J. (1986). Phys. Rev. A34, 3642-3645. Danos, M., Stahlhofen, A. A , , and Biedenharn, C. (1989). Ann. Phys., NY 192, 158-203. Dzhelepov, V. P., Ermolov, P. F., Moskaler, V. I., and Fil’chenkov, V. V. (1966). Sov. Phys. JETP 23,820-831. Faifman, M. P., Menshikov, L. I., Ponomarev, L. I., Puzynin, I. V., Puzynina, T. P., and Strizh, T. A. (1986). Z. Phys. D2,79-85. Frank, F. C. (1947). Nurure 160,525-527. Froelich, P., Szalewicz, K., Monkhorst, H. I., Kolos, W., and Jeziorski, B. (1989). In “AIP Conf. Roc. 181: Muon-Catalyzed Fusion” ( S . E. Jones, J. Rafelski, and H. J. Monkhorst, eds.), pp. 259-273. American Institute of Physics, New York. Gershtein, S . S., and Ponomarev, L. I. (1977). Phys. Len. B72, 80-82. Gershtein, S. S., Petrov, Y. V., Ponomarev, L. I., Popov, N. P., Resnyakov, L. P., and Somov, L. N. (1981). JETP (Zh. Exp. Teo. Fiz.) 53, 872-876. Hale, G. M., Brown, R. E., and Jarmie, N. (1987). Phys. Rev. Lett. 59,763-766. Hale, G. M., Struensee, M. C., Pack, R. T., and Cohen, J. S. (1989). In “AIP Conf. Proc. 181: Muon-Catalyzed Fusion” (S. E. Jones, J. Rafelski, and H. J. Monkhorst, eds.), pp. 344-354. American Institute of Physics, New York. Harley, D., Miiller, B., and Rafelski, J. (1990a). J . Phys. G: Nucl. Phys. 16, 281-294. Harley, D., Miiller, B., and Rafelski, J. (1990b). Z. Phys. 336A, 303-312. Harley, D., Miiller, 8 . . and Rafelski, 1. (1990~).In “Proceedings of an International Symposium on Muon Catalyzed Fusion pCF-89” (J. D. Davies, ed.), pp. 56-58. Rutherford Appleton Laboratory, Oxon, UK. Harley, D., and Rafelski, J. (1990a). Pseudo-resonant d-t muon catalyzed fusion. Submitted for publication. Harley, D., and Rafelski, J. (1990b). Non-resonant fusion reactions in the tp d system. pCF ‘90 International Conference on Muon Catalyzed Fusion, Vienna. Harley, D. et al. (1991) “Muon Regeneration in degenerate plasma’’ (in preparation). Hartmann, F. J., Bossy, H., Daniel, H., von Egidy, T., Neumann, W., Plendl, H., Schott, W., and Weissgerber, P. (1988). Muon Curalyzed Fusion 2, 53-61. Hartmann, F. J., Baumann, P., Daniel, H., Grunewald, S.,Lipowsky, R., Moser, E., Neumann, W., Plendl, H. S . , Schmidt, G., Schott, W., von Egidy, T., Ackerbauer, P., Breunlich, W. H., Cargnelli, M., Fuchs, M., Jeitler, M., Kammel, P., Marton, J., Nagele. N., Werner, J., Zmeskal, J., Bossy, H., Crowe, K., Sherman. R. H., Lou, K., and Petitjean, C. (1990). The observation of quantum radiation as a tool for investigations on muon-catalyzed fusion processes. p C F ’90 International Conference on Muon Catalyzed Fusion, Vienna. Haywood, S. E., Monkhorst, H. J., and Szalewicz, K. (1988). Phys. Rev. A37,3393-3398. Haywood, S . E., Monkhorst, H.J., and Szalewicz, K. (1989). Phys. Rev. A39, 1634-1638. Haywood, S. E., Monkhorst, H. J., and Alexander, S. A. (1991) “Sticking fractions of rdfi and d d p using random tempered basis sets”, Phys. Rev. A, in press. Hu, C.-Y. (1986). Phys. Rev. AM, 2536-2539. Hu, C.-Y. (1989). Phys. Rev. A36,4135-4138.
+
214
J . Rafelski and H . Rafelski
Jackson, J. D. (1957). Phys. Rev. 106,330-339. Jandel, M., Danos, M., and Rafelski, J. (1988). Phvs. Rev. C37,403-406. Jandel, M., Froelich, P., Larson, G., and Stodden, C. D. (1989). Phys. Rev. A40, 2799-2802. Jeziorski, B., Moszynski, R., Szalewicz, K., Scrinzi, A., Zhao, X., Kolos, W., and Velenik, A. (1990). University of Delaware, Newark. Unpublished. Jones, S. E., Anderson, A. N., Caffrey, A. J., Walter, J. B., Watts, K. D., Bradbury, J. N., Gram, P. A. M., Leon, M., Maltrud, H. R., and Paciotti, M. A. (1983). Phys. Rev. Lett. 51, 17571760. Jones, S. E., Anderson, A. N., Caffrey, A. J., Van Siclen, C. D. W., Watts, K. D., Bradbury, J. N., Cohen, J. S., Gram, P. A. M., Leon, M., Maltrud, H. R., and Paciotti, M. A. (1986). Phys. Rev. Lett. 56, 588-591. Jones, S. E. (1986). Nature321, 127-133. Kamimura, M. (1989). In “AIP Conf. h c . 181: Muon-Catalyzed Fusion” (S. E. Jones, J. Rafelski, and H. J. Monkhorst, eds.), pp. 330-343. American Institute of Physics, New York. Kravtscov, A. V., Popov, N. P., and Solyakin, G . E. (1984). JETP Lett. 40, 124- 126. Kulsrud, R. M. (1989). In “AIP Conf. Proc. 181: Muon-Catalyzed Fusion” (S. E. Jones, J. Rafelski, and H. J. Monkhorst, eds.), pp. 367-380. American Institute of Physics, New York. Lane, A. M. (1983). Phys. Lett. 98A, 337-339. Leon (1984). Phys. Rev. Lett. 52, 605. Markushin, V.E. (1988). Muon Catalyzed Fusion 2,395-420. Men’shikov, L. I., and Ponomarev, L. 1. (1984). JETP Lett. 39,663-667. Men’shikov, L. I., and Ponomarev, L. I. (1985). JETP Lett. 42, 13- 16. Men’shikov, L. I., and Ponomarev, L. I. (1986). Phys. Lett. B167, 141-144. Men’shikov, L. I., and Ponomarev, L. I. (1987). JETPLetr. 46, 312-315. Men’shikov, L. I. (1988). Muon Catalytic Processes in a Dense Low-Temperature Plasma. Insritut Atomnoj Energii im. I.B. Kurchatova March 1988. Preprint IAE-458912, Moscow, USSR. Miiller, B., Rafelski, H. E., and Rafelski, J. (1989a). Phvs. Rev. A40, 2839-2842. Miiller, B., Rafelski, J., Jandel, M., and Jones, S. E. (1989b). In “AIP Conf. Proc. 181: MuonCatalyzed Fusion” (S. E. Jones, J. Rafelski, and H. J. Monkhorst, eds.), pp. 105-110. American Institute of Physics, New York. Miiller, B., Rafelski, J., Jandel, M . , and Jones, S. E. (1989b). 105-1 10. Nagarnine, K. (1987). Muon Catalyzed Fusion 1, 137- 150. Nagamine, K., Matsuzaki, T., Ishida, K., Watanabe, Y., Sakamoto, S., Miyake, Y., Nishiyama, K., Torikai, E., Kurihara, H., Kudo, H., Tanase, M., Kato, M., Kurosawa, K., Sugai, M., Fujie, M., and Umezawa, H. (1989). In “Proceedings of an International Symposium on Muon Catalyzed Fusion pCF-89” (J. D. Davies, ed.), pp. 27-32. Rutherford Appleton Laboratory, Oxon. Nagamine, K. (1990). X-ray observation of a-sticking phenomena in muon catalyzed fusion for liquid and high T concentration D-T mixture. FCF ‘90 International Conference on Muon Catalyzed Fusion, Vienna. Ostrovskii, V. N., and Ustimov, V. I. (1980). Sov. Phys. ZETP 52,620-625. Petitjean, C., Lou, K., Ackerbauer, P., Breunlich, W. H., Fuchs, M., Jeitler, M., Kammel, P., Marton, J., Nagele, N., Werner, J., Zmeskal, J., Bossy, H., Crowe, K. M., Sherman, R. H., Baumann, P., Daniel, H., von Egidy, T., Gruenewald, S., Hartmann, F. J., Lipowsy, R., Moser, E.,and Schott, W. (1990). The p p d fusion cycle. pCF ’90 International Conference on Muon Catalyzed Fusion, Vienna. Petrov, Y. V. (1980). Nature 285,466-468. Petrov, Y. V. (1985). Phys. Lett. 163B, 28-30. Petrov, V. Y., and Petrov, Y. V. (1989). In “AIP Conf. Proc. 181: Muon-Catalyzed Fusion” (S. E. Jones, J. Rafelski, and H. J. Monkhorst, eds.), pp. 124-144. American Institute of Physics, New York.
MUON-CATALYZED FUSION
215
Petrov, Y. V., Petrov, V. Y., and Shlyakhter, A. I. (1988). Muon CatalvzedFusion 2, 261-272. Rafelski, J. (1980). Hydrogenic mesomolecules and muon catalyzed fusion. Exotic Atoms '79, Vol. 4 (K.Crowe et al., eds.), pp. 177-205. Plenum, New York. Rafelski, H. E., Miiller, B., Rafelski, J., Trautmann, D., Viollier, R. D., and Danos, M. (1987). Muon Catalyzed Fusion I, 315-332. Rafelski, H. E., Muller, B., Rafelski, J., Trautmann, D., and Viollier, R. D. (1989). Progr. Part. and Nucl. Phys. 22,219-338. Rafelski, H. E., and Muller, B. (1990). Comp. J . Phys. Comm. 59, 521-525. Rafelski, J., and Miiller, B. (1985). Phys. Lerr. 1648, 223-227. Rafelski, H. E., Harley, D., Shin, G . R.. and Rafelski, J. (1991). J. Phys. B: Arom. Mol. Opt. Phys. 24, 1469-1517. Sakharov, A. D. (1948). Report of the Physics Insritute. Academy of Sciences, Unpublished, referred to by Zel'dovich (1954). Scrinzi, A., Szalewicz, K., and Monkhorst. H. I. (1988). Phvs. Rev. A37, 2270-2276. Scrinzi, A., and Szalewicz, K.(1989). Phvs. Rev. 39, 2855-2861. Shin, G. R., and Rafelski, J. (1990a). Nucl. Instr. Methods A287, 565-569. Shin, G. R.. and Rafelski, J. (1990b). J. Phvs. G: Nucl. Part. Phvs. 16, L187-Ll95. Shin, G. R., and Rafelski, J. (1990~).Phvs.Rev. A43, 601-602. Soff, G., and Rafelski, 1. (1990). Z. Phvs. D14, 187- 190. Stodden, C. D., Monkhorst, H. J., Szalewicz, K., and Winter, T. G . (1990). Phys. Rev. A41, 128I - 1292. Struensee, M. C., Hale, G. M., Pack, R. T., and Cohen, J. S. (1988). Phvs. Rev. A37, 340-348. Swe Myint, K., Akaishi, Y.,Tanaka, H., Kamimura, M., and Narumi, H. (1989). 2. Phvs. A334 423-428. Szalewicz, K.,Monkhorst, H. J., Kolos, W.. Scrinzi, A. (1987). Phvs. Rev. A36, 5494-5499. Takahashi, H., Kouts. H. J. C., Grand, P., Powell, J. R . , and Steinberg, M. (1980). AtomkernenergieKerntechnik 36, 195- 199. Takahashi, H. (1988). Muon Caralvzed Fusion 2,453-458. Vesman, E . A. (1967). JETP Lett. 5,91-93. Vinitskii, S. I., Ponomarev, L. I . , and Faifman, M. P. (1982). Sov. Phys. JETP 55, 578-581. Zel'dovich, Y. B. (1954). Dokl. Akad. Nauk S. S. R. 95, 493-498. Zmeskal, J . , Kammel, P., Scrinzi, A, Breunlich, W.H., Cargnelli, M.. Marton, I . , Nagele, N., and Werner, J. (1990). Phys. Rev. A42, 1165-1177.
This Page Intentionally Left Blank
.
.
.
ADVANCES IN ATOMIC MOLECULAR AND OPTICAL PHYSICS VOL . 29
MULTIPLE-ELECTRON EXCITATION. IONIZATfON. AND TRANSFER IN HIGH-VELOCITY ATOMIC AND MOLECULAR COLLISIONS J . H. McGUIRE Department of Physics Kansas State University Manhattan. Kansas
I . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . I1. Theory . . . . . . . . . . . . . . . . . . . . . . . . . . .
A . Formulation . . . . . . . . . . . . . . . . . . . . . . . B . Independent-Electron Approximation . . . . . . . . . . . . . C . Multiple-Electron Effects (Electron Correlation) . . . . . . . . . D . ExpansioninZp/v . . . . . . . . . . . . . . . . . . . . . E . Many-Body Perturbation Expansions in Both Zp/vand u . . . . . . F. Special Features of Electron Capture, Nonorthogonality,and Special Kinematics in Many-Body Scattering . . . . . . . . . . . . . G . Projectile Electrons and a Classical Atomic Limit . . . . . . . . H . Molecules and the Transformation Between a(B) andf(Q) . . . . . I . Methods of Computation for Few- and Many-Electron Transitions . . I11. Observations and Analysis . . . . . . . . . . . . . . . . . . . A. Simple Analysis . . . . . . . . . . . . . . . . . . . . . . B . Multiple Ionization. . . . . . . . . . . . . . . . . . . . . C. Multiple Excitation . . . . . . . . . . . . . . . . . . . . D. Multiple-Electron Capture . . . . . . . . . . . . . . . . . . E. Ionization and Excitation . . . . . . . . . . . . . . . . . . F. Transfer and Ionization . . . . . . . . . . . . . . . . . . . G . Transfer and Excitation . . . . . . . . . . . . . . . . . . . H . Projectile Electrons . . . . . . . . . . . . . . . . . . . . I . Molecules . . . . . . . . . . . . . . . . . . . . . . . . IV. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . . Appendix: Correlation . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . .
. . . . . .
. . .
. . . . . .
. . . . . .
. . . . . . . . .
217 219 219 226 230 232 237 239 244 247 250 263 264 270 286 288 291 292 297 299 302 312 313
314 315
.
I Introduction The many-body and many-electron problem is common in various areas of physics as well as in chemistry and biology. Basic understanding of phenomena ranging from the nature of matter at the creation of time to the properties of useful 217
.
Copyright Q 1992 by Academic Press Inc. A11 rights of reproduction in any form reserved. ISBN 0-12-003829-3
218
McGuire
materials in the human environment is limited by the boundaries of our knowledge of the many-body problem. There is an advantage in studying the manybody problem in atomic physics since the two-body and parts of the three-body problem are understood. Furthermore, both the mystery of the meanings of quantum mechanics and the mystery of the transition from microscopic time-reversible atomic processes to the dynamics of macroscopic time-irreversible aggregates of atomic particles is inherent in the many-body problems of atomic interactions. Thus, by studying the many-body problem in atomic physics we are able to develop effective tools to discover insights that provide both meaning and utility in our lives. The few-body problem in atomic physics is not completely understood. Although our knowledge is growing (Stolterfoht, 1989, 1990; Reading and Ford, 1987a,b; Briggs and Macek, 1991; McGuire, 1987; McGuire and Straton, 1990a,b), understanding is limited in a number of key areas. One limit of understanding, for example, is encountered in collisions of atomic hydrogen with charged particles at moderately high velocities (Madison, 1990; Walters, 1988; Williams, 1986). While there is good agreement between theory and experiment for total and differential cross sections, relative phase information observed in electron impact excitation to the first excitation state of atomic hydrogen differs even in sign at large scattering angles from present calculations done on some of the largest computers now available. Another limitation in the Coulomb fewbody problem is the asymptotic wavefunction for three charged particles needed for calculations of ionization of atoms, ions, and molecules by charged particles (Brauner et al., 1989). Despite these limitations, some reasonable understanding of multielectron processes such as multiple-electron ionization is still possible under certain conditions. For example, if electron correlation is small, then it may be possible to apply the independent-electron approximation to describe total cross sections, even when a sum over many final states is included (McGuire and Macdonald, 1975; McGuire and Weaver, 1977; Sidorovitch and Nikolaev, 1983). As more details about the final states are specified, it is often necessary to include more and sometimes new details (e.g., many-electron effects) in the analysis of atomic and molecular reactions. While looking closely at atomic interactions tends to require new understanding of many-particle effects for individual atoms, this many-body understanding is also useful in dealing with aggregates of atoms. The many-electron effects needed to understand specific detailed cross sections in many-electron atoms and molecules may also be required to understand the transition from microscopic atomic properties to properties of macroscopic systems of atoms. Specifically, we may get a clearer notion of correlation. In studies of both microscopic and macroscopic effects (Goldberger and Watson, 1964; Balescu, 1975) the notion of correlation is an important concept. In a general sense correlation is the interdependency of individual members of a group or, perhaps more simply, how individuals affect one another. This concept of correlation is used in both atomic physics and statistical mechanics. (Cf.Appendix.) In these contexts correlation
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
2 19
usually refers to electron correlation which is caused by the Coulomb interaction (or correlation potential) acting between electrons, although other kinds of correlation (e.g., time [memory] or spin) are possible. While a physical definition of correlation could be based on the correlation interaction between electrons, the more common definition is mathematical (Balescu, 1975; Huang, 1987; Van Kampen, 1981; Eadie et al., 197 1 ; Gardiner, 1985). Mathematically correlation is that which is not described by a product of terms. For example, a wavefunction is correlated if "(r,, rz)# $(r,)+(rJ.The physical and mathematical definitions of correlation differ a little. In this chapter the mathematical definition of correlation is used and some, but not much, distinction is made between correlation and multielectron effects. The focus of this chapter is multielectron transitions in high-velocity atomic (and molecular) transitions, i.e., multiple ionization, electron transfer, excitation, and combinations thereof. Multiple-electron transitions are intrinsically many-body (many-electron)phenomena in contrast to single-electron transitions which may or may not depend significantly on multielectron effects. At high collision velocities the description of both single and multiple transitions is relatively simple both conceptually and mathematically. The emphasis here is on multielectron effects in single atoms and molecules, although brief attention is given to an example of a transition to a classical limit of quantum mechanics in a few-electron system that is observable. A little consideration is given to some statistical methods. Auger transitions and dielectronic recombination are largely ignored. Unless otherwise specified the projectile is considered to be a fastcharged particle that is much heavier than an electron, and the target is a neutral atom that is not strongly correlated. This means that the projectile moves faster than the target electrons and that the independent-electron approximation can be sensibly defined. (Cf.Fig. 11 .) Where possible different processes are analyzed in a unified way. For example, some connection is made among collisions of atoms (and molecules) with photons, electrons, positrons, protons, antiprotons, fully stripped ions, partially stripped ions, and neutral atoms.
II. Theory A. FORMULATION The Hamiltonian for scattering of an atom of nuclear charge Z, with N electrons by a particle of charge Z, and mass M using atomic units is (McGuire, 1987)
= K
+ V + Ho,
(2)
220
McCuire
where M,is the reduced mass,
K = --V2 2M,'
(3)
and
-cv,+ N
- N)
zP(zT
j= I
R
where long-range Coulomb effects vanish when 2, = N and
Here R is the internuclear coordinate and rJ the coordinate to the jth electron. The (rt - rj - I Coulomb (electron-correlation) interaction gives rise to spatial electron-electron correlation in the unperturbed (static) atomic Hamiltonian H,. It is now assumed that the internuclear motion may be separated from the electron motion and that the internuclear motion may be treated classically, so that the internuclear trajectory R(t) is well defined, e.g., R(t) = B vt where B is the impact parameter of the projectile. The resulting equation (McGuire and Weaver, 1986) for the electron motion is now time-dependent, namely,
+
The Hamiltonian for the electrons, H e 1 ,defined as H,1 = Ho
+
i
V, = H,
+V-
-
zF'(zT
R
N)
(7)
is now explicitly time-dependent since the V, depend on R which is explicitly time-dependent. Consequently, the evolution operator U is not simply given by exp[ - iH,, ( t - t o ) ] ,but rather by a more complicated expression containing time ordering, namely, EQ.(10) which follows. It is useful to work in the intermediate representation, where one takes full advantage of the fact that the eigenfunctions of H, are known (or nearly known). In the intermediate representation, the evolution operator U ( t , t o ) is governed by
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
22 1
V,(t) is not a sum of single-electron operators because H,, in Eq. ( 5 ) is not a sum of single-electron terms due to the Irk - r,l- electron-electron interaction. Equation (8) may be formally solved for U ( t , f a ) (with T as the time-ordering operator):
namely, Dyson’s equation (Goldberger and Watson, 1964). The probability amplitude for scattering from the asymptotic initial state I#~to the asymptotic final state 4fis given by a =
(4fllCli) = (+flu( +
- m)l6i).
~7
( 1 1)
The probability, P(B),for a transition from 4, to r$f is given by the absolute square of a, and the corresponding cross section by
mif =
I
dB P(B) =
I
ds luI2.
(12)
This result holds for an arbitrary number N of electrons, including the fundamental case of N = 1. Some special considerations are as follows.
1 . Mulfielectron Eflects and Electron Correlation Electron correlation is a multielectron effect. That is, electron correlation arises because there is more than one electron in our system and the electrons affect one another. Correlation (discussed in the Appendix) is rigorously defined as the difference between exact and uncorrelated quantities. A quantity is uncorrelated if that quantity is determined from a sum of single-electron Hamiltonians, even though that quantity obeys the Pauli exclusion principle and appropriate symmetry requirements such as rotational symmetry (conserving angular momentum) and parity. For example, the total wavefunction, the probability amplitude, and the probability are uncorrelated if (ignoring symmetries) q ( r , , r z ) = $(rl)+(r2),a(1, 2) = a(l)a(2) and P(1, 2) = P(l)P(2), respectively. Note, however, that u( 1, 2) f a(l)a(2) due to the integration over impact parameters in Eq. (12). As the Hartree-Fock approximation is usually defined to be the uncorrelated static wavefunction, we use the time-dependent Hartree-Fock (TDHF)
222
McGuire
approximation (Devi and Garcia, 1984; Stich et al., 1985;Gramlich et al., 1986) to define the uncorrelated dynamic wavefunction. In practice this requires defining a new basis set of wavefunctions using the variational principle every time the system changes significantly during the collision (which can be difficult). Also, for highly correlated systems the uncorrelated limit is sometimes not easily defined (e.g., there may be no sensible single-electron configuration). In this chapter we shall primarily consider systems where correlation is not very strong so that the uncorrelated (independent-electron) limit is well defined.
2. Scattering, Relaxation, and Asymptotic Regions In a scattering event the boundary between the scattering region and the asymptotic region is not easily localized. In this chapter we shall assume that the interaction usually occurs within about one Bohr radius of the target nucleus. At high velocities the collision time is usually shorter than the relaxation time required for the final state to decay, usually by either X-ray decay or Auger emission. So the excitation and de-excitation process can be decoupled in most cases we consider. The correlation that occurs outside the scattering and relaxation regions (in both space and time) is referred to as atomic, static, or asymptotic correlation. The Auger effect corresponds to a de-excitation process in which the electron-electron interaction causes a metastable doubly excited state to decay with one electron going to a lower state and the other to a higher state. This is sometimes referred to as dynamic correlation (Stolterfoht, 1989). Correlation occurring during the collision is called scattering correlation.
3 . Waves and Particles
Quantum amplitudes may be formulated in either the wave picture using the scattering amplitude, f,or in the particle picture using the probability amplitude, a. As shown by McCarroll and Salin (1966) and demonstrated explicitly in Section ILH., the amplitudes f and a are related by a Fourier transform, namely,
f(Q) =
I
dB eiQsa(B)
where Q is the momentum transfer, p is the reduced mass, v the velocity of the projectile, and B is the impact parameter of the projectile. Heref = 0 if there is no scattering (e.g., if the target does not interact with the projectile) so that U is replaced by U - I in Eq. (1 1). In this chapter both f and a are used according to convenience of presentation. The duality of these quantum mechanically equivalent representations presents a difficulty, however, when correlation is
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
223
considered. Uncorrelated probability amplitudes do not usually correspond to uncorrelated scattering amplitudes since a product of a’s does not transform by Eq. (13) to a product off’s. In studying correlation we shall generally use the particle picture so that uncorrelated means that the electron probability amplitude is expressed in product form. We shall consider in Sections 1I.G and III.1.2.d examples where the scattering amplitude is in product form. Since this amplitude arises from the electron-electron interaction, we are reluctant to say that electron correlation is not present in this case. A resolution of this dilemma is suggested by Balescu (1975), who suggests the use of Wigner distributions. In most experiments the wavepacket of the projectile is much larger than the size of the target. Under these conditions it is always possible to determine the scattering angle accurately, but it is not always possible to determine the impact parameter accurately. The conditions under which there is a well-defined one-toone relationship between the impact parameter and the scattering angle of the projectile (namely 2v/ZpZT< 1, where v is the collision velocity) were established by Bohr in 1948. Sometimes these conditions are not met in the examples we consider, and then the impact parameter, B, is considered as a mathematical rather than a physical parameter.
4. Interaction Representation In dealing with probability amplitudes we generally use the interaction representation, the projectile is localized so the trajectory is well defined (often well approximated by a straight line), and the influence of the target electrons on this trajectory is neglected (i.e., the trajectory is determined by elastic scattering from the nucleus) as discussed by McGuire and Weaver (1986). In principle one could work in another representation such as the Heisenberg representation. In practice this is difficult because of the coupling between the projectile and the target expressed (Magnus, 1954) as an infinite series of commutator terms in an exponential. Without these terms the exp{iwt} term (e.g., Eq. (40)) is missing in the first-order amplitude in an expansion in Z , and the resulting cross sections become infinite as with the Magnus results of Eichler (1977). 5 . Long-Range Coulomb Terms
The long-range nature of the Coulomb interaction causes some difficulties. The total two-body Coulomb cross section is infinite. The kernel of the LippmannSchwinger equation is not compact so that there is no guarantee of uniqueness of the solution and corresponding numerical difficulties arise (Joachain, 1975; Merkuriev, 1976; Chen and Chen, 1972). The on-shell limit of the off-energy-
224
McGuire
shell Coulomb T matrix does not exist due to a phase divergence (Gau and Macek, 1975; Roberts, 1985). Fortunately, the long-range Coulomb terms are almost shielded in nature at large distances. For scattering from neutral atoms the long-range problems may be avoided by formulating scattering (McGuire and Weaver, 1986) in terms of the difference between the instantaneous Coulomb interaction and the average of this Coulomb interaction over the initial- (or final-) state electronic densities of the target. This difference of terms (similar to Eq. (4))eliminates the long-range Coulomb tail in interactions of a charged projectile with electrons of a target atom. In the case of neutral targets, the Coulomb interaction between the projectile and the target nucleus is replaced by the nonCoulombic static potential between the projectile and the entire atomic target with a frozen density distribution of target electrons. Thus unnecessary asymptotic Coulomb terms are removed, e.g., in the case of p + H + H + p where asymptotic Coulomb terms are not required. This also provides a starting point for density functional theory mentioned in Section II.1 where the electron density is no longer frozen. For scattering of charged particles with ions in this formulation, the electronic interactions remain non-Coulombic (i.e., short range) and scattering from the nucleus involves (if necessary) only two-body Coulomb terms that are well understood. This formulation also removes to first order the difficulty arising if nonorthogonal initial and final states are used. The longstanding problem of the asymptotic wavefunction for three charged particles is not, however, solved by this formulation. In many cases, however, there is an advantage in retaining long-range Coulomb terms since mathematical techniques have been developed that are useful in analytic evaluation of matrix elements. A useful formulation of scattering amplitudes in terms of Coulomb wavefunctions, for example, has been developed by Belkic et al. (1979).
6. Exclusive and Inclusive Cross Sections In a rigorous quantum mechanical calculation a theorist may evaluate a probability amplitude a, for a transition from a known n. Finally we note that it is not necessary to use the same value of P for all electrons in a given shell. Different electrons may have different values of P (e.g., due to different binding energies). In this case a multinomial distribution (Eadie et al., 1971; Aberg, 1987) may be used, corresponding to
where Zy=, Pi = 1. In this distribution the final state is specified more completely than in the binomial distribution of Eq. (83). Both binomial and multinomial distributions are applicable in the independent distinguishable electron approximation (Section I1.B.; Section I1I.B. 1). Binomial distributions apply if the single-electron transition probabilities, P, are the same (or nearly the same) for a group (or shell) of N electrons. Overall energy conservation is not applied. For (heavy or very fast light) projectiles the energy loss of the projectile is usually negligible in contrast to the statistical energy deposition model considered next. b. Statistical Energy Deposition Model. In the statistical energy deposition (SED) model developed by Russek (1963), it is assumed that energy, ET, obtained by an atomic (or molecular) target during a collision is statistically distributed to all of the N electrons in the system. One simply counts the number of ways that ET can be distributed among the N equivalent electrons. If any electron
256
McGuire
has an energy greater than the ionization energy, E,,, then it is ionized. P; is the probability function for ionizing n of the N electrons. The following mathematical development is largely due to Tulkki ( 1987). According to Russek (1963) P;(m) =
(3
j=O
K,,(j)Q&-"(M- rn - j)lK,,(m)
(90)
gives the probability that when m units of energy are statistically distributed among N electrons, n and only n electrons will receive r units of energy each allowing their ionization. In Eq. (90)
where [vlh] is the integer part of ulh. The task is to determine P;(ETIEion)= lim P',")(m)
(93)
m--
such that x = mlr = ET/Eio,stays finite. Tulkki observed that the summation over j can be carried out exactly. Substituting Eqs. (91) and (92) in Eq. (90) and observing from Eq. (91) that j S m - rn - ri gives
where
c (I+4- l ) ( m
m - m - ri
L',h3(m,r; i) =
-
j=O
rn
+
- j - ir N - n - 1 m - m - j - ir (95)
Using the abbreviations a = n - 1 b = m - rn - ir N c = N - n - 1 +a.
+
- n - l + a
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
257
Eq. (95) can be written as b-c
b + l
(97)
c - a
j-0
This sum rule is given by Vilenkin (1971). Thus Eq. (94) reduces to
(3[(mz)’r’; ~l,i(~
n ) ( m - rn
i=O
- ir N - 1
+N
-
1
PN(m) =
(98)
which is exact. Now let m and r go to infinity:
[ ( m - m)/r]+ [x - n](x
-
n - 1 < i,,,
x
- n)
for all i except for i = x - n which gives a negligible contribution if it occurs. Using the formula
+w
v fixed
in Eq. (98) yields the Russek formula
x = ETIEi,
where k must be chosen such that n
+k
S x< n
+k+
1.
(102)
Equation (101) implies that the energy has been shared statistically between the N - n electrons that remain and the n electrons that are ejected. Thus, it corresponds to a distribution where the ions are left in excited states. Since decay of excited states often occurs by autoionization, the SED distribution is not usually observed experimentally. Hence, justification of this simple method is incomplete (Aberg, el al., 1984). The binomial P;;(B) and the SED PE;(AE,) are compared in Figs. 7 and 8. The impact parameter B and energy transfer AEm may in some cases be related
N
FIG.7. The probability P i for removing n of N electrons from a target for N = 8. The binomial distribution corresponding to the independent distinguishable electron approximation is shown on the left for various values of n as a function of the impact parameter.
2M
400 600 AE (eV) FIG.8. The probability Pg for removing i of N electrons for N = 8 in the statistical energy deposition model (Mueller ef al.. 1985) is shown on the right as a function of the energy transfer AE to the system. See Fig. 15 for a comparison of these models to observations.
0
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
259
classically (Cocke, 1979). In any case the binomial and SED P,! are not the same. For example, the maximum values of Pg differ for 0 # n # N . For N = 6, one has Max P; n
IEA
SED
1 2 3
40% 33% 31%
70% 59% 43.5%
The SED distribution tend to more successful at low collision velocities, while the binomial distributions are often successful at high collision velocities. ( ' C Section 1II.B.1 .) In chemistry the basic concept of SED has been modified by using quantum transition probabilities (e.g., harmonic oscillator probabilities for vibrational transitions in molecules). This modifies the Pg probability functions. This procedure (Rosenstock et al., 1952) goes by the name of quasi-equilibrium theory. (Cf.Sec. 111.1.1.) c . Fokker-Planck Method. The Fokker-Planck method (Balescu, 1975) is a statistical method, widely used in nuclear physics, for analyzing distributions of events as a function of time. This method may also be applied to atomic collisions. In an ion-atom collision the atomic charge can change during a collision according to a probability distribution, W ( N , t), which changes during the time of the collision, 0 < t < T. Initially the number of electrons N = No is well defined, corresponding to W ( N , t < 0 ) = 6 ( N - N o ) . During the collision the width of the distribution, r, increases. When the collision is over and the systems have readjusted, W ( N , t > T ) no longer changes. This is the basic picture that directly corresponds to observed data in some cases.
(i) Fokker-Planck Equation. The equation governing the probability distribution W ( N , t ) is the Fokker-Planck equation, dW(N' ') = [ V ( N , t)W(N, r)] dt dv
61
+[D(N, i)W(N, t)]. dN'
(103)
This is a diffusion equation where V is a velocity coefficient and D is a diffusion coefficie5t. This equation is the same as the Schrodinger equation i dqtdr = H ( - V, V z ) 9 with time rotated by 90" in the complex plane. Equation (103) corresponds to Eq. (92) of Balescu who gives a clear derivation and explanation, with W = W , , N = y, V = A, and D = 112B. The basic assumption is that the process is a Markov process: a statistical process where in each step infor-
260
McGuire
mation regarding previous steps is lost (i.e., the memory time is short). The first term on the right-hand side of Eq. (103) corresponds to a frictional slowing corresponding, as is later illustrated, to the fact that the system is losing electrons. The second term corresponds to diffusion, or the time rate of change of the width in Fig. 9. Note that if V and D go to zero, then dWldt = 0 and W remains frozen in time. Mean values of physical observables may be determined from W, e.g.,
r
(N") =
J cw N"W(N, r)
(moments of N )
,-
(E) =
J cw E,W(N,
r)
( E is energy, for example).
We may also note the following mathematical properties:
I
W ( N , r) 3 0
CW W(N;t )
= 1
(nonnegative probability distribution)
(105)
(conservation of total probability).
Finally, V and D are related to the first two moments (corresponding to a linear second-order partial differential equation) via
d v=(N) dt
( 106)
(ii) Application to Atomic Collisions. The statistical distribution function W may be applied to atomic collisions by relating V and D to atomic collision quantities. There are various ways in which this may be done. For example, one may introduce transition rates and Fermi-Dirac distributions for colliding atomic systems, as has been done in nuclear collisions. An alternate method, used here,
OctcT
0 N NO FIG. 9. The Fokker-Planck distribution function W ( N , t ) before, during, and after a collision occurring during at time T.
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
26 1
is to consider one shell of an atom that is losing electrons (e.g., via direct Coulomb ionization) with a single-electron probability, p, that is the same for all electrons within that atomic shell. With this picture in mind, one may express V and D in terms of the singleelectron transition probabilities, P ( B , t) by using the definitions of V and D from Eqs. (106) and (107). Noting that n N = No where n = number of vacancies, N = number of electrons, and No = total number of states in that atomic shell, it is easily shown that
+
d
(N) dt d dt
=
-d
(n) and dt d dt
- {(N2) - (N)2} = - {(n’)
-
(n)2}.
If we use, for example, a binomial distribution for the probability of producing n vacancies from N electrons, namely P f = (:)pn(l - p)N--n, it is straightforward to show that
where p = P(B, r) is a single-electronprobability for an atomic transition. Then
D
=
d +-{(n’ dt
-
dP (n)2)}= N-(1 dt
- 2p).
Now V and D are known functions of N and t. They may be used in the FokkerPlanck equation, Eq. (103), to find W(N, t). Note that after the collision, V and D go to zero, so that W(N, t > T) does not change. Next let us consider some limitations of this method. The Fokker-Planck equation assumes that a Markov process is valid, i.e., the system forgets its history before the collision. This implies that the collisions are well separated in space and time (e.g., Eq. (73) holds). From the atomic point of view, each shell is treated independently, each B is treated independently, and p(r) may be computed accurately. Also, electrons only leave and do not return. (This can be corrected by a more complicated model with particles moving back and forth, as has been done in nuclear physics.) Since the Einstein relation is not used, W does not go over to a Maxwellian distribution. (C.’ Balescu, 1975. This may not always be a limitation.) It is also required that No >> 1, so that Jp dN W(N, t) = J; dN W(N, t) or Jg0 CW W(N, r) T, W ( n , t > T ) = W ( N , T ) . The behavior is illustrated in Fig. 10. This corresponds to taking p ( t ) and dpldt as step functions is time. Also, Vo and Do are independent of N . Thus take
Note that if Po
FIG. 14. Comparisons of multiple-ionization data with the statistical energy deposition model (Mueller et al., 1985). Here the energy deposited in the system, A&, is given in units of the ionization energy I , .
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
273
FESIDUAL CHARGE STATE FIG. 15. Comparisonsof multiple-ionizationdata to binomial distribution (IEA) and to statistical energy deposition (SED model) (Cocke, 1979).
data and analysis of Cocke (1979) shown in Fig. 15. Generally, however, the independent-electron approximation is usually used at moderately high collision velocities while the statistical energy deposition model is used for low-velocity collisions. Multiple ionization by electron impact (Hazi, 1981) has been reviewed by Schartner (1990). Collisions of highly charged ions at low velocities have been reviewed by Barany (1990).
2. Double ionization Two-electron transitions are the simplest of the multielectron transitions that are the subject of this chapter. Because helium and molecular hydrogen are two
274
McGuire
of the simplest targets in which to study two-electron transitions, these targets have been more thoroughly considered than more complicated systems. a. Independent-Electron Approximation. Application of the independent distinguishable electron approximation to analysis of data for single and double ionization of helium is shown in Figs. 16 and 17 (Sidorovitch et al., 1985) which does not include the relatively accurate data of Shah and Gilbody (1985). At the lower velocities shown, first-order perturbation theory used for the singleelectron probability, P, is inadequate in both single and double ionization. The independent-electron approximation agrees with single ionization within about 40% at higher velocities and misses the data for double ionization by a little
SINGLE IONIZATION 10-1
=
10-16
10-l~
100
1000 E ( KeV/arnu) FIG. 16. Cross sections for the single-electron ionization of He by the H+,He2+,and Liz+nuclei versus nuclei energy (Sidorovitch et al., 1985). The various calcuations are all based on perturbation theory.
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
275
more than twice the error in single ionization as one might expect since double ionization varies as P2 compared with single ionization which varies as P (for small P). Unless one uses different binding energies per electron for single and double ionization it is difficult to fit both data sets better than shown in Figs. 16 and 17. For highly charged ions ( Z , up to 44) incident at 1.4 MeV/amu on helium, the Giessen group observed a saturation in the increase in both single and double ionization with increasing Z , shown in Fig. 18. Salin (1987) has interpreted this saturation with the independent-electron approximation using a multiple expansion defined on one center that limits the single-electron probability to values less than one. A simple cutoff of first Born results gives similar results (Ben-Itzhak and McGuire, 1988). Thus, while the overall trends of the
I
I
DOUBLE I ONlZATlON
E (KeV/amu)
FIG. 17. Cross sections for the two-electron ionization of He atoms by the H + , He*, and Li3+ nuclei versus nuclei energy (Sidorovitch er al., 1985). Theoretical curves are all based on perturbation theory.
276
McGuire
r'7
-r--IEA - wder V 1st
in
Incident ton Chorpc
, Zp
FIG. 18. Ratio of single to double-ionization cross sections in helium versus projectile charge, Zp (McGuire et al., 1987). The curve labeled IEA is a calculation of the independent-electronapproximation. The curve labeled 1st order in V is the observed high-energy limit of the ratio.
data are generally qualitatively reasonable (except for the very high velocity data for proton impact explained in terms of shake-like models in Sections III.A.3 and 1II.B.b.-e), it is difficult to obtain accuracy better than a factor of two using the independent-electron approximation in the case of helium. b. Simple Shake. At sufficiently high collision velocities it is expected that effects first order in Z,,/v will dominate. (C' Sections II.D.l and III.A.3.) The simplest picture in this regime is the simple shake picture in which double ionization follows single ionization via final state rearrangement independent of the properties of the projectile. In this simple picture the ratio of double to single ionization is expected to be independent of Z , or v. Ratios of double to single ionization were first considered by the Aarhus group in 1982. Figure 19 shows that this ratio does go to a constant value independent of both v and 2, at high v as qualitatively predicted by the simple shake model. The observed ratio of 2.67 X for 40 MeV electrons (80GeV/amu) from Giessen (Salzborn and Mueller, 1986) nicely confirms this trend. Comparable results for H, observed by Edwards er al. (1988) are shown in Fig. 20.
MULTIPLE-ELECTRONEXCITATION, IONIZATION, AND TRANSFER
r
0.5
\
!.
I
mi mm
2
5
I
1
10
277
20 I
He
'J 0.5
E, (MeV) 5
2
I
10
20
FIG. 19. Ratio of single- to double-ionization cross section in helium versus collision velocity (Haugen et al., 1982). The curve labeled IEA is a calculation of the independent-electron approximation. The curve labeled 1st order is the observed high-energy limit of the ratio. The top energy scale is for electrons and the bottom scale is for protons.
7
6 - I H'on H2 i e- on H2 5-
,Q 4 -
0
3-
-
2?
, € €
I-
0
1
I
I
,
,
,
McGuire
278
c. Relation between Charged Particles and Photons. In the simple shake theory (Mittleman, 1966; Amusia et al., 1975; Sampson, 1990) the ratio of double to single ionization is the same for impact by charged particles and photons. As pointed out by Eckhardt and Schartner (1983) this prediction is not in agreement with observed data. The ratio of total cross sections for double to single ionization for charged particles differs by an order of magnitude from the ratio for photons. This difference can be largely explained by the fact that photons impart all of their energy to the ionized electron while charged particles impart a distribution of mostly relatively small energies. From Eq. (68) without the inelastic projectile electron terms (or from McDowell and Coleman, 1970, Chapter 7) the scattering amplitudes may at high v be expressed as
The amplitude for photon impact may be expressed in terms of
f = (4f1e-E 220
(+[
IE
*
PI 44
*
PI
+i)
-
(+f
IE rl *
(photons)
( 130)
+i)
where K is the momentum of the photon (K = w / c ) . Here E is the polarization vector of the photon and p = iV. Below about 1 keV the value of Kmi. is small enough to use the dipole approximation. Since the matrix elements are the same up to overall factors, the cross sections for ionization (or dipole-allowed excitation) by charged particles and photons are related at small momentum transfer as shown by Byron and Joachain (1967), namely
where E is the energy of the ejected electron, and E is the energy of the incident particles. For charged particles, there is a distribution, p ( E ) , of ejected electrons ~ each , ~electron ~ energy E correwhile for photon impact E = E ~ - ~I , i.e., sponds to a single photon energy. Consequently, if we take a ratio of double- to single-ionization cross sections
where R = cr++/c+ for both Z and y. Thus, the ratio for ionization-charged
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
PHOTON ENERGY, E
279
lev)
FIG. 21. Energy distribution, p t , of the final state (left-hand side) and ratio, R,, of single to double photoionization (right-hand side) in helium as a function of photon energy, E (McGuire, 1984). Here E = I E where I(&) is the binding (ejected-electron) energy and A is the difference in the single- and double-ionization threshold energies. Note that I p$ d~ = 1, and that R, is equal to the area under the product of the curves illustrated here. The data (Schmidt et al., 1976; Wright and Van der Wiel, 1967; Holland er al.. 1979; Carlson, 1967) are taken from McGuire (1984). A useful datum at 600 eV due to Carlson (1967) is included in the next figure.
+
particles is equal to the ratio for photons weighted by density of ejected electrons and integrated over the ejected-electron energies. This relation illustrated in Fig. 21 was first suggested by Hvelplund (1982). (Cf.McGuire, 1984.) Consequently, the ratios of total cross sections for total double to single ionization can be quite different for photons and charged particles, but are nonetheless related at high particle velocities. Numerical calculations (McGuire, 1984) are in qualitative agreement with the observed results in Table 2.
d . Aberg's Generalized Shake Probability. A conceptually useful generalized shakeoff probability has been introduced by Aberg (1976) who treats the first outgoing electron as a plane wave and ignores antisymmetrization of the wave functons, i.e., the exchange terms are ignored. Following Aberg, let state & b e an uncorrelated antisymmetrized wave function with an outgoing plane wave electron of momentum k and a residual electron wave function Cbf. (For double ionization one might simply use a plane wave of momentum k2 for 4f.) Thus,
280
McGuire
For a two-electron system the scattering amplitude is given by
+
I
- 4TZp -{F(Q
-
v5Q
I
1
dr, +r(r,)eiQrl drz e-ik.r2+i(r,,r 2 )
k)
+ exchange term}.
If we ignore the exchange term, then the matrix element for a shake process to a particular final state cbf is expressed in terms of generalized shake probability amplitudes, Ff(Q - k). It is easily shown that f di(Q - k) (+f I + i ) + (+f )eiQrll &(k) where +i(p)is a momentum space wave function and + i ( r i ,r z ) = +1(rl)+i(r2). Also it is easily shown that the exchange term (i.e., the second term) may be ignored when Q - k = 0 for large Q and large k. Taking the ratio of cross sections for double to single ionization or excitation, and ignoring the exchange term one may obtain Aberg's generalized shake result, namely,
-
-
P A , = N ( Q - k) =
PAbe,g(Q
I
I
dr,ei(Q-k).ri dr2 +f(r2)+i(rL,r2)
(135)
- k).
where N(Q - k) is a normalization factor given explicitly by Aberg. Here Q is the momentum transfer of the projectile, k the electron momentum. This probability is a function of the momentum Q - k, transferred to the residual ion. At a fixed Q - k this probability is independent of how the state f was formed. Thus, in this approximation shake processes for photons and charged particles, capture, and Compton scattering may be interrelated. e. Shakeoff Observation. Aberg's generalized shake probability suggests that shakeoff may be different for photons than for fast ejected electrons in binary encounter collisions with charged particles since Q - k is different. For photons Q = 0 so that the residual momentum transfer of the target is k, which can be large. For binary encounters with charged particles Q = k so that Q - k = 0. The experimental results of Kamber er al. (1988) for binary encounters with charged particles combined with photon results are shown in Fig. 22 with data from Schmidt et al. (1976), Wright and Van der Weil (1976), and Holland et al. (1979). In the electron velocity range shown, k = 15. A discontinuity from the photon results to the charged particle results is evident indicating that the shake
28 1
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
L
6.0
d,k(Carter
8,Kelly
50 -
&?
v
40-
+m
5
30-
m
I
20 -
i 4 ’
1.0 -
I
0 01
p’ data
(Q=k) , , ,b, , ,
,
01
,I
1.0
1
0
I
I ! *
10.0
PHOTON ENERGY (keV) FIG. 22. Ratio of double to single ionization in helium in equivalent photon energy. Data and theory are taken from Kamber et al. (1988). Note the discrepancy between Q = 0 data for e - and y and Q = k data for p + . The Q = 0 data corresponds to the data in Fig. 21 except that the open triangles in Fig. 21 are omitted and a useful observation of Carlson (1967) has been included at 600 eV. The high-energy limit for photoionization is not well determined experimentally.
limit does depend on Q - k. The physical picture is that shake is different if the collision is a close encounter (large Q) or a distant encounter (small Q). A larger change in screening could occur for collisions at close encounters than for distant encounters. This suggests that there is not just one simple shake limit, but rather a range of high Zdv limits for the ratio of double to single ionization. (C’ Section II.ZA.4.) However, more data would be helpful including more complete data for photoionization.
f. MBPT Calculations. The limit of high ejected-electron velocities for highvelocity collisions may be one of the simplest limits in which multielectron effects dominate. It is clear from the previous discussion that this limit is not well understood as of 1991. The range of ratios with Q is not clear, and the interrelation (if any) of various processes is also not entirely clear. (CJ Section III.A.4.) Many-body perturbation theory calculations done by Ishihara and Hino (1989) indicate that the shake concept may not be sufficient to describe this limit. In addition to contributions from the many-body perturbation theory first-order shake diagram, there are also contributions for photoionization from other firstorder diagrams (Cf.Section 1I.E) as shown in Fig. 23. Similar calculations for impact of charged particles have not yet been done. The observations of Kamber et al. (1988) as well as Aberg’s calculation (discussed in Section III.B.2d) sug-
282
E (eV) 23. MBPT calculation for ratio of double to single photoionization versus photon energy, E (Ishihara and Hino, 1989). The total result includes amplitudes for ground state correlation (GSC), two step-l (TSI). and shakeoff (SO) illustrated in Fig. 1. Calculations using each MBPT amplitude alone are also shown. No one amplitude dominates at high energies. FIG.
gest that the limit of high ejected-electron velocities and high collision velocities depends on the momentum transfer, but neither complete calculations nor reliable qualitative understanding is available. g. Zp?Terms. The ratio of double- to single-total-ionization cross sections compiled by Haugen (Hangen et a/., 1982) in Fig. 19 presented at the first European Conference on Atomic Physics in 1981 posed an interesting problem. In the velocity regime considered single-ionization cross sections are well described by the first Born approximation which varies as Zp?and gives no significant difference between cross sections for ionization by protons or electrons. At such high velocities apparently well within the region of validity of perturbation theory it was difficult to understand the clear experimental factor of two in the difference between double ionization by protons and electrons. Data confirming this effect in Hz taken by Edwards et al. (1988) is shown in Fig. 20. In 1982 McGuire suggested that the observed difference between protons and electrons was due to Z; interference arising from the square of a sum of firstand second-order amplitudes in the absence of multielectron effects. (C’. Section 11.D.4.) The second-order amplitude includes effects of the independent-electron
283
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
approximation. It was counterargued that the observed difference could be due to trajectory effects caused by the large difference in the masses of the protons and electron. An experiment of Edwards et al. (1986) showed that protons and deuterons gave the same ratios. This data provided some evidence against the argument for a mass effect. The issue of charge versus mass was resolved by an experiment by Andersen er al. (1986, 1987) using protons and antiprotons in LEAR at CERN. Their data, shown in Fig. 24, show that for double ionization the antiproton data lie above the proton data by a factor of about two, confirming a clear 2, dependence in double but not single ionization. 10-16
I?I'-lL84
0
PROTON DATA
o
ANTIPROTON DATA
-
10-l~
N
1420-
0
I
I
I
I
I
I
I
1
2
3
4
5
6
7
8
E (MeV) FIG.24. Observed cross sections for single and double ionization of helium by protons and antiprotons (Andersen ef af.. 1986, 1987).
284
McGuire
As of early 1991, theorists tend to agree that the difference between protons and antiprotons (and electrons) is due to correlation. However, there is much disagreement as to the physical mechanism for correlation. McGuire's suggestion of interference of quantum amplitudes is challenged by Becker (1985; McGuire and Burgdoerfer, 1987) who argues that the interference may be restricted by selection rules. Olson (1987) suggests that the effects may be explained classically without quantum interference. Vegh (1988) suggests that polarization of the charge cloud of the target may be especially important in double ionization. The relatively complete forced impulse method C (' Section 11.1.1) calculations of Reading and Ford (1987a,b) indicate that correlation must occur between at least two collisions with the projectile. Sorensen suggests (Andersen er al., 1987) that correlation can occur after one collision with the projectile when one electron hits another. Meanwhile data for positron impact by Charlton er al. (1988) shown in Fig. 25 nicely completes observations for total cross sections. Threshold effects at lower
0.1
0.2
0.5
1
2
5
10
20 "80,000
E [MeV arnu-'1
FIG. 25. Observed ratios of double to single ionization of helium by proton ( p + ) ,antiprotons ( p - ) , electrons ( e - ) , and positrons ( e + )of the same incident velocity in MeV/amu. Data taken from Charlton er al. (1988). The 80-GeVlamu electron point is from Salzborn and Mueller (1986). The differences between negative and positive projectives above one MeV/amu are thought to be caused by correlation, but the physical mechanism for this correlation is not clear.
MULTIPLE-ELECTRONEXCITATION, IONIZATION, AND TRANSFER 100
i
a x
n
V
A
+
285
ELECTRON PROTON NITROGEN HELIUM
10
z 1
1
0
10
20
30
40
50
ENERGY (MeV/amu)
FIG. 26. Comparison of He double-to-single-ionization cross section ratios (Heber et al.,1990). The solid curve shows the predictions of a semiempirical calculation of Knudsen er al. (1984) and the dotted curve shows the predictionsof Reading and Ford. The high energy limit for N'' projectiles appears to differ from the lower Z projectiles.
velocities are evident for the projectiles with a small mass. The Z 3 effects are related to a difference in stopping powers observed for positive and negative pions by Barkas (Barkas et al., 1956; Basbas, 1982; Andersen et al., 1989). Clear conceptual understanding, however, of collision mechanisms for the 2; terms remains a challenge.
h. High v Limit for Large Z,. Heber et al. ( 1990) observed the ratio of total cross sections for double to single ionization by heavy ions with Z, = 7 at values of Zp/v comparable to protons at 5-10 MeV. Heber et al. observed a ratio that differs from the ratio observed for projectiles with Z , = 1 (C'. Table LI) and with the empirical fit in Zplv obtained by the Aarhus group (Knudsen et al., 1984) using electrons and some ions with Z , 2 1. Heber et d ' s results shown in Fig. 26 are in possible conflict with present theory which agrees with the fit of the Aarhus group. (Also see Andersen et al., 1990.) The high u limit is not well established for large Z, . 3. Differential Cross Sections It is evident from the preceding discussion that observations of multielectron transitions are likely to be useful in developing understanding of few- and manybody effects. In principle differential cross sections (Rudd et al., 1985) can yield more detailed information. Unfortunately, relatively few differential studies of multielectron transitions at high collision velocities are available. One differential cross section that has been observed is the ratio of double to
McCuire single ionization of helium by protons as a function of the scattering angle. The observed data of Giese and Horsdal (1988) show a peak at 8 = 28, where 8, is the scattering angle for a single binary encounter. Vegh (1989) first proposed a three-step model including a double binary encounter between an electron and the projectile. A more complete classical calculation by Olson et al. (1989a) (including a double binary encounter events) gives qualitative agreement with the data. Single- and double-ionization cross sections triply differential in the angle and energy emission of the electron and in the recoil angle of the projectile have been observed by Skogvall and Schweitz (1990). For double ionization differences are found with coupled channel calculations indicating a breakdown of the independent-electron approximation. Another set of cross sections differential in the recoil angles and energies of the ejected electrons have been observed by Schmidt-Bocking and analyzed by Olson (Olson et al., 1989a). These observations show that a large amount of recoil momentum can be carried by both electrons and the recoiling nucleus. The topic of recoil distribution has been reviewed by Cocke and Olson (1990). C. MULTIPLE EXCITATION Understanding multiple excitation uses many of the ideas and methods applied in the previous subsection to multiple ionization. However, multiple excitation is easier to understand than multiple ionization where the asymptotic Coulomb wavefunctions are not known. Furthermore, one may search for effects of selection rules more directly in multiple excitation. On the other hand, observations of multiple excitation tend to be more difficult than for multiple ionizations so that only a little data for double excitation is available as of 1991. 1. Fano Profiles
It is common to measure double excitation in helium by observing resonances in either photon or electron emission. The resonances are imbedded in a continuum background as illustrated in the case of Auger electron emission in Fig. 27. Analysis of these resonant (Fano) profiles has been discussed by various authors (Fano, 1961; Mehlhorn, 1978, 1985; Shore and Menzel, 1968; Stolterfoht, 1971; Bordenave-Montesquieu and Benoit-Cattin, 1971; Prost et al., 1977; Bordenave-Montesquieu et al., 1982; Arcuni and Schneider, 1987; Gundnov et al. 1989). At a given energy and angle of the scattered electron the autoionization resonance and the continuum background must be added coherently. Fano (1961) briefly notes after Eq. (110) in his paper that one may unitarity to separate the resonance peak from the continuum background, i.e., using J l'I'E)('PE[ = J (14)(41 IJIE)(+EI) = 1. Here a sum is taken over the energy and direction of the ejected particles. Under this condition (i.e., summing over
+
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
33
35
287
37
e- ENERGY (eV)
FIG. 27. The fits of Fano profiles to the doubly differential emission cross sections for (a) electrons and (b) protons (Giese et al., 1990). The dashed line is the total fit, the dotted line includes contributions of only 2 p resonances, and the dashed-dotted line includes only the 2s2p resonances in the main peak.
both energies and angles of the ejected particles) the resonance is decoupled from the background of continuum electrons if the continuum background is constant. This decoupling does not hold, however, if the background varies with energy and the q parameter (Fano, 1961) is not larger than one.
2. Double Excitation Cross sections for double excitation of helium into the first excited states have been reported by Pedersen and Hvelplund (1989) and by Giese et al. (1990), although other results were available earlier (e.g., Rudd, 1956; Bruch et al. 1984; Stolterfoht, 1985). The more recent results are shown in Fig. 28 together with various calculations. Calculations of Fritsch and Lin (1990) are based on the method of coupled channels (Section 11.1.1). Preliminary results (not shown) of Reading and Ford using the forced impulse method (Section 11.1.1) are close to those of Fritsch and Lin for the (2p2) ID transition. McGuire and Straton (1990b) have done calculations complete through second order in ZJv (Section 1I.D). The theoretical differences are relatively large. Differences between Zp =
288
McGuire
0.1 0.0
'
'
"
"
'
'
'
"
I.o
-.-. ._ '
'
, . - ,
'
'
'
2.0
Projectile Energy (MeV/amu) FIG. 28. Cross section for double excitation of helium to 2s2p('P) (McGuire and Straton, 1990b). Experimental data due to Giese et al. (1990) and Pedersen and Hvelplund (1989). FL is the result of a coupled channel calculation of Fritsch and Lin (1990). Interference with the singleionization continuum background is not included in the theory.
+
1 and Z , = - 1 appear to be smaller than the factor of two found for double ionization. This means that the effects of space and time correlation that give rise to Z; effects (C' Section II.D.4) may be relatively small for double excitation to the first excited level. Observations of excitation of both electrons in A+16 in collisions with He at high collision velocities have also been made (Chetioui, 1990), but no theoretical calculations have been done as of early 1991. Double excitation of helium by photon impact has been considered by Lindle et al. (1985) who quote a constant high energy calculated ratio of double to single excitation into all states of n = 2 of about 4.8% in rough agreement with observation. The connection of ratios of double to single exctiation by various projectiles (e.g., charged particle and photons) has not yet been considered. (See Table 11.) D. MULTIPLE-ELECTRON CAPTURE Multiple capture, like multiple excitation, may be regarded as transitions to discrete bound states. However, as noted in Section KF, capture differs at high velocity from excitation both kinematically (e.g., the Thomas process) and in its scaling laws with projectile charge and velocity.
I . Multielectron Capture Capture of as many as six electrons in a single fast collision has been discussed by Schlachter et al. (1990). The observed data are well fit by a binomial distri-
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
289
0.3 0.2 0.1 0.0
0
1
2
3
4
5
6
7
0
1
2
3
4
5
8
7
0.3 0.2 0.1 0.0
L- and M-&W electrons ~aptwed FIG. 29. Typical fit of data (solid) to binomial distribution (hatched) of probabilities given by the independent-electron approximation (Schlachter er al., 1990) for Ca+” Ar at 47 MeV.
+
bution using a single capture probability, P, as shown in Fig. 29, corresponding to the independent-electron approximation (Section 1I.B). Similarity to the binomial analysis for ionization in Fig. III.B.2 may be noted. Binomial distributions have been previously used for multiple capture by Anholt et al. (1987) including capture in relativistic heavy-ion collisions.
2 . Double Capture Data for total cross sections for single- and double-electron capture for He++ He from Sidorovitch et al. (1985) are shown in Figs. 30 and 3 1 together with calculations in the independent-electron approximation (Theisen and McGuire, 1979) using the Bates-Born approximation. (C’ Section 1I.F.) At high collision velocities both single and double capture go primarily to the ground state. As in the case of single and double ionization (Figs. 1II.B. 5 and 6) the error for double capture is about twice that for single capture, and the independent-electron approximation appears to be somewhat incomplete. More recent calculations using a continuum distorted wave approximation in an independent-electron model using correlated asymptotic wave functions by Crothers and Dunseath (1990) give some improvement in the comparison of theory to observation. Coupled channel calculations at intermediate energies by Jain et al. (1989) are in fair agreement with data by Zouros er al. (1987) for formation of 2 t 2 t ’ double excited projectile states for He++ + He. As the collision velocity decreases, an
+
290
McCuire
SINGLE CAPTURE
10-22
-
10-23
too
1000 E (KeV/amu)
FIG. 30. Cross sections for the single-electron capture to the ground state in the charge transfer of the H + , He2, and Li3+ nuclei in helium versus nuclei energy (Sidorovitch et al., (1985). The theoretical curves are based on the first-order Bates-Born approximation.
increasing fraction of multiple capture goes into excited states. A brief overview of multiple capture at lower velocities, including an examination of simultaneous vs. successive transitions, is given by Barat (1988). In the high-velocity limit it has been suggested that multiple capture may be dominated by shakeover (Section III.A.3; Miraglia and Gravielle, 1987; McGuire et af., 1988). Gravielle and Miraglia (1990) have developed a two-step model for double capture including shakeover. No experimental evidence for shakeover is available as of early 1991.
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
29 1
E (KeWarnu) FIG. 31. Cross sections far the two-electron capture to the ground state in the charge transfer of the H', He2, and Li3+nuclei in helium versus nuclei energy (Sidorovitch er al., 1985). The theoretical curves are based on the first-order Bates-Born approximation.
E.
IONIZATION AND
EXCITATION
Relatively little work has been done on ionization and excitation at high veIocities. A few observations for ionization and excitation in He and Ne in collisions with various ions have been made by Bmch et al. (1979) Comparison of ionization plus excitation in H, by protons and by electrons has been made by Edwards (1990). The only calculation was done by Kocbach (Bruch et al., 1985) using the independent electron approximation with first-order single-electron
McGuire
292
0
u
V
9
m
H
on H e e on H e
+
" 3
1
2
3
4
5
6
7
8
9
Velocity [v/v,]
FIG. 32. Cross sections for excitation plus ionization of helium by impact of protons and electrons (Bruch et al. 1990).
probabilities for excitation and ionization. Qualitative agreement between these calculations and data for C4+ + He at velocities from 167 to 417 keV/amu. Data of Bruch et al. (1990) comparing ionization plus excitation of helium in collisions with protons and electrons, shown in Fig. 32, is consistent with observations of Pedersen and Folkmann (1990). A review is given by Hippler (1988).
F. TRANSFER AND IONIZATION At high collision velocities most studies have been of single and single-electron transfer (Salzborn, 1990). An overview of transfer ionization in helium is shown in Fig. 33 developed by Mueller (McGuire et al., 1987; Hvelplund el al., 1980). In this figure the ratio of transfer ionization to single transfer is plotted over a wide range of projectile energies for projectiles of charge Z , ranging from 1 to 3. At the higher velocities shown this ratio is consistent (HorsdalPedersen and Larsen, 1985; Kristensen and Horsdal, 1990) with a combination of independent-electron approximation plus shakeoff, i .e. the probability for transfer ionization is approximated by (Pion+ Pshkeofl) Pumrfer. (Cf. Section III.F.3.) The value of PshkeoE is approximated by the high-energy photon limit R, = 3.5% since the transferred electron leaves the collision region with a ve-
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
r-
293
II
Energy
(KeVIamu)
Fic. 33. Ratio of ionization plus capture to single-capturecross section in helium versus velocity (McGuire er nl., 1987). Note approximate similarity of limiting value of R at high energy with the limit for photoionization shown in Fig. 21.
locity equal to the velocity of the projectile, which is large. The data for proton impact is anomalous in that the observed ratio shows no peak at intermediate velocities (i.e., about a few hundred keV/amu). Also, studies of transfer ionization in collisions with helium with U+q and I + *about several hundred keV/amu have been reported by Datz et al. (1990), who suggest that electron correlation may be significant due to the large projectile charge which promotes correlated double capture. At lower collision velocities transfer ionization is indistinguishable from double-electron transfer followed by autoionization. At these lower velocities double transfer rates are expected to be greater than direct ionization plus single capture.
1. Independent-Electron Approximation Comparison of a calculation (Sidorovitch et al., 1985) using the independentelectron approximation with data is shown in Fig. 34. Since the single-electron probabilities are computed using first-order perturbation theory which overestimates the true result at low velocities and capture to excited states is ignored (also a source of significant error at lower velocities), the theory is not reliable at the lower velocities. At the higher velocities the agreement of these calculations with observation is consistent with the error in single capture and single ionization. (C’ Sections 1II.B and D.)
McGuire
294
10-16
10-23 10- 24 10- 25
E (KeV/amu) FIG. 34. Cross sections for the ionization of helium with simultaneous electron capture to the ground state of nuclei in collisions of He atoms with the H+,He2+,and Li’ nuclei versus nuclei energy (Sidorovitch er nl., 1985).
2. Thomas Ridge Since electron capture is classically forbidden in a one-step process (cf. Section ILF), classically allowed two-step processes may play an important role in differential cross sections that include electron transfer. To illustrate the nature of second-order ridges, the distribution of recoil particles, daldv, , given by a second Born calculation (Briggs and Taulbjerg, 1979; Ishihara and McGuire, 1988; McGuire et al., 1989) for the four-body reaction p+ He -+ H He++ + e - is shown in Fig. 35. Here M , = M, and
+
+
MULTIPLE-ELECTRON EXCITATION, IONIZATION. AND TRANSFER
295
M' = M 2 = M 3 = mc so that r = 1 in Fig. 3 and v3 is the velocity of the ionized (recoil) electron. In this case the He nuclear charge Z , localizes the two electrons but plays no direct role in the transfer process. The presence of this fourth particle, however, broadens the constraint of overall three-body energy and momentum conservation. A sharp ridge corresponding to overall energy and momentum conservation, and a broad ridge cresting v3 = v are both evident for 200-MeV proton impact shown in Fig. 35. The broad ridge corresponds to a pole in the Green's function. The crest of this ridge occurs when the intermediate energy equals the total energy. The classical Thomas peak is found at the intersection of these two ridges. The widths of both ridges are consistent with the uncertainty principle. The broad energy-conserving ridge has a width P E = Z,V' x hla,,, corresponding to the discussion in Section 1I.F. with v' = v = v3. The sharp ridge has a width given by the momentum distribution of the target electron which is proportional to Z , . This width goes to zero as ZT+ 0. In the ridge shown in Fig. 35 one target electron collides with another in the second step of a two-step collision (somewhat like the TSl process in double ionization of Section IILB). Because of the interactions between the electrons, this ridge corresponds to a multi-electron effect. Observation of this Thomas singularity was reported by Palinkas et al. (1989, 1990) and their data are shown in Fig. 36 together with the original approximate second Born calculations of Briggs and Taulbjerg (1979) as well as more recent
Recoil Angle, 7
+ He + H + He+ + e - for 200-MeV p+ (Ishihara and McGuire, 1988). Here vois the Bohr velocity. The sharp ridge, which permits energy nonconservation in intermediate states, intersects the broad energy-conserving ridge at u, = v and y = 90". (Cf. Fig. 34.) At lower projectile energies both ridges are broader and weaker. The broad peak is indistinct at energies near or below a few MeV where experiments are feasible. FIG.35. Differential cross sections duldv, in the velocity of the recoil electron for p+ +
296
McGuire
t
102*
t . 60
90
120
Emission angle [deg.] FIG. 36. Observations in one-MeV P + He by Palinkas el al. (1989, 1990) of the Thomas peak due to electron-electron rescattering.The solid line is the theory of Briggs and Taulbjerg (1979) and the dotted line of Ishihara and McGuire (1988) corresponding to a cut at u = u,, in the ridge in Fig. 35 (for a 1-MeV proton).
calculations of Ishihara and McGuire (1988). Here v3 is fixed at vj = v and the emission angle y is varied. Data suggesting this effect were reported by Horsdal et al. in 1986. At the collision velocities observed the second Born calculations shown in Fig. 35 are inaccurate (and show little structure). A different mechanism for the observed peak has been proposed by Gayet and Salin (1990) who use a diffraction mechanism without electron correlation. Their calculations give qualitative agreement with the observation of Horsdal et al. shown in Fig. 37. 3. High-Velocity Limit In the limit of high collision velocity, one might use Ptmnsre, PsMeofffor the probability for transfer ionization. Then the ratio of total cross sections for transis independent of the fer ionization to single transfer, proportional to Pshakeoff, collision velocity, v (Sections III.A.3 and 4).The data of Kamber et al. (1988) (Section ILI.B.3) then suggest that this ratio depends on Q - k, the momentum transfer to the residual ion, consistent with the generalized shakeoff probability
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
0.15
-
0.10
-
0.05
-
297
+-
B ++
B
v
\
+
x
0.00
1
2
A
Scattering angle (mrad)
FIG. 37. Ratio of transfer ionization to single transfer. The theory, diffraction model of Gayer and Salin (1990). is one of various possible explanations. Data are from Horsdal er al. (1986).
of Aberg (Eq. (135)). For transfer ionization IQ - kl lies halfway between the value of k for photoionization and zero for ionization via a binary encounter between the projectile and a target electron. Assuming that the ratio varies monotonically with Q - k, one may expect that the ratio for transfer ionization to be between the ratio for photoionization (large )Q - kl) and the ratio for ionization via binary encounters with charged projectiles (small lQ - kl). Data for this ratio (Kristensen and Horsdal, 1990, listed in Table 11) is consistent with this prediction, but the data also appear to increase with projectile velocities. An alternative possibility for this high v limit of the ratio of transfer ionization to single transfer is suggested using the simple shakeover probability of Eq. (127). In the preceding discussion it was assumed that at very high v, the transfer ionization probability is PmSfer Psh&eoB. If shakeover (collapse of the electronic wavefunction onto the moving projectile) occurs, then the dominant transfer ionization probability is given by Pionnation Psh&eover. Using this one has a ratio of transfer ionization to single capture that increases linearly with v at in the high v limit. As noted already, observations (Kristensen and Horsdal, 1990) are inconclusive.
G . TRANSFER AND EXCITATION The two-electron transition in which a target electron is transferred and a projectile electron is excited has received considerable attention experimentally stimulated by an observation in 1982 by Tanis et al., who discovered a resonance in
298
McGuire
the total cross section. The observed resonance was interpreted as an inverse Auger process (or dielectronic recombination) in which the projectile electron is excited by interaction with a captured target electron. This process occurs when the kinetic energy of the projectile electron matches the transition energy, i.e., a resonant condition in the collision velocity. This process has been referred to as resonant transfer and excitation (RTE). In some cases RTE dominates the total cross section for transfer excitation. In 1983 Pepmiller et al. observed a nonresonant process for transfer and excitation occur due to independent interactions of the two electrons with the nuclear charges Z , and 2, of the projectile and target, respectively. Thus, RTE is a process with electron correlation and NTE is an uncorrelated process. Theoretically Shakeshaft and Spruch ( 1979) had briefly suggested the possibility of RTE in 1979. The first calculation of RTE was done by Brandt (1983) using quasi-free electrons in the impulse approximation (Goldberger and Watson, 1964, Chapter 11). In Brandt’s approximation the RTE cross section is equal to the cross section for dielectronic recombination weighted with a velocity distribution for the target electron, namely,
where J is the probability for finding a target with a velocity component v, along the beam axis, i.e., J is the Compton profile (Richard, 1990). In 1984 Feagin er al. formulated an amplitude for transfer excitation that includes both RTE and NTE, namely, =
acornelated
+
aunconelatcd
(137)
-
where auncomtated - aexcitation atrangrer, corresponding to NTE, and acornlared is an amplitude for RTE proportional to the electron interaction that includes Brandt’s RTE result. However, the relative phase between the RTE and NTE amplitudes may be difficult to obtain correctly. Coupled channel calculations for He+ H at intermediate velocities (Fritsch and Lin, 1988) are in good agreement with observations. The experimental technique of zero-degree Auger spectroscopy has been developed by Stolterfoht and coworkers (Itoh er al., 1985; Swenson et al., 1986; Stolterfoht et al. 1987). The Fano profiles of angular distributions of Auger electrons and photons have been considered by Bhala (1990) and Kunikeev and Senashenko (1989). Hahn (1990) has discussed the intercorrelation of various resonant processes and developed a formulation that includes contributions from a third electron. Evidence for this two-center correlation is unclear (Schulz et al., 1989; DePaola et al., 1990). A large number of observations and several reviews of transfer excitation are now available (Tanis, 1990; Graham, 1990; Schuch et al., Mokler, 1989). One comparison between theory and experiment is shown in Fig. 38.
+
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
299
FIG. 38. Cross secton o$ for transfer excitation (TE) in S"+ + He collisions (Tanis, 1990). The dashed curve is a calculated RTE cross section multiplied by 0.85. The solid curve is a calculated NTE cross section normalized to the data.
H. PROJECTILE ELECTRONS Previously in this section we have considered projectiles that do not carry electrons into the collision (although they may carry electrons out). Now we consider the effect of projectile electrons. There are two simple limits. If a projectile electron tightly bound, then the electron simply screens the nuclear charge by one atomic unit. If the electron is loosely bound, then the projectile electron and the projectile nucleus may scatter incoherently from the target corresponding to the free collision model of Bohr in 1948 (also Dewangen and Walters, 1979; Andersen et al., 1988; Richard, 1990). In the context of the first Born approximation the projectile may be considered as a point particle with an effective charge, Z,, , given from Eq. (7 1) as
ZfC = (ZP
-
o2- S,",
(138)
where F is the atomic form factor and Sin,is the incoherent scattering function defined by Eq. (72). F and Sinearise from the interaction between a projectile and a target electron, i.e., F and S,,, are mutltielectron terms. Zeffgoes to the
McGuire
300
coherent limit of full screening at small Q and goes to the free collision limit at large Q. In Fig. 4,Z,, is illustrated. Anholt (1986) has shown that Sin=may be zero (or nearly zero) at velocities below the threshold for ionization by free electrons so that there may be a threshold for the effects of the projectile electron-target electron interaction carried by Sin,. The full screening limit is well established experimentally. For example, 07+ and N7+are known to yield the same scattering cross sections in collision with systems of loosely bound electrons. A clear example of the limit of incoherent scattering by the electrons and the nucleus is shown by the data of Wang et al. (1986) in Fig. 39. Here the cross section for collisions of a high Rydberg hydrogen atom with argon is equal to the sum of cross sections by impact with a free proton and by a free electron. The threshold for multielectron effects has been observed by Zouros et al. (1989) and Huelskoetter el al. (1989) shown in Figs. 40 and 41. Zouros et al. (1990) analyzed their results using an impulse approximation similar to that introduced by Brandt ( 1983) for resonant transfer and excitation. (Cf.Section 1II.G.) Liu and Starace (1990) have used the first Born approximation described pre-
--
1
2
I
I
5 I
I
1
20
10 I
I
I l l 1
-
50 -
-
:&++$A aC."".
0
E
20
's
-.-.-.-
I
*.a-
-.*.
-.I.
*'., A.-.-. .
c
P
10-
z : u l ul
e
u
'a.
I'
..J'
..*.#
-
.8'*
-
'. .'
...'
5-
-
'. '.
f
'.\
*.*'
-
E .'
..a
,8*.
21
2
I
I
I
5
I l l l l
70
1
20
I
1
!
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
1 22r-l , ,
PROJECTILE ENERGY (MeV)
II
tI
, , , , , , , , , , , , , , ,,
,
,
,
30 1
~
o F6*+He F6**t-$
8-
J
1
0
' 0
4
8
12 16 20 24 28
32 36 40
PROJECTILE ENERGY (MeV) FIG. 40. Cross sections for the production of ls2s2p4P states by Is + 2p projectile excitation in collisions of (a) (Y and (b) of F6+ (1~2s)projectiles with He and H2 targets versus projectile energy (Zouros er al., 1989). Calculation: electron-electron excitation (eeE) cross sections using ls2s2p4P theoretical electron impact excitation cross sections folded by the Compton profile of the target. Dashed lines, calculated eeE for H1 targets; dash-dotted lines, calculated for eeE for He targets. Arrows (at 16.3 and 25.0 MeV) indicate the projectile energies corresponding to the threshold for 1s + 2p electron-impact excitation in 05+and F 6 + ,respectively.
viously to analyze data of Duncan and Menendez (1977) for doubly differential detachment cross sections for collisions of fast H - with argon. A distinct peak for excitation of the n = 2 level of H is apparent in Fig. 42. The large peak at small angles is almost entirely due to excitation of H (n = 2) in the Sincterm of Q. (138). The remaining discrepancy between theory and experiment may be due to large experimental angular resolution. Wang et al. (1990) have recently used this first Born approximation to predict the ordering of ratios of single to double ionization of helium by projectiles with and without electrons. The analysis of Sec. III.B.2.c was extended to include
McGuire
302 2..oe+s
(a)
I
B Rojecrile a ~ g WeVN y Fic. 41. Electron-loss cross section for Ot7 + HI(Huelskoetter efa / ., 1989). Effects of electronelectron interactions are included in the calculations with the solid line (C'. Eq. 71) but not included with the dashed line where Zen = Zp.
projectile electrons by replacing 2; in Eq. (13 I) by 2iff.Using R to denote the ratio of double to single total-ionization cross sections Wang et al. (1990) predict RH+ = RHe++< RHe+< RHeo< Rphoton and R H + < R,Q < R H e o .
(139)
Partial confirmation of this prediction is given by data of DuBois. (CJ Wang et al., 1990; DuBois and Kover, 1989; and Table 11.)
I. MOLECULES Since molecules are groups of atoms, some molecular properties are similar to atomic properties, but other properties are not describable in atomic terms. For example, since H, has an internuclear separation of 1.4 Bohr radii, one might expect the cross section for H2 to differ from the cross section for two atoms of hydrogen.
I . Application of Atomic Mechanisms Understanding of atomic collisions may to some extent be applied to molecular collisions. The fragmentation of a molecule, for example, is conceptually somewhat similar to ionization of an atom. Data for fragmentation of methane, CH4, into various fragments, CH,' ,
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
?SO
200
240
290
320
360
303
A00
DETACHES Z L C R C N 3 E X Y (4)
160
200
2A0
290
320
360
1
10
60 45
30 15
0 1
DETACHED E:CT;(ON ENmGY (sV)
FIG. 42. Differential cross section for 0.5-MeVH - + Ar 4 H+ + e + Ar* (Liu and Starace, 1990). The contribution from n > 1 excited states comes from electron-electron ineractions.
CH: , CH: , CH+, and C +, was collected and analyzed by Malhi et al. (1987). The ratios of cross sections for CH:/CH; in collisions with electrons and protons are shown in Fig. 43 as a function of the collision velocity, v. Except for the lowest value(s) of n the ratios of CH,+/CH,+ are independent of both v and the sign of the projectile charge. This is consistent with the shake-type mechanism we used previously. (C’. Section III.B.2.) An alternative picture that may be applicable is a statistical energy distribution model (cf Section 11.1.2) in
304 100.0
-
t*
I 90.0
0 \
tm
-
16.0 14.0 12.0 10.0
-
I
c
0 .c d
2.5
0
-
-
-
*
-- +
-
+ R3
+
-+
80.0
0 70.0
-
- -
+
-
-*
*
*
-RP\ +
-
+
+
* +
+
+++
1 .
\
+
0
1.0 0.5
+ +
+
+
+
+
I
I
2
3
+
*+++*
+
I
4
5
v (10' c m / w c ) FIG. 43. Ratios of fragmentation cross sections for methane, CH,, collision versus velocity (Malhi et al.. 1987). Note similarity to fragmentation of atoms (i.e., ionization) shown in Fig. 19. which energy deposited by the projectile is statistically distributed after the collision producing various degrees of fragmentation. In chemistry this statistical model is referred to as quasi-equilibrium theory (Rosenstock et al., 1952). The shake and the statistical mechanisms are not necessarily different. These data are consistent with a combination of shake and independent-particle mechanisms. This picture is reinforced by observations of the Z, dependence of the fragmentation ratios shown in Fig. 44. Here the ratios for lower-order fragmentation are independent of Z, while again the C+/CH,+data are different. The CH+ data could be intermediary. Another example of a similarity of collision mechanisms is found in the case of particle transfer. In Section 1I.F mechanisms of transfer via two binary collisions were described. Confirmation of one Thomas mechanism in electron transfer was first obtained in 1983 by Horsdal-Pedersen et al. This Thomas mechanism was first observed, however, in capture of atoms from molecules by Cook
MULTIPLE-ELECTRONEXCITATION, IONIZATION, AND TRANSFER
-+ u ; +* I
0
0
90.0
0
85.0 80.0. 75.0 70.0.-
+* I
20.0
-
15.0
-
0
305
P
1
~
n
0
O
D
A A
-lR2 A
a
A
A
10.0
-
0
.c!f? +'
a,'
16.0 12.0 8.0 4.0 - 0 0.0
--
o o o o
o o o o o 0
16.0 12.0 0.0 4.0 -
P
I
20.0
0
O.O..?
# 0 0
. 0 0 0
' 2
1
1
4
1
'
I
6
Charge
I
I
8
'
1
I
1012
Zp
FIG. 44. Ratios of fragmentation cross sections for methane versus projectile charge, Z, (Malhi et al., 1987). Note similarity to fragmentation of atoms (i.e., ionization) shown in Fig. 18.
et al. in 1975 and further observed by Breinig and Sellin (1981). Another predicted Thomas mechanism, corresponding to the ridge in Fig. 35, has been observed in electron transfer but has not been observed in atom transfer in molecular collisions.
2 . Two-Center Scattering Two-center scattering introduces at least one more particle into scattering at high velocities. In Section 1I.H scattering from a diatomic molecule was approximated as scattering from two atomic centers. Here we consider this problem with an extra center in a little more detail. The molecular T matrix was expressed in Eqs. (73 and 75) as TM -- TI + T2 = f,e-i(I-fiK)v+ t2eifiQ.P
(141)
306
McGuire
where TI and T2 have a relative phase difference of exp{iQ * p } where Q is the momentum transfer of the projectile, p is the displacement vector between the two atomic centers, and pfi is the distance of center 2 from the molecular center at mass. To illustrate application of two-center effects and specific evaluation of probability amplitudes, we evaluate 1s- 1s cross sections for electron capture from diatomic molecules at high collision velocities following Wang ed al. (1989). First, we consider homonuclear molecules, in particular H, . Then we consider heteronuclear molecules. In this section we shall apply the simple BrinkmanKramers approximation, i.e.,rj = V,. a. Homonuclear Molecules. For homonuclear molecules, since by symmetry, ti = t, = t for gerade state, rl = - t, = t for ungerade state, and EL. = 112, we have from Eq. (141)
TM = t(e-iQ.(p/2)
+
e-iQ.(p/Z) )3
(142)
where ( - ) stands for gerade (ungerade). The wave function of a diatomic molecule can generally be obtained by linear Sections I1I.B-C.) combination of atomic orbitals (LCAO) or other methods. (Cf. However, the complete LCAO wave function may not be necessary when computations involve high-velocity collisions in which only inner-shell orbitals are important. Therefore, we begin by representing the molecular wave function by a simple product of 1s hydrogenic wave functions centered at each scatterer. The spherical symmetry of 1s wave functions on each center allows the simplest algebra without losing the basic physics of the process. Similarly, the ionic wave function can be expressed as a sum of 1s hydrogen wave functions centered on each scatterer. Thus we have for the simplest symmetrized molecular wave funcand ionic wave function ai tion aM
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
307
where NM and N,' are normalization constants for the molecular and ionic wave functions; 4, 4, u , and u are hydrogenic wave functions. Since the two atoms are identical, we have 3, and 4 ( u and u ) with the same form of function. The effective charge carried by molecular 3, and Q, (ionic u and LJ) has been written as ZM(Z,)in Eq. (145). We point out that AM(A,) is an overlap integral of the molecular (or ionic) wave function. In the preceding development we have introduced two T matrices: the molecular matrix TM and the effective atomic matrix t. A further relationship holds for t and its atomic limit tA which describes scattering from a free atom. Following a similar procedure to Tuan and Gerjuoy (1960), it is straightforward to show, ignoring an exchange term (not justified for excitation and ionization in H2),
W, ITMI*J
= N&"(A,M X
*
rA(elQ IP)~)
* elQ XIM)
( 146)
IP'~)),
with the initial wave function 'Po = erKxR'PM and the final wave function qb= elK,R'@Iui,in which QM and Ql are defined in Eq. (143) and (144), ul, is the 1s bound state hydrogenic electronic wave function on the projectile. Here A I M and xIM are overlap integrals between ionic and molecular wave functions given (Wang et al., 1989) by AIM
=
uTs(r)3,dr) dr
= 8 ( z , z ~ ) ~ / ~f / (ZzM ~)~,
(147)
+ Z,(PV + 4Z,)e-Z+'] where v = Z,Z - Z& . In Eq. (146) we have replaced the BK amplitude I,, of Tuan and Gerjuoy (1960) by the more general matrix t A . Comparing Eqs. (146) and (141) we have t = NMN,?(A,M
* X,M)
NMN,N,'ltA E NjtA.
(149)
The effective atomic matrix t differs from its atomic limit tA by an overall conand t = N j f A , it is primarily N: that stant N: . Since T M = t ( e - ' Q p ' 2 ? erQ@z) determines the difference between the molecular and twice the atomic total cross sections. By the preceding definition, N j is a constant combining the molecular wave function (NM), the ionic wave function (N,),and their overlap (N,'l), which depends on the form of the wave functions. The preceding expression for the overlap factor N&, derived for wave functions given by Eqs. (143) and (144), may be generalized.
308
McGuire
The reflection symmetry provides properties of a homonuclear diatomic molecule may be used to determine N: under two interesting limits, i.e., the separated atom limit ( p + m) and the united atom limit ( p + 0), without specific m, the total normalization constant for gerade calculations. For the limit as p N: and ungerade N; are the same. Namely, when p + 03, we have NM = N: = N; = (1/2)’”,N& = 1, and N: = N; = 1/2. The limit as p +S 0 is considered following Eq. (162). As an example, we now use the simple BK approximation amplitude tlK
-
in Qs. (146) and (142), where 2, is the charge of the projectile and ZT is the effective charge of the nuclei in the molecule, also denoted as Z, , to obtain T&K
= NZI t BAK ( e - i Q ( p / Z )
(151)
eiQW2)).
By Q. (77), the corresponding probability amplitude afCIKis found as
B,
=
B
2
p1/2.
These integrals can be done analytically, namely,
where Kz is the Bessel function of the third kind and second order and
x+
=
(B,I(G +
Qt)’/* -+ B p , COS(&-
X
(Z+
+ QT)’/2,
1
4p)
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
309
where +B and +p are the azimuthal directions of B and p . Thus,
X
[x2_K2(x-) e - iQlp:/2
?
x: K , (x
+
)eiQ+:i2].
The probability as a function of impact parameter is given by
P$K(B) =
The total single-capture cross section ukKis
crhK = 2
PLK(B)dZB.
(159)
The factor of 2 comes from the indistinguishability of the two electrons. When substituting Eq. (159) into (158), uIKcan be written as
where u1and u2are due to individual centers and u12 is due to the interference. The integrals for ul,2in Eq. (159) can be computed analytically to give the BK cross section, i.e.,
while vI2may be computed numerically. For the limit as p -+ m, the interference term uI2approaches zero for both gerade and ungerade. Meanwhile, in Eq. (160), the individual terms u1and vz summed over gerade and ungerade give (TI
=
+.-
+,-
=
gz =
2(4)*gA
=
(N:)CTA
+ (N;)'~A
(162)
4CA.
Thus, the total cross section uM in Eq. (164) yields 2UA9 twice the atomic cross section. For two electrons in the limit as p + 0, again N,? + 1/2, but cos(a,p,) -+ 1 and the ungerade contribution vanishes, i.e., N ; + 0, ZM+ 22, so that uiKp=o = 2uuA.Here uAdenotes the cross sections for an isolated atom,
McGuire
310
and vUAis the 1s-1s cross section per electron for the atom with the united nuclear charge. If there are four electrons, two go to the 1s level and two to the 2p level in the united atom limit. b. Application to H, . The preceding development provides a method to calculate capture from homonuclear diatomic molecules using a simple, but generalizable, molecular wave function. For H, a somewhat better wave function is the Weinbaum wave function, also used by Tuan and Gerjuoy (1960) and other authors. Specifically, we choose @M
=
+ $Is(~)~Is(~) + C[$ls(1)$lr(2) + 4lS(l)+lS(2)1}
NM{$ls(1)41s(2)
(163)
and ionic wave function = Ni’[uls(l)
@i
+-
~1s(1)1,
( 164)
where the normalization constant N,‘ is given by Eq. (145) and NM is given by
+
+
CAM]}"'. NM = 1/{2[(1 C2)(1 + A&) (165) The preceding Weinbaum wave function differs from the previous sample LCAO wave function in Eq. (149) by a bonding term characterized by c # 0. The overlap constant Ni&in this case is found to be slightly different from that in Eq. (165), namely, N& =
(AiM
5
XiM)(l
? C).
(166)
The form of Eqs. (142) and (149) is not changed. If we take the bonding term c + 0, then N& reduces to the simpler form in Eq. (149). For H, at equilibrium, weusep = 1.4, 2, = 1.193, c = 0.256, andZ, = 1.4.
c. Heteronuclear Diatomic Molecules. For heteronuclear diatomic molecules we again choose a simple LCAO wave function with only s wave contributions. Now we use I and 2 to denote the electrons, and A and B to denote the two different atomic nuclei. The molecular wave function is given by NM[auMA(l) + buME(1)l[auMA(2) + buMM(2)1, where NM = [(a4 + b4) + 2a2b2(1 + 2AL) 4ab(a2 + b2)AM]-’”. (167)
4M
Here A,
=
=
+
J ~ & A ( l ) ~ w ( 1 ) and, d r l for example,
The ionic wave function is given by
4i
+
=
+
Ni[a’~,(2) + b’ui,(2)1,
( 170)
with N, = [(a’)2 (b’)2 2a’b‘Ai]1’2and A, = J u ~ A u , B dr. Using these wave
3 11
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
functions, it is straightforward to show that to an overall phase
TM
TA
+
TBeiQ'p,
(171)
-
where it is convenient to choose the origin at the center of mass of the molecule. The mathematical evaluation of probability amplitudes is now similar to the b and a' + b', corresponding to homonuclear case, which is recovered as a c = 1 in Eq. (163). Although the algebra is quite similar for the homonuclear and heteronuclear cases, the identification of final states and counting of electrons is a little different. If A # B in molecule AB, then A + B is a different final state than AB+.Thus TAdiffers from TB. At the same time gerade and ungerade labels no longer apply since there is reflection symmetry. We also note that for most molecules, there are four electrons in the ground state, twice as many as in H, . Hence, except for where c A , B is a 1s atomic two-electron systems such as HZ,c&zm2(cA cB) cross section per electron for center A and B .
+
d . Other Examples. This two-center analysis may be applied, if an exchange is included, to ionization and excitation as well as to the case of electron transfer considered here. In each of these cases this theory predicts that the cross section depends on the orientation of the molecule as illustrated in Fig. 45. This depenIO4
H++H,
3 0
3
10'
-
I
5 MeV
.-
0
40
80
120
160
ORIENTATION ANGLE, 8, (degree) FIG. 45. Capture cross section from H2as a function of the orientation angle O,, of the H2 molecular axis relative to the beam direction (Wang e t a / . . 1989).
McGuire
312
dence on molecular orientation may be observed in Coulomb recoil distributions from collisions of charged particles with molecules. Some data were taken in 1963 by Dunn and Kieffer corresponding to single ionization of H, in collisions with electrons and were analyzed by Zare (1967). Coupled channel calculations for electron impact on H, and other molecules are common (Collins and Schneider, 1988; McConkey et al., 1988; Shimamura and Takayanaki, 1984), but few calculations for impact by heavy projectiles have been done. A few observations of Coulomb explosions in collisions with charged heavy particles are in progress as of early 1991. Data dealing effects of multielectron transitions in Coulomb explosions may be available soon. For multiple-electron transitions one may apply the independent distinguishable electron approximation. One must ensure, however, that the two amplitudes p/2 and B - p/2 are properly separated (especially at large p ) when for B applied to Eq. (29). This can be done by distinguishing which center has a vacancy in the final state. For symmetric molecules such as H,, proper separation may result from proper application of Pauli exclusion. For ionization and excitation of molecules by projectiles carrying electrons the scattering amplitude may be evaluated in the first Born approximation by simply replacing 2, by Z,, as discussed in Section 1I.G. I . The resulting T matrix (or scattering amplitude) is then given by
+
T = 4r/Q2Zf'(0) cos(Q
*
P)fatornie(p)
(172)
corresponding to a product of internuclear scattering times the effective projectile charge times the two-center interference times an atomic scattering amplitude. ( C ' discussion at the end of Section 1I.G. 1.) Data testing the validity of this simple form is not available as of 199 1.
IV. Conclusion Multielectron transitions are many-body phenomena. When more than one electron undergoes a transition in an atomic or a molecular collision, the event contains information about the dynamics of many-body systems. The electrons in these dynamic events may act independently or they may influence one another. In this review both independent and correlated systems of electrons in highvelocity collisions have been considered, leading to a few general observations as follows. Our understanding of many-electron effects is incomplete, even in fast collisions involving only a few electrons, i.e., relatively easy examples of manyelectron phenomena. Some understanding has been achieved in systems in which electron correlation is weak so that an analysis using the independent-electron
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
3 13
approximation is sensible. However, incorporating exclusion effects of identical electrons in the independent-electronapproximation is incomplete. Correlation arising from the Coulomb interaction between electrons is often complicated and hard to understand. Some analysis of the role and nature of correlation in two-electron transitions has been possible. Both perturbative methods and coupled channel methods have been useful in the analysis of observed data. However, the nature of specific physical mechanisms responsible for dynamic correlations remains mostly unclear, although experiments in the last several years using antiprotons and positrons have led to some clarification of our understanding. Nevertheless, even the simplest cases-for example, ratios of double- to single-electron transitions in helium in the high energy limit (summarized in Table II)-remain only partially understood at present. Various approaches have been used to help understand multielectron effects in atomic and molecular collisions. Kinematic restrictions imposed by conservation of energy and momentum have been helpful in interpreting some many-body effects in terms of combinations of two-body interactions. Effects of mutual interaction of electrons on two centers in relative motion have been observed and analyzed with some success within the confines of first-order perturbation theory. Two-center effects in the analysis of molecular dynamics have been predicted and may be soon tested by experiment. But both theory and experiment becomes more difficult as the number of centers, the number of electrons, and the number of states involved increase. Eventually as the complexity of the systems grows, our understanding may be limited to statistical descriptions, an intellectual position that is common with our understanding both in basic quantum mechanics and in classical chaos. In summary, understanding the transition from how two bodies interact to how many bodies interact provides a conceptual basis for understanding various aspects of our environment. The human environment is sufficiently rich in both complexity and possibility to challenge the human mind as well as the human spirit. The truths that emerge from this challenge can provide intellectual satisfaction and at the same time provide meaning and alternative in our lives.
Acknowledgments I gratefully acknowledge discussion with 0. L. Weaver, J. Straton, J. P. Giese, C. D. Lin, C. L. Cocke, I. Ben-Itzhak, P. Richard, R. Shingal, T. Ishihara, N. Stolterfoht, W. Meyerhof, T. Aberg and Y.D. Wang as well as some guidance over many years from D. R. Bates. This work has been partially supported by the Division of Chemical Sciences, Office of Basic Research, U. S. Dept. of Energy.
McGuire
314
Appendix: Correlation In a dictionary correlated means interdependent; that is, correlated particles influence one another. Electron-electron correlation occurs because electrons interact with each other. In a technical sense there are different ways to specify this kind of correlation. It is possible to specify that electron correlation is any effect arising from the Coulomb interaction between electrons. It is also possible to define correlation as a deviation from a product of electron probability densities. This latter definition of correlation is sensible because the probability for any reaction may be written as a product of individual single-electron probabilities if the electron-electron interaction is ignored. This definition is also sensible since it corresponds to the definition of correlation used in stochastics and statistical mechanics. To the extent that understanding macroscopic properties in terms of microscopic atomic properties is interesting, it is sensible to choose the broader stochastic definition of correlation (Balescu, 1975; Huang, 1987) since statistics are ultimately useful in dealing with large numbers of atoms. Correlation is often introduced in stochastics and statistical mechanics (Balescu, 1975) as a generalization of the notion of statistical standard deviation of a distribution,
&=
((x -
(173)
(x))2).
If c2is not zero, then the distribution is not localized at a single value ( x ) , but rather the distribution is spread out. This leads to the idea that correlated functions are also spread out, i.e., not confined to a single x, (denoting a single particle), but connect x,and x,. Consider the one-particle distribution function given by f i ( x , ) = p(x,)
=
4*4.
(174)
The two-particle distribution function is uncorrelated iff2(xl,x2) = f l ( x , )fl (xz). In general, however, f2 is correlated and may be written as f2(XIr
x2)
= fl(xl)fl(x2)
+
gz(x1,
(175)
x2).
g, is called the two-particle correlation function. For three particles one may generally write h(XI
9
x29
x3)
=
fl(xl)fl(x2)fl(x~)
+ fl(xl)g2(x29
+ fi(X3)g2(XIr
x3)
+ fl(X2)g2(x1,
x2)
+ g3b-1,x 2 r
x3)
(176)
x3)
and so forth for f4, fs, . . . fN, . . . . This is called a cluster expansion of the distribution function fN in terms of the jth-order correlation functions g, . Correlation has been defined and studied in both space and time. Here we
MULTIPLE-ELECTRONEXCITATION, IONIZATION, AND TRANSFER
3 15
shall consider spatial correlation. Examples of time correlation include timeordering effects and memory. Most studies have been confine to two-particle or pair correlation since experimental studies of higher-order correlation are often difficult. We do note, however, that the N-particle distribution function can be 1 particle distribution function via the two-particle correlation related to the N potential u(xi - x j ) . This is referred to as BBGKY hierachy. The generalized correlation coefficient T is defined as follows. If P is a oneparticle operator, then for a two-electron system
+
75 =
1I (1fi
&(XI,
x * ) p ( ~ I ) p ( rdrl z ) dr*
( r V W dr) -
(1f,
(177)
(r)P(r)dr)*
such that (i) T = 0 if the electrons are uncorrelated and (ii) [TI < 1. Most importantly, T is independent of choice of basis functions. This generalized correlation coefficient has been applied by Christensen-Dalsgaard ( 1988) to atomic structure, but as of early 1991 application has not yet been made to dynamic correlation in atoms and molecules. Finally, let us note that detailed effects of correlation tend to disappear at both large and small space (and time, etc.) scales. On the large scale, correlation effects may average out. For example, fluctuating phase terms may average to a relatively small value on a sufficiently large scale. On a small scale the independent particle limit may dominate if the separation between the particles is large compared to the size of the region of interaction. For example, if the separation between electrons is much larger than the typical interaction distance of a probe (e.g., projectile), correlation effects may be small for relatively well separated electrons. REFERENCES Aberg, T.(1967). Phys Rev. 156, 35. Aberg, T. (1976). In “Photoionizationand Other Probes of Many Electron Interactions” (F. Wuillemier, ed.), p. 49. Plenum, New York. Aberg, T. (1987). Nucl. Instr. Merh. A262, 1 . Aberg, T. (1990). Private communication and Atlanta Proceedings (1973) (R. W. Fink er nl., eds.), US At Eng. Corn. No. Conf. 720404, p. 1509. Aberg, T., Blomberg, A . , Tulkki, J . , and Goszinski 0. (1983). Phvs. Rev. Left. 52, 1207. Aberg, T., Blomberg, A , , and MacAdam, K. B. (1987). J . Phvs. B20, 4795. Alston, S. (1989). Phys. Rev. A. 40,4907. Alston, S . , and Macek, J. (1982). Phys. Rev. (SPB). Amusia, M. Y . , Drukarev, E. G., Gorshkov, V. G., and Kazachkov, M. D. (1975). J . Phys. BS, 1248.
316
McGuire
Andersen, L. H., Hvelplund, P. Knudsen, H., Moller, H. P., Elsner, K., Rensfelt, K. G.,and Uggerhoj, E. (1986). Phys. Rev. Lett. 57,2147. Andersen, L. H., Hvelplund, P. Knudsen, H., Moller, S. P., Sorensen, A. H., Elsner, K., Rensfeld, K. G.,and Uggerhoj, E. (1987). Phys. Rev. 36, 3612. Andersen, L. H., Nielsen, L. B., and Sorensen, J. (1988). J . Phys. BZ1, 1587. Andersen, L. H., Hvelplund, P., Knudsen, H., Moeller, S. P., Pedersen, J. 0. P., Uggerhoj, E., Elsner, M., and Morenzoni, E. (1989). Phys. Rev. Lett. 62, 1731. Andersen, L. H., Hvelplund, P., Knudsen, H., Moller, S. P., Pedersen, J. 0. P., Sorensen, A. H., and Uggerhoj, E. (1990). Phys. Rev. Letters 65, 1687. Anholt, R. (1986). Phys. Lett. A114, 126. Anholt, R., Meyerhof, W. E., Xu, X. Y., Gould, H., Feinberg, B., McDonald, R. J., Wegner, H. E., and Thieberger, P. (1987). Phys. Rev. A36, 1586. Arcuni, P. W., and Schneider, D. (1987). Phys. Rev. A36, 3059. Auger, P. (1926). Ann. Phys. 6, 183. Balescu, R. (1975). In “Equilibrium and Non-Equilibrium Statistical Mechanics.” Wiley, New York. Barany, A. (1990). In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P. M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.), p. 246. American Institute of Physics, New York. Barat, M. (1988). In “Comment Atom and Molec.” Phys. 21,307. Barkas, W. H.,Birnbaum, W., and Smith, F. H. (1956). Phys. Rev. 10,778. Barnett, C. F. (1990). “Atomic Data for Fusion.” Controlled Fusion Atomic Data Center, Oak Ridge, Tennessee. Basbas, G.(1982). Nucl. Instr. Meth. B4, 227. Bassell, R. H., and Gerjuoy, E. (1960). Phys. Rev. 117,749. Bates, D. R. (1958). Proc. Royal SOC.A247,294. Bates, D. R. and Griffing, G.(1955). Proc. Phys. Soc. (London) A66, 961. Becker, R. L. (1985). Private communication. Becker, R. L. (1988). In “High Energy Atomic Collisions” (D. Berenyi and G.Hock, eds.), p. 447. Springer Verlag, Berlin. Becker, R. L., Ford, A. L., and Reading, J. F. (1980). J. Phys. B13, 4059. Belkic, Dz., Gayet, R., and Salin, A. (1979). Phys. Rep. 56, 279. Ben-Itzhak, I., and McGuire, J. H. (1988). Phys. Rev. A38,6422. Ben-Itzhak, I., Gray, T. J., Legg, J. C., and McGuire, J. H. (1988). Phys. Rev. A37, 3685. Berg, H., Doerner, R., Kelbch. C., Kelbch, S., Ullrich, J., Hagmann, S., Richard, P., SchmidtBoecking, H., Schlochter, A. S., Prior, M., Crawford, H. J., Engelange, J. M.,Flores, J., Loyd, D. H., Pedersen, J., and Olson, R. E. (1988). 1.Phys. B21,3929. Bhalla, C. P. (1990). Phys. Rev. Lett. 64, 1103. Bhalla, C. P., and Karim, K. R. (1989). Phys. Rev. A39,6060. Blomberg, A., Aberg, T., and Goscinski, 0. (1986). J. Phys. B19, 1063. B o b , N. (1913). PhilMag. 26, 857. Bohr, N. (1948). Danske. K . Vidensk. Selsk. Mat. Fys. Meddr. 18, No. 8 Bordenave-Montesquieu, A., and Beniot-Cattin, P. (1971). Phys. Lett. 37A, 243. Bordenave-Montesquieu, A., Gliezes, A., and Benoit-Cattin, P. (1982). Phys. Rev. A25, 245. Bottcher, C. (1982). Phys. Rev. Lett. 48,85. Bottcher, C., Bottrell, G.J., and Strayer, M. (1990). Reprint. Brandt, W. (1983). Phys. Rev. A27, 1314. Brauner, M., Briggs, J. S., and Klar, H. (1989). J. Phys. B22,2265. Breinig, M. and Sellin, (1981). Bull. Am. Phys. Soc. 76, 601. Briggs, J. S . , and Macek, J. (1991). In “Advances in Atomic and Molecular Physics” (D. R. Bates and B. Bedersen, eds). In press.
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
3 17
Briggs, J. S. (1988). In “Electronic and Atomic Collisions” (H. B. Gilbody, W. R. Newell, F. H. Read, and A. C. H. Smith, eds.), p. 13. Elsevier Science, New York. Briggs, J. S.. and Roberts, M. J. (1974). J . Phys. B7, 1370. Briggs, J. S . , and Taullbjerg, K. (1978). In “Topics in Current Physics” (I. Sellin, ed.), Vol. 5, p. 105. Springer Verlag, New York. Briggs, J. S., and Taulbjerg, K. (1979). J . Phys. B12,2565. Bruch, R., Schneider, D., Schwartz, W. H. E., Meinhart, M., Johnson, B. M., and Taulbjerg, K. (1979). Phys. Rev. A19,587. Bruch, R., Fuelling, S. et al. (1990). Private communication. Bruch, R., Alteck, P. L., Trabert, E., and Heckmann, P. H. (1984). J. Phys. B17, L655. Bruch, R., Kocbach, L., Trabert, E., Heckmann, P. H., Raith, B., and Will, U. (1985). Nucl. Insirum. Meih. B9, 438. Byron, F. W., and Joachain, C. J. (1967). Phys. Rev. 164, 1 . Carlson, T. A. (167). Phys. Rev. 156, 142. Carlson, T. A., Nestor, C. W., Tucker, T. C., and Malik, F. B. (1968). Phys. Rev. 169, 127. Carter, S. L., and Kelly, H. P. (1981). Phys. Rev. 24, 170. Charlton, M., Andersen, L. H., Brun-Nielsen, L., Deutch, B. I., Hvelplund, P., Jacobsen. F. M., Knudsen, H.. Lariccha, G., Poulsen, M. R., and Pedersen, J. 0. (1988). J . Phys. B21, L545. Chen, J. C. Y., and Chen, A. C. (1972). I n “Advances in Atomic and Molecular Physics” (D. R. Bates and I. Estermann, eds.) Vol. 8, p. 71. Academic Press, New York. Chetioui, A. (1990). Private communication. Christensen-Dalsgaard, B. (1988). J . Phys. B21, 2539. Cocke, C. L. (1979). Phys. Rev. A20, 749. Cocke, L. L., and Olson, R. E. In press. Collins, L. A., and Schneider, B. 1. (1988). In “Electronic and Atomic Collisions” (H. B. Gilbody, W. R. Newell, F. H. Read, and A. C. H. Smith, eds.) Elsevier Science, 57. Cook, C. J., Smith, N. R. A., and Heinz, 0. (1975). J . Chem. Phys. 63, 1218. Crasemann, B. (1975). “Atomic Inner-Shell Processes.” Academic Press, New York. Crothers, D. S. F., and Dunseath, K. M. (1990). Preprint. Datz. S., Hippler, R., Anderson, L. H., Dittner, P. F., Knudsen, H., Krause, H. F., Miller, P. D., Pepmiller, P. L., Rassel, T., Stolterfoht, N., Yamazaki, Y., and Vane, C. R. (1987). Nucl Insir. Meih. AM2, 62. Datz, S. Hippler, R., Anderson, L. H., Dittner, P. F., Knudsen, H., Krause, H. F., Miller, P. D., Pepmiller, P. L., Rassel, T., Shuch, R., Stolterfoht, N., Yamazaki, Y., and Vane, C. R. (1990). Phys. Rev. A41, 3559. De Paola, B. D., Parameswaran, R. and Axmann, W. S. (1990). Phys. Rev. A41,6533. Devi, K. R., and Garcia, J. D. (1984). Phys. B18,4589. Devi, K. R. S., and Garcia, R. D. (1990). Preprint. Dewangen, D. P., and Walters, H. R. J. (1979). J . Phys. B11,3893. Dreizler, R. M., Ast, H. J., Henne, A., Luedde, H. J., and Starg, C. (1990). In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P. M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.) p. 258. American Institute of Physics. New York. DuBois, R. D., and Kover, A. (1989). Phys. Rev. AM, 3605. DuBois, R. D., and Manson, S. T. (1987). Phys. Rev. A35, 2007. Duncan, M. M., and Menendez, M. G. (1977). Pys. Rev. A16, 1799. Dunn, G. H., and Kieffer, L. J. (1963). Phys., Rev. 132, 2109. Eadie, W. T., Drijard, D., James, F. E., Roos, M., and Soudoulet, B. (1971). “Statistical Methods in Experimental Physics.” North Holland, Amsterdam. Eckhardt, M., and Schartner, K. H. (1983). Z. Physik. A312, 321. Edwards, A. K. (1990). Private communication.
318
McGuire
Edwards, A. K., Wood, R. M., and Ezell, R. L. (1986). Phys. Rev. AM,441 I . Edwards, A. K., Woods, R. M., Beard, A. S., and Ezell, R. L. (1988). Phys. Rev. A37, 3697. Eichler, J. (1977). Phys. Rev. A15, 1856. Fano, U. (1961). Phys. Rev. A124, 1866. Feagin, J. M., Briggs, J., and Reeves, T. M. (1984). J. Phys. B17, 1057. Fetter, A. L., and Walecka, J. D. (1971). “Quantum Theory of Many-Particles Systems.’’ McGrawHill, San Francisco. Fink, R. W., Manson, S. T., Palms, J. M., and Rao, P. V. (1973). “Roc. Intl. Conf. Inner Shell Ionz. Phenomena and Future Appl.” Atlanta, US Atomic Energy Comm. Rep. No. CONF720404. Oak Ridge, Tennessee. Ford, A. L. (1981). Private communication. Ford, A. L., and Reading J. F. (1989). Nucl. Instr. Merh. B40/41,362. Ford, A. L., Reading, J. F., and Becker, R . L. (1979). J. Phys. B12,2905. Fritsch, W., and Lin, C. D. (1983). Phys. Rev. A27,3361. Fritsch, W., and Lin, C. D. (1986). J. Phys. B19, 2683. Fritsch, W., and Lin, C. D. (1988). Phys. Rev. Lett. 61, 690. Fritsch, W., and Lin, C. D. (1990). Phys. Rev. 41,4776. Fritsch, W., and Lin, C. D. (1991). PhysicsReports. In press. Fuelling, S. and Bruch, R. (1990). Private communication. Gardiner, C. W. (1985). “Handbook of Stochastic Methods.” Springer Verlag, Berlin. Gau, J. N., and Macek, J. (1975). Phys. Rev. A12, 1760. Gayet, R . , and Salin, A. (1990). J . Phys. B. In press. Giese, J. P., and Horsdal, E. (1988). Phys. Rev. Lett. 57, 1414. Giese, J. P., Schulz, M., Swensen, J. K., Schoene, H., Benhennu, M., Varghese, S. L., Vane, C. R., Dittner, P. F., Shafroth, S. M.,and Datz, S. (1990). PhysRev A42, 1231. Goldberger, M. L., and Watson, K. M. (1964). “CollisionTheory,” Chap. I t . Wiley, New York. Graham, W. G. (1990) In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P. M. Koch, M. S. Lubell. and T.B. Lucatorto, eds.), p. 544. American Institute of Physics, New York. Gramlich, K., Grun, N., and Scheid, W. (1986). J. Phys. B19, 1457. Gravielle, M. S . , and Miraglia, J. E. (1990). Reprint. Gryzinski, M. (1985). Phys. Rev. 138, A349. Gudnov, A. L., Kunikeev, Sh.D., Novikov, N. V., and Senashenko, V. S. (1989). In “Abstracts of Contributed Papers, 16th International Conference on the Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, M. S. Lubell, and T. B. Lucatorto, eds.), p. 502, American Institute of Physics, New York. Hahn, Y. (1990). In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P. M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.) p. 550. American Institute of Physics, New York. Hall, J., Richard, P., Gray, T. J., Newcomb, J., Pepmiller, B., Lin, Itoh, A., Zouros, T. J., Schneider, D., Stettner, U., Zeitz, W., and Stolterfoht, N. (1985). J. Phys. Bl8, 4581. C. D., Jones, K. Johnson, B., and Gregory, D. (1983). Phys. Rev. A28,99. Hall, J., Richard, P., Grey, T. J., Newcomb, J., Pepmiller, P., Lin, C. D., Jones, K., Johnson, B., and Gregory, D. (1983). Phys. Rev. A28,99. Halprn, A., and Thomas, B. K. (1979). Phys. Rev. Lett. 43, 33. Hansteen, J. H., and Mosebekk, 0. P. (1972). Phys. Rev. Lett. 29, 1961. Hansteen, J. M., Johansen, A. M., and Kocbach, L. (1975). At. Data Nucl. Data Tables 15, 305. Haugen, H. K . , Anderson, L. P., Hvelplund, P. and Knudsen, H. (1982). Phys. Rev. A26, 1950. Hazi, A. (1981). Phys. Rev. A23, 2232. Heber, 0..Bandson, B. B., Sampoll, G., and Watson, R. L. (1990). Phys. Rev. Letters 64, 851. Hertal, C. and Salin,A. (1980). J . Phys. B13,785.
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
3 19
Hippler, R. (1988). In “Coherence in Atomic Collision Physics” (H. J. Beyer, K. Blum, and R. Hippler, eds.) 137. Plenum, New York. Holland, D. M. P.. Codling, K., West, J. B., and Marr, G. V. (1979). J. Phvs. B12, 2465. Horbatsch, M. (1986). J. Phys. B19, L193. Horbatsch, M., and Dreitzler, R. M. (1986). Z. Phvsik D2, 183. Horsdal-Pederson, E. and Larsen, L. (1979). J . Phys. B . 12,4085. Horsdal, E., Jenson, B., and Nielson, K. 0. (1986). Phys. Rev. Lett. 57, 1414. Horsdal-Pedersen, E., Cocke, C. L., and Stocki. M. (1983). Phvs. Rev. Lett. 50, 1910. Huang, K. In “Statistical Mechanics.” Wiley, New York. (1987). Hvelplund (1982). Private communication. Huelskoetter, H. P., Meyerhof, W. E., Dillard, E. D., and Guardola, N. (1989). Phys. Rev. Lett. 63, 1938. Hvelplund, P., Haugen. H. K., and Knudsen, H . (1980). Phys. Rev. A22, 1930. Inokuti, M. (1971). Rev Mod. Phys. 43, 297. Inokuti, M. (1979). Argonne National Laboratory Report no. ANL-76-88, p. 177 (unpublished). Ishihara, T., and Hino, K. (1989). Private communication. Ishihara, T., and McCuire, J. H. (1988). Phys. Rev. A38, 3310. Itoh, A., Zouros, T.J., Schneider, D., Stettner, U., Zeitz, W., and Stolterfoht, N. (1985). J . Phvs. B18,4581. Jacobs, V. L., and Burke, P. G. (1972). J. Phvs. BS, L67. Jain, A., Lin, C. D., and Fritsch, W. (1989). Phys. Rev. A39, 1741. Joachain, C. J. (1975). “Quantum Collision Theory,” Chap. 19. North Holland, Amsterdam. Kamber, E. Y. Cocke, C. L. Cheng, S . , and Varghese, S. L. (1988). Phvs. Rev Lett. 60, 2026. Kauffman, R. L., McGuire, J. H., Richard, P., and Moore, C. F. (1973). Phys. Rev. AS, 1233. Kimura, M., Sato, H., and Olson, R. E. (1985). Phvs. Rev. A28, 2085. Knudsen, H., Andersen, L. H., Hvelplund, P., Astner, G., Cederquist, H., Darnared, H., Liljeby, L., and Rensfelt, K. G. (1984). J. Phvs. B17, 3545. Kohland, H., and Dreizler, R. M. (1986). Phys. Rev. Lett. 56, 1993. Kristensen, F. G., and Horsdal, P. E. (1990). J. Phvs. B. (submitted). Kunikeev, Sh.D., and Senashenko, V. S. (1989). In “Abstracts of Contributed Papers, 16th International Conference on the Physics of Electronic and Atomic Collisons” (A. Dalgarno, R. s. Freund, M. S. Lubell, and T. B. Lucatorto, eds.), p. 503. American Institute of Physics, New York. Lindhardt, J. (1988). Private communication. Lindle, D. W., Ferrett, T. A., Becker, U., Korbin, P. H., Truesdale, C. M., Kerkhoff, H. G.,and Shirley, D. A. (1985). Phys. Rev. A31,714. Liu, C. R., and Starace, A. F. (1990). Phys. Rev. A42, 2689. tuedde, H. J., and Dreizler, R. (1985). J. Phvs. B18, 107. Macek, J. (1989). Private communication. Madison, D. H. (1990). In “The Physics of Electronic and Atomic Collisons” (A. Dalgarno. R. S. Freund, P.M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.) p. 149. American Institute of Physics, New York. Madison, D. H., and Merzbacher, E. (1975). In “Atomic Inner Shell Processes” (B. Crasemann, ed.) Magnus, W. (1954). Commun. Pure Appl. Marh. 7,649. Mahli, N. B., Ben-ltzhak, I., Gray, T. J., Legg, J. C., Needham, V., Carnes, K., and McGuire, J. H. (1987). J . Chem. Phys. 87,6502. Massey, H. S. W.,Burhop, E. H. S . , and Gilbody. H. B., (1969). In “Electronic and Ionic Impact Phenomena.” Oxford University Press, Oxford. Matsuzawa, M. (1980). J . Phvs. B13, 3201. McCarroll, R . , and Salin, A. (1969). Compt. Rend Acad. Sci. 263,329. McConkey, J. W.. Trajmar, S., and King, G. C. M. (1988). Comments At. Mol. Phys., 17.
320
McGuire
McDowell, M. R. C., and Coleman, J. P. (1970). “Introduction to the Theory of Ion-Atom Collisions.” North-Holland, New York. McGuire, J. H. (1982). Phys. Rev. Len 49, 1153. McGuire, J. H. (1984). J. Phys. B17, 1779. McGuire, J. H. (1985). J. Phys. B18, L75. McGuire, J. H. (1987). Phys. Rev. AM, 1114. McGuire, J. H. (1988). In “High Energy Ion-Atom Collisions” (D. Bereneyi and G. Hocks, eds.), p. 415. Springer Verlag, Berlin. McGuire, J. H., and Burgdorfer, J. (1987). Phys. Rev. A X , 4089. McGuire, J. H., and Heil, T. G. (1987). Nucleurlnstr. undMethods B24/25, 243. McGuire, J. H., and Macdonald, J. R. (1975). Phys. Rev. A l l , 146. McGuire, J. H., Mueller, A., and Salzborn, E. (1987). Phys. Rev. A35, 3265. McGuire, J. H., Stolterfoht, N., and Simony, P. R. (1981). Phys. Rev. A24, 97. McGuire, J. H., and Straton, J. C. (1990a). In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P. M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.) p. 280. American Institute of Physics, New York. McGuire, J. H., and Straton, J. C. (1990b). Phys. Rev. A43, 5 184. McGuire, J. H., and Weaver, 0. L. (1977). Phys. Rev. A16, 41. McGuire, J. H., and Weaver, 0. L. (1984). J . Phys. 817, L583. McGuire, I. H., and Weaver, 0. L. (1986). “Long Range Terms in Atomic Collisions.” Phys. Rev. AM, 2473. McGuire, I. H., Deb, N. C., Aktas, Y., and Sil, N. C. (1988). Phys. Rev. A38, 3333. McGuire, J. H . , Straton, J. C., Axmann, W. J., Ishihara, T., and Horsdal, E. (1989). Phys. Rev. Lett. 62, 2933. Mehlhorn, W. (1978). Summer lecture notes at University of Aarhus. Mehlhorn, W. (1985). In “Atomic Inner-Shell Processes” (B. Crasemann, ed.). Plenum, New York. Menendez, M. G., and Duncan, M. M. (1989). Phys. Rev. A39, 1534. Merkuriev, S. P. (1976). Yad. Fir. 24, 289. Merzbacher, E. (1984). In “Electronic and Atomic Collisions’’ (J. Eichler, I. V. Hertel, and N. Stolterfoht, eds.) p. 1, North Holland, Amsterdam. Messiah, A. (1965). “Quantum Mechanics,” Chap. XIX, Sec. 24. Wiley, New York. Meyerhof, M. E. (1990). Private communication. Meyerhof, W. E., and Huelskoetter, H. P. (1990). U.S. Japan Seminor, in press. Anchorage, Nucl. Inst. Meth. Miraglia, J. E., and Gravielle, M. S. (1987). In “Abstracts of the Fifteenth Intl. Conf. on Physics of Electronic and Atomic Collisions” (J. Geddes, H. B. Gilbody, A. E. Kingston, and C. J. Latimer, eds.) p. 517, North Holland, Amsterdam. Mittleman, M. H. (1966). Phys. Rev. Leu. 16, 498. Mokler, P. H., Reusch, S.,Stoehlker, T., Schuch, R., Schulz, M., Wintermeyer, G., Stachura, Z., Warczak, A., Mueller, A,, Awaya, T., and Kambara, T. (1989). In “Radiation Effects and Defects in Solids” (Gordon and Breach, N.Y.). p. 39. Mueller, A., Groh, W. and Salzborn, E. (1983). Phys. Rev. Lett 51, 107. Mueller, A., Tinschert, K., Achenbach, C., and Salzborn, E. (1985). Phys. Rev. Letrers 54,414. Mueller, A,, Schuch, B., Groh, W., and Salzborn, E. (1987) Z. Phys. D7,251. Mukoyama, T. (1986) Bull. Chem. Res. Koyoto Univ. 64, 1. Olson, R. E. (1987). Phys. Rev. A36, 1519. Olson, R. E. (1988). In “Electronic and Atomic Collisions” (H.B. Gilbody, W. R. Newell, F. H. Read, and A. C. H. Smith, eds.) p. 271. Elsevier Science, London. Olson, R. E., Ullrich, J., and Schmidt-Boecking, H. (1989a). Phys. Rev. A40,2843. Olson, R. E., Ullrich, J., and Schmidt-Boecking, H. (1989b). Phys. Rev. A39,5572. Palinkas, J., Schuch, R., Cederquist, H., and Gustafsson, 0. (1989). Phys. Rev. Lett. 22, 2464. Palinkas, J., Schuch, R., Cederquist, H., and Gustafsson, 0. (1990). Reprint.
MULTIPLE-ELECTRON EXCITATION, IONIZATION, AND TRANSFER
32 1
Peach, G., Willis, S. L., and McPowell, M. R. C. (1985). J . Phys. B18, 3921. Pedersen, J. 0.P., and Folkman, F. (1 990). J . Phys. B23,44 1 . Pedersen, J. 0. P., and Hvelplund, P. (1989). Phys. Rev. Leu. 62,2373. Pepmiller, P. L., Richard, P., Newcomb, J., Dillingham, R., Hall, J. M., Gray, T. J., and Stockli, M. (1983). IEEE Trans. Nucl. Sci. NS30, 1002. Prost, M., Morgenstern, R., Schneider, D., and Stolterfoht, N. (1977). In “Abstract of the Tenth International Conference on the Physics of Electronic and Atomic Collisions” (M. Barat and J. Reinhardt, eds.) Commissariat a1 Energie Atomique, Paris). Reading, J. F. (1973). Phys. Rev. AS, 3262. Reading, J. F., and Ford, A. L. (1980). Phys. Rev. A21, 124. Reading, J. F., and Ford, A. L. (1987a). Phys. Rev. Len. 58,543. Reading, J. F., and Ford, A. L. (1987b). J. Phys. B20,3747. Reading, J. F., Ford, A. L., Smith, J. S., and Becker, R. L. (1984). I n “Electronic and Atomic Collisions” (J. Eichler, I. V. Hertel, and N. Stolterfoht, eds.) Elsevier Science, Berlin. Reinhold, C. O., and Falcon, C. A. (1986). 1.Phys. A33, 3859. Richard, P. (1975). In “Atomic Inner-Shell Processes” (B. Craseman, ed.), p. 74. Academic Press, New York. Richard, P. (1990). In invited talks of “X-90, 15th Intl. Conf. on X-Ray and Inner-Shell Processes,” Knoxville, Tennessee, Am. Insr. Phys. Conf. Proceedings 215 (T.A. Carlson, M. 0. Krause, and S. T. Manson, eds.), p. 315. American Institute of Physics, New York. Roberts, M. J. (1985). J. Phys. B18, L707. Rocin, P., Barat, M., Gaborland, M. N., Guillemot, L., and Laurant, H., (1989). J. Phys. B22, 509. Rosenstock, H. M., Wallenstein, M. B., Wahrhaftig, A. L., and Eyring, H. (1952). Proc. Nurl. Acud. Sci. USA 38,667. Rosseland, S . (1923). J. Phys. B14, 1972. Rudd, M. E. (1956). Undergraduate thesis, unpublished. Rudd, M. E., Kim, Y.K., Madison, D. H., and Gallagher, J. W. (1985). Rev. Mod. Phys. 57,967. , Russek, A. (1963). Phys. Rev. A132, 246. Ryufuku, H., and Watanabe, T. (1979). Phys. Rev. A19, 1538. Salin, A. (1987). Phys. Rev. A36,5471. Salzborn, E. (1990). In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P. M. Koch, M. S. Lubell, and T. B., Lucatorto, eds.), p. 290, 366, American Institute of Physics, New York. Salzborn, E. and Mueller, A. (1986). Rivate communication. Sampson, J. R. (1990) Phys. Rev. Lett. 65, 2861. Schartner, K. H. (1990). I n “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P. M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.), p. 215. American Institute of Physics, New York. Schlachter, A. S., Steams, J. W., Berkner, K. H.,Bernstein, L. M., Clark, M. W., DuBois, R. D., Graham, W. G., Morgan, T.J., Mueller, D. W., Tanis, J. A., and Woodland, W. T.(1990). In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P. M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.), p. 366. American Institute of Physics, New York. Schmidt, V., Sander, and Kuntzmuller, H. (1976). Phys. A13, 1748. Schuch, R., Justiniano, E., Schulz, M., Mokler, P. M., Reusch, S., Datz, S., Dittner, P. F., Giese, J., Miller, P. D., Schoene, H., Kambara, T., Mueller, A., Stachura, Z., Vane, R., Warczak, A,, and Wintermeyer, G. (1990). In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno, R. S. Freund, P.M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.) p. 562. American Institute of Physics, New York. Schulz, M., Giese, J. P., Swenson, J. K., Datz, S., Dittner, P. F., Krause, H. F., Schoene, H., Vane, C. R., Benhenne, M., and Shafroth, S. M. (1989). Phys Rev. Lert. 62, 1738. Shah, M. B., and Gilbody, M. B. (1985). J. Phys. B18, 899.
322
McGuire
Shakeshaft, R., and Spruch, L. (1979). Rev. Mod. Phys. 51, 369. Shimamura. I., and Takayanagai, K. (1984). “Electron-Molecule Collisions.’’ Plenum, New York. Shingal, R., Brandsen, B. H., Ermalov, A. M., Flower, D. R.. Newby, C. W., and Noble, C. J. (1986). J . Phys. B19, 309. Shingal, R., and Lin, C. D. (1989). Phys. Rev. A40, 1302. Shore, B. W., and Menzel, D. H. (1968). “Principles of Atomic Spectra.” Wiley, New York. Sidorovitch, V. A., and Nikolaev, V. S. (1983). J. Phys. B16, 3743. Sidorovitch, V. A,, Nikolaev, V. S . , and McGuire, J. H. (1985). Phvs. Rev. A31, 2139. Skogvall, B., and Schweitz, G. (1990). Phvs. Rev. Letters. 65, 3265. Stich, W., Luedde, H. J., and Dreizler, R. M. (1985). J. Phvs. Bl8, 1195. Stolterfoht, N. (1971). Phys. Lett. 37A, 117. Stolterfoht, N. (1985). Private communication. Stolterfoht, N. (1989). In “Spectroscopy and Collisions of Few Electron Ions” Bucharest. (M. Ivaseu, V. Floreseu, and V. Zoran, eds.) Worldwide Scientific, Singapore, p. 342. Stolterfoht, N. (1990). Physica Scriptu 42, 192. Stolterfoht, N., Miller, P. D., Krause, H. F., Yamazaki, Y., Swenson, J. K., Bruch, R., Dittner, P. F., Pepmiller, P. L., and Datz, S. (1987). Nucl. Ins?. Merhods B24/25, 168. Swenson, J. K . , Yamazaki, Y., Miller, P. D., Krause, H. F., Dittner, P. F., Pepmiller, P. D., Datz, S., and Stolterfoht, N. (1986). Phys. Rev Letters 57, 3042. Tanis, J. A. (1990). In “The Physics of Electronic and Atomic Collisions” (A. Dalgarno. R. S. Freund, P. M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.) p. 538. American Institute of Physics, New York. Tanis, J. A., and Woodland, W. T. (1990). In “The Physics of Electronic and Atomic Collisions” (A. Dalgamo, R. S. Freund, P. M. Koch, M. S. Lubell, and T. B. Lucatorto, eds.) p. 366. American Institute of Physics, New York. Tanis, J. A., Bernstein, E. M., Graham, W. G., Clark, M., Shafroth, M., Johnson, B. M., Jones, K. M., and Meron, M. (1982). Phys. Rev. Lett. 49, 1325. Tanis, J. A., Bernstein, E. M., Clark, M. W, McFarland, R. H., Morgan, T. J., Muller, A,, Stockli, M. P., Berkner, K. H., Gohil, P., Schlachter, A. S., Steams, J. W., Johnson, B. M., Jones, K. W., Meron, J., and Nason, J. (1985). In “Electronic and Atomic Collisions” (D.C. Lorents, W. E. Meyerhot, and 1. R. Peterson, eds.), p. 425. North Holland, Amsterdam. Taulbjerg, K. (1990). Phys. Ser. 42, 205. Taulbjerg, K., Barrachina, R. O., and Macek, J. H. (1990). Phys. Rev. A41, 207. Theisen, T. C., and McGuire, J. H. (1979). Phys. Rev. A20, 1406. Thomas, L. H. (1927). Proc. RoyalSoc. A114, 561. Thompson, J. J. (1913). Phil Mug. 26,792. Tuan, T. F., andGerjuoy, E. (1960). Phvs. Rev. 117, 756. Tulkki, I. (1987). Private communication. Van den Bos, J., and De Heer, F. J. (1967). Phvsica 33, 333. Van den Bos, J., and De Heer, F. J. (1968). Cerreta Physica 40, 161. Van Kampen, N. G. (1981). In “Stochastic Processes in Physics and Chemistry.” North Holland, Amsterdam. Vegh, L. (1988). Phys. Rev. A37,992. Vegh, L. (1989). J. Phys. B22, L35. Vilenkin, N. Ya. (1971). In “Combinatorics,” p. 38. Academic Press, New York. Vogt, H., Schuch, R. Justiniano, E., Schulz, M., and Schwab, W. (1986). Phvs. Rev. Left. 57, 2256. Walters, H. R. J. (1988). In “Electronic and Atomic Collisions’’ (H. B. Gilbody. W. R. Newell. F. H. Read, and A. C. H. Smith, eds.), p. 147. North Holland, Amsterdam. Wang, L. J., King, M., and Morgan, T. J. (1986). J . Phys. B19,623. Wang, Y. D., McGuire, J. H., and Rivarola, R. D. (1989). Phvs. Rev. A40, 3673.
MULTIPLE-ELECTRON EXCITATION. IONIZATION, AND TRANSFER
323
Wang, Y.D., Straton, J. C., McGuire, J. H., and DuBois, R. D. (1990). J. Phvs. B23, L133. Weaver, 0. L. (1990). Private communication. Weaver, 0. L., and McGuire, J. H. (1985). Phvs. Rev. A32, 1435. Williams, J. F. (1986). A m . J. Phvs. 39, 621. Wright, G. R., and Van der Weil. M. J. (1976). J. Phvs. B9, 1319. Zare, R. N. (1967). J. Chem Phvs. 47, 204. Zajfman, D., and Maor, D. (1986). P h w . Rev. Leffers56,320. Zhang, X..Whelan, C. T., and Walters, H. R. J. (1990). J. Phys. B23, L173. Zouros, T. J. M.. Schneider. D., and Stolterfoht, N. (1987). Phv.5. Rev. A35, 1963. Zouros, T. J. M.. Lee, D. H..and Richard, P. (1989). Phvs. Rev. Leu. 62,2261. Zouros, T. M.. Lee. D. H., Sanders, J. M., Shinpaugh, J. L., Tipping, T. N.. Richard, P., and Varghese. S. L. (1990). Phys. Rev. A42,678.
This Page Intentionally Left Blank
ADVANCES IN ATOMIC. MOLECULAR, AND OFTICAL PHYSICS. VOL. 29
COOPERATNE EFFECTS IN ATOMIC PHYSICS J . P. CONNERADE Blackett Laboratory Imperial College London, Great Brirain
I. 11. 111. IV. V. VI. VII. VIII. IX. X. XI. XII.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . Many-Body Effects and the Conservation of Angular Momentum . . . . . Rydberg Series . . . . . . . . . . . . . . . . . . . . . . . . . . Non-coulornbic Potentials and the Periodic Table . . . . . . . . . . . Giant Resonances . . . . . . . . . . . . . . . . . . . . . . . . . Atomic Giant Resonances in Other Environments-Controlled Collapse and Instabilities of Valence . . . . . . . . . . . . . . . . . . . . . . . Giant Resonances in Nuclei and in Atomic Clusters . . . . . . . . . . Are Giant Resonances in the d and f Sequences Atomic Plasmons? . . . . Extending Mean Fields beyond the Hartree-Fock Scheme. . . . . . . . Can One Blow Off Complete Shells by Laser Spectroscopy?. . . . . . . Interactions between Giant Resonances and Rydberg Series-Intershell and Intersubshell Couplings. . . . . . . . . . . . . . . . . . . . . . . Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . References. . . . . . . . . . . . . . . . . . . . . . . . . . . .
. .
. .
. .
325 328 330 333 337 34 1 344 348 352 355 358 364 364
I. Introduction In low-energy physics and, indeed, in physics at achievable energies, the main unsolved problem is the many-body problem. The three-body problem was uncovered by Newton in classical mechanics. In quantum mechanics, exact solutions for the few-body problems are not known, although the existence of closed shells in central fields is extremely helpful in developing excellent approximations. In relativistic quantum field theory, there is no such thing even as a system of just two particles. The atom provides the ideal few-body quantum system: not only can it be isolated as a “free” atom, but it actually consists of a selectable number of bodies (electrons) that interact with each other through the best known law of force, the inverse square law. The only aspect of the problem that then remains unsolved is precisely the one of interest. However, in spite of all these good reasons, many-body effects do not figure prominently in the dictionary of most atomic physicists. In fact, the subject is
325 Copyright 0 1992 by Academic F’ress, Inc. All rlghtr of reprcduclion in any form reserved. ISBN 0-12-M)3829-3
326
Connerade
dominated by the independent particle approximation (IPA), which takes a diametrically opposite point of view, namely that the electrons are dynamically independent. The predictions of the IPA are uncannily accurate over a wide range of energies and phenomena. Indeed, they seem so good that physicists have come to treat atoms as systems whose many-body characteristics have thankfully disappeared from view with the help of some excellent approximations. An important question to ask is why the independent particle approximation works so well for atoms, and whether conditions can be found under which significant breakdown might be expected. At first sight, almost every aspect of the independent particle model, apart from some QED corrections, seems well worked out. The chances of discovering dynamic many-body effects therefore seem rather small. The first ray of hope in this quest is glimpsed when we ask the question: Is there a single atomic mean field that can be used to represent not only ground state atoms, but also excitation and ionisation processes? It soon transpires that there is not, that there cannot be, and indeed that many different choices of potential are appropriate according to circumstances. The issue of how best to represent many-body phenomena in the calculation of excitations is arguably the most important one remaining in atomic physics. One may also ask the question: Should a mean field or an atomic potential be used in the first place? It would be surprising, in a many-electron atom, if electrons could be regarded as even remotely “independent.” Of course, they are not truly so (despite the name) even within the IPA since they are coupled through an averaging over all the motions enshrined in the central field approximation. What one really means is that they are dynamically independent, and the search for cooperative effects can be interpreted as a search for the breakdown in this aspect of the description. There are, however, more subtle points as to the nature of the breakdown. For example, the price one pays, in the self-consistent field theory of the atom, for reducing a many-electron problem to what appears to be a one-electron description is that each nonequivalent particle “sees” a different effective potential, due to averaging over the motions of all the others. Furthermore, when we come to excited states, each configuration (configuration average approximation) and, in some cases, each term (LS-dependent Hartree-Fock approximation) of the atom may require a different potential, at which point (regarding physical validity as opposed to computational convenience) the whole idea of using potentials to represent the atom becomes rather contorted. How could one excite the system from one state to another without completely altering its mean field? What, then, is the use of a potential? The point at which one decides that a description in terms of average potentials should be given up is to some extent a matter of taste, since quantum mechanics, being a linear theory, allows even quite complicated subsidiary excitations to be represented through the superposition of different configurations. As a rough rule, the more the fully self-consistent potential of the excited state differs from
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
327
that of the ground state, the more complicated and contorted the IPA interpretation becomes. When dealing with photoionisation, a further problem is to decide which system the potential should be calculated for, since one may choose the atom (called V N ) ,the ion ( V N - I ) ,or indeed some kind of interpolation between the two. Alternatively, one may prefer to adopt Wigner’s scattering theory, a formulation that has the advantage that it does not depend explicitly from the outset on the choice of wavefunctions but merely on the existence of a Schrodinger-type equation. Precisely how one extends the IPA can all too easily turn into a technical discussion of computational methods, which would obscure the real issue: Are there new e$ecrs due to the presence of many electrons in the system that one might justifiably call cooperative effects? Our first task is to explore what they might be in order to ensure that appropriate experiments can be performed. However, the question posed here does not originate from experiment, but from fundamental principles. An independent particle model, augmented by low-order perturbation theory, successful though this combination may be, cannot fully represent the dynamic response of a manybody system. The experimental (and theoretical) challenge is then: Where are the missing parts of the description or where (in energy) do cooperative effects arise in the spectral response of many-electron atoms? This is a more specific question than merely asking where theory might break down or becomes inaccurate. Rather, one seeks to identify behaviour involving the correlated motion of many electrons and then relate it to an observable breakdown in the independent electron description. One thinks at first of double excitations, which are the most obvious example of a dynamic correlation effect. In this context, the question would become: can the simultaneous excitation of many electrons result in a specific spectral feature (an atomic plasmon)? The uncertainty as to whether an effective single electron potential may or may not adequately represent many-body effects is a familiar feature of manybody theory: one speaks of quasi particles, which resemble particles in all respects except that they are dressed, i.e., that they experience a potential different from that of free particles, and also quite often of the other particles in the system. Usually, this provides a very satisfactory description, which atomic physicists use and accept under the name independent (or “single”) electron theory. For example, an inner shell vacancy or hole in a nearly complete shell is just such a quasi particle. When this description breaks down, one speaks of a many-body or cooperative effect. From this standpoint, the Hartree-Fock model is merely the first term in a series (the linked cluster expansion, see, e.g., Brout and Carruthers, 1963) which, if convergent and pursued to all orders, should represent the many-body system correctly. Unfortunately, the solution to the many-body problem with an inverse square
328
Connerade
law of force is not known, even in classical mechanics. It was observed a long time ago by Landau and Lifshitz (1977) at a fundamental level that a problem cannot be solved in quantum mechanics unless its solution in classical mechanics is also known. The dynamics of the few-electron problem in quantum mechanics are unsolved. A further complication is that, for systems with N > 2 particles, classical motion becomes a nonlinear problem, so that irregular or chaotic orbits may appear simply as a result of the increase in the number of degrees of freedom. Since free atoms with many electrons are the ideal quantum analogue of such a classical system, an interesting (and unsolved) question arises: Do correlations involving many electrons trigger something similar to chaotic behaviour in the spectral response of many-electron systems? Such behaviour, if observed, would clearly signal a new form of breakdown of the independent electron approximation and would also qualify as a cooperative effect. This chapter addresses the issues just raised and attempts to illustrate by experimental examples many-body effects that signal pronounced departures from the independent particle approximation. There are certain aspects that the present chapter will not address. Beyond a brief discussion, cooperative motion involving only two electrons (as reviewed, for example, by Berry, 1986) will not be covered. Attempts to expand the N body problem in terms of a two-electron basis (Sinanoglu, 1960, 1961) do not seem so far to have been particularly fruitful. It seems fair to comment that helium is not the hydrogen of correlations and that the ideas needed to understand and to classify doubly excited states in He (Herrick, 1978) translate even less into more complex spectra than do the properties of hydrogen for the development of the independent particle picture. Likewise, the double excitation, double escape, and Wannier ridge problems (Wannier, 1953; Fano, 1976) are interesting as a manifestation of the breakdown of the independent particle model but are not central to the subject of multielectron correlations. We shall concentrate on the case where an atom contains several subshells. In other words, we are interested in the atom as a many-body system rather than in the long-range correlations of the three-body problem.
II. Many-Body Effects and the Conservation of Angular Momentum A historical approach to the subject of cooperative effects in light atoms (specifically, in two-electron systems) has been given by Berry (1986), who has attempted to project out from the best available correlated wavefunctions for systems with two outer electrons, probability distributions that would somehow convey the nature and symmetry of their correlations.
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
329
The origin of the subject can be sought in an early controversy between Bohr and Langmuir: when setting up the aufbau principle, Bohr (1913) took the bold step of supposing that the angular momenta of individual electrons could be quantised. It was remarked by Langmuir (1921) that this assumption was unlikely to be quite correct since, in many cases, electron-electron interaction would be significant, so that it might be more reasonable to quantise, e.g., the angular momentum of a pair of electrons. In his own words: Bohr assumes that the angular momentum of every electron round the centre of its orbit is h. However, it is an attractive hypothesis to assume that, in the case of coupled electrons, the quantum theory is concerned not with the angular momentum possessed by one electron but rather with the angular momentum which, by being transferred from electron to electron, circulates in each of two directions about the nucleus.
Various models for helium were developed using the old quantum theory. (See, e.g., Kramers, 1923.) However, Bohr’s early point of view soon predominated even for many-electron systems. At the outset neither Bohr nor Langmuir could have suspected the reason why Bohr’s approach was successful, since it really emerged from the new wave mechanics: by a happy coincidence, the existence of closed shells, the Pauli principle, and the mathematical simplicity that the biaxial theorem introduces in sums of spherical harmonics all conspire to save the central field approximation, within which the conservation of angular momentum is the natural approximation. In this respect, the quantum many-body problem is a good deal simpler than any classical analogue. However, Langmuir’s (1921) approach was indeed a sensible one: when we examine how many-body effects can arise even as a breakdown of the independent particle model, the most likely cause is a breakdown of the angular momentum labeling for individual electrons. Formally, this is how configuration mixing and two electron jumps are introduced in the standard atomic theory. Recently, there has been a great revival of interest in the study of the dynamic aspects of two-electron correlations. It has borne fruit: much progress has been made in the search for new quantum numbers to describe the two-electron system (e.g., Lin, 1974; Herrick, 1978; Berry, 1986). The essential conclusion of the work of Berry (1986) and others is that the electronic states of atoms with two outer electrons in a closed subshell exhibit collective rotations and vibrations rather than predominantly independent particle behaviour. Thus, the low-lying excited states of helium and the alkaline earths resemble collective rotor-vibrators. They are not the quantum analogue of solar systems. The purest independent particle system, hydrogen, possesses +fold degeneracy of its levels because of symmetry under the O(4) rotation group. The independent particle symmetry of the two-electron problem would therefore be
330
Connerade
O(4) x 0(4),but this is broken by correlations. When electron-electron repulsion is included, the appropriate symmetry turns out to be O(4) with an additional invariant (A, - A2I2,where A, and A, are the Runge-Lens vectors of the two electrons, i.e., vectors proportional to the Kepler ellipticities of their orbits, and pointing in the directions of their semimajor axes. The invariance of /A, - A$ signifies simply that the ellipses are “locked,” i.e., that the orbits precess together (Herrick and Sinanoglu, 1975). This is an example of cooperative motion. It was also noted (Nikitin and Ostrovsky, 1976, 1978) that the principal quantum numbers n , and n2 must be the same for correlations to be strong, because, if n, is very different from n 2 , one simply reverts to the independent particle model. Thus, electrons from the same shell are the most highly correlated. It is also found that valence and inner valence electrons are more highly correlated than inner electrons. This arises because they are held out from the deep core region and can never achieve a high kinetic energy. Consequently, their motion is much affected by electron-electron repulsion within the subshell. There has been much interest in manufacturing double Rydberg states with n, n2 that would exist as large, highly correlated states of the atom. This leads to the double-excitation/double-escapeproblem and manifestations of the longrange correlations of the three-body system which, as stated in the introduction, are not part of the main thrust of this chapter. How much of what has been said in this section concerning cooperative motion can be transferred to the many-electron problem is an open question. There can be little doubt that progress in extending the considerable increase in our understanding of two-particle collective motions to the full N-electron atom would constitute the most significant breakthrough in atomic physics since the work of Hartree. There might also be important implications for theoretical chemistry: the electronic shells of some atoms (as remarked by Berry, 1986) might possess an internal geometry that would then only need to be rotated in order to form directional bonds.
-
III. Rydberg Series We are primarily concerned with excitations in many-electron atoms, and we must first decide what the spectral signature of a clean independent electron system is likely to be. The one-electron atom has the signature En = E , - R/n2, where E, and R depend only on fundamental constants. When we move to a many-electron system, the following correlations are required: (i) There is no longer a single ionisation potential E,, i.e., several series limits occur for excitation from the ground state. (ii) These series limits no longer depend only on the fundamental constants, but become properties specific to the atom (dressing
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
33 1
of the quasi particle called the vacancy), whereas only small numerical corrections to R are required. (iii) Finally we introduce a quantum defect that appears as a new quantity in the Rydberg formula and that is approximately constant for all the excited states nl in which 1 is constant. These Rydberg states form the discrete part of a single-excitationchannel. The quasiconstancy of the quantum defect for a given angular momentum I signals that, to a good approximation over most of the space occupied by the excited electron, angular momentum is conserved. It is intuitively obvious, and is established by a beautiful mathematical theory (Seaton, 1966a), that the region of conservation of angular momentum lies outside a core, and that the region of breakdown in which the electron is scattered by many-body interactions lies inside the many-electron core, where the potential is non-coulombic. The origin of the quantum defect p is well understood, and the mathematical formalism for it has been reviewed so often in the literature (see, e.g., the elegant and pedagogical review by Starace, 1976) that there is no point in repeating it here. Rather, it may be useful to extract some simple physical arguments that are relevant to our theme. The near-constancy of p for a given 1 can be traced back (as remarked by Mulliken, 1964) to the recapitulation of atomic orbitals, i.e., to the fact that the nodes of the wavefunctions, as n increases, tend to repeat at nearly the same radial distance from the nucleus. This distance is modified in the presence of a core of finite radius r, , when a matching condition of the coulombic (large r) and core (small r ) wavefunctions is applied at r,, and thus the quantum defect p can also be interpreted as resulting from a phase shift, generalised to bound states of high n. It is then a fundamental result of quantum defect theory that this phase shift v p joins smoothly to the phase shift 6 of the adjoining continuum states. Taken together, the Rydberg series and its adjoining continuum form a whole, described as an excitation channel. It will emerge later in this chapter that continuity across the threshold (first noted by Sugiura, 1927, for hydrogen, and extended by Gailitis, 1963, and Seaton, 1966b, to other atoms) is a fundamental property, valid even when the constancy of the quantum defect breaks down: if a channel is perturbed or distorted by many-body interactions, oscillator strength may actually be transferred across the threshold, from the continuous to the discrete part of an excitation channel or vice versa. This principle has been verified experimentally even for a situation where a resonance straddles the threshold (Connerade el al., 1988). Because of (i) and (iii) together, it is also possible to find situations in which two channels will overlap significantly in energy. Elementary perturbation theory tells us that the coupling between two channels increases. dramatically if the difference in energy between the channels is small. If the electrons move nearly independently of each other, we expect that the two quantum defects p , and p2 remain nearly constant. (See the dashed lines in Fig. 1.) However, if the electrons are not independent, the individual quantum defects cease to be constant
332
Connerude 1
-
h
71
.
Z O
v CI
s
0
0.5 1.0 h/z(mod 1) FIG. I . A two-dimensional plot of quantum defects representing the interaction between a doubly excited and a core-excited channel of Yb I. In the absence of any coupling between the channels, the quantum defects would be constant and follow the dashed lines. Correlations are responsible for the avoided crossing.
and, in cases of moderately strong coupling involving just two channels, can result in the simple graph shown in Fig. 1. It is a major achievement of multichannel quantum defect theory that patterns of the kind shown in Fig. 1 and indeed much more complex ones (see, e.g., Armstrong et ul., 1977) are interpretable in detail using but few adjustable parameters. When p ceases to be approximately constant as a result of interchannel coupling, we are dealing with a breakdown of the independent electron approximation, i.e., the first step beyond the standard atomic model referred to previously. Quantum defect theory confines the many-body effects within a core region, but deals with excited states that lie well outside r, . Thus, the excited electron does not have much opportunity to “see” rearrangements involving many electrons. It is therefore a natural question to ask whether the many-body effects that are beginning to emerge in perturbed Rydberg series might become much more dramatic if somehow the excited state were more compact or loculised. In the language of quantum mechanics, one looks for situations in which the spatial overlap between the initial and the final-state wavefunctions is as large as possible. The hope is that such cases will favour the observation of dynamic rearrangements within the atom. Unfortunately, the near-constancy of quantum defect, which served as a “background” against which many-body effects could be recognised, then disappears from the problem. The constancy of quantum defects is definitely a feature of extended Rydberg orbitals and does not apply to compact states. Thus, we lose an important reference point when we try to make many-body effects large. The familiar Rydberg series becomes unrecognisable because the concept of a quantum defect is not truly applicable. In other words, the regularity of the spectrum no longer exists.
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
333
There are other ways in which Rydberg series may become absent from the spectra of many-electron atoms. For example, if a dramatically large resonance appears in a Rydberg channel, then the sum rule for oscillator strengths implies that oscillator strength must be redistributed from elsewhere in the spectrum: thus, the presence of a strong resonance in the continuum would imply that the intensity of the discrete transitions would be much reduced. Another form of disappearance arises through lifetime effects involving the series limit or parent ion state in cases where it lies above the double-ionisation threshold: if the lifetime is shortened by Auger decay, then once the lifetime broadening becomes comparable to the energy interval between two successive series members, all higher members of the Rydberg series are wiped out. Interestingly, it is also possible for autoionising states to be “stabilised” by interactions, so that they look much narrower (longer lifetimes) than other Rydberg members within the same channel, i.e., that they do not conform to the rAu,olonlsa,,on/n*3 expected within regular law for lifetime breadths A r = rAugcr a series of resonances. A spectacular example of this stabilisation effect is to be found in the spectrum of CaI, where interference between two different doubly excited channels leads to a strong perturbation, and since each of the excitation channels itself arises by breakdown of the 1 characterisationof the outer electrons of Ca, the mixture of the two channels implies that the one electron labels are even further corrupted. Likewise, intensities may be enhanced among upper members at the expense of lower ones, an effect that is ubiquitous in perturbed series. All the preceding effects, which lead to the spoiling and eventual disappearance of Rydberg series, are many-body effects. In a very general sense, we can regard the “quality” of Rydberg series and the accuracy with which their number is predicted by the independent electron model as the best markers for the significance of many-body effects within a given spectral range: the longer and more regular they are, the less the situation is interesting as far as extracting manybody effects is concerned. We thus arrive at the idea that non-Rydberg specfroscopy is the study of many-body effects (Connerade, 1978a, 1984).
+
IV. Non-coulombic Potentials and the Periodic Table As we have seen, the conservation of angular momentum outside the core and the occurrence of closed subshells within the core underpin the aufbau principle and are fundamental to our understanding of the Periodic Table. Since we are searching for breakdown in the independent electron scheme, we should begin by asking just how good this understanding is, and whether it really rests completely on the aufbau principle, or whether non-coulombic effects due to the presence of many electrons in the core must also be taken into account. Elementary textbooks present the explanation of the Periodic Table of the
334
Connerade
Elements as one of the early triumphs of quantum mechanics-a perception so widespread that one often forgets to qualify this statement, which is only strictly applicable to the first three periods. The aujbuuprinzip is usually attributed to Bohr, who actually divided the electrons equally among the subshells. However, it was really Stoner (1924) who first put the appropriate number of electrons in each subshell and arrived at the correct structure. The Bohr-Stoner principle was of course worked out before modern quantum theory, and so the notions of an independent particle model, with a static mean central field and the use of spherical harmonics to represent closed shells, were incorporated later. Even so, the argument neglects important features of the radial equation and only works properly for the first few rows, where the ordering of the levels in the many-electron system is not too markedly different (apart from the lifting of 1degeneracy) from that of hydrogen. Beyond argon, matters become more complicated, and the qualitative, semiclassical notion of screening that has to be introduced as an ad hoc supplement of the aufbau principle (see., e.g., Herzberg, 1935) is quite unsatisfactory, as was first noted by Fermi (1928). The elements of a proper understanding were advanced by Goeppert-Mayer (1941) following Fermi’s (1928) suggestions, but the numerical methods then available did not suffice to follow the explanation through convincingly, and it was only with the progress made possible by the advent of fast computers that a quantitative explanation became possible (Griffin eral., 1969, 1971). The crucial point is that the 1-dependence of the radial Schrodinger equation manifests itself in subtle ways that can be understood by quantum mechanics but not by semiclassical models based on progressive screening. Interestingly, it is a manifestation of elementary quantum mechanics (the filling properties of shortrange asymmetric potential wells) that governs the sudden d and f subshell contraction, and it is the non-coulombic nature of the radial potential (i.e.. a manybody effect which can partly be “absorbed” into a static mean field) that controls the filling of the long periods. The basic physics of the lanthanide contraction is associated with the emergence of a double-well effective radial potential due to the combined effect of a non-Coulombic core and the positive, repulsive centrifugal term in the radial l)h2/2rnr2.If we consider, Schrodinger equation. This term has the form l(1 first, the case of hydrogen, it is clear that the potential at large r must be dominated by the attractive coulombic part which falls off only as llr, whereas the potential at small r is dominated by centrifugal repulsion. Thus, the high 1 states in hydrogen live in a potential well with a more or less hard center depending on the value of I, which is why the f electrons are held well away from the nucleus. is All atoms resemble hydrogen at large r (i.e., outside the core radius r,-this the basis of the quantum defect theory discussed in Section III), and this accounts for the outer well of the 1 = 3 double-well potential in the lanthanides.
+
335
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
As the electron penetrates further inward (inside r o ) ,the screening due to the core electrons becomes less effective, and there is a very rapid increase in absolute magnitude of the attractive Coulombic term, so that the effective potential becomes binding again. At very small values of r, the repulsive centrifugal term must eventually win again, since it depends on l/r2, and this accounts for the inner well of the double-well potential. Asymmetric short-range wells have very different properties from long-range wells: the former may be completely empty of states and will suddenly acquire bound states, one by one, as the binding strength of the well is progressively increased. The number of bound states is always finite. This is a purely wave mechanical effect, not a consequence of semiclassical screening theory. Longrange wells, on the other hand, always contain an injinite number of Rydberg states. The atomic double-well potential contains both, separated by a potential barrier. As the atomic number is increased, a state may thus suddenly drop from the outer into the inner well, where filling is energetically very favourable. The Mayer-Fermi theory of the lanthanide contraction explains (i) why the filling of the long periods occurs out of the expected order for quite high 1 values (1 = 2 and 3), (ii) why the effect is sudden and depends critically on atomic number, (iii) why this filling, when it occurs, happens deep inside the core rather than in the outer reaches of the atom, and (iv) why there is s-d competition, and so on. Since the Mayer-Fermi model is covered by earlier reviews (Connerade, 1978a, 1984; Griffin et al., 1987), we concentrate here on aspects that are relevant to possible breakdown of the quasi-particle approximation, i.e., to the characterisation of vacancies. As an example of the problems that can arise, consider the ionisation of atoms in the long d periods of the periodic table. From available information on the ground states of the atoms and ions, we can construct a table of configurations (Table I). It is a corollary of the aufbau principle that, if one electron is removed from TABLE I TABLEOF GROUND CONFIGURATIONS OF THE ATOMSA N D IONS IN THE FIRSTLONGPERIODS OR TRANSITION ELEMENTS Ca sc Ti V Cr Mn Fe
co Ni cu
zn
. . . 4s’ 3d4’ 3d24s2 3d34s2 3d54s 3d54sZ 3db4s2 3d84s2 3d84s’ 3dIU4s 3dI04s2
. . .4s 3d4s 3d24s 3d4. . . 3d5. . . 3d54s 3d64s 3d8. . . 3d9. . . 3dI0 3d1°4s
Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd
. . . 5s’ 4d5s2 4d25s’ 4d45s 4dS5s 4d55s2 4d75s 4d85s 4 d i 0 .. , 4dL05s 4dIU5s2
. . . 5s . . . 5s’ 4d5s 4d4. . . 4dS.. . 4d55s 4d7. . . 4ds . . . 4d9. . . 4diu 4d%
Ba La Hf Ta W Re 0s Ir Pt Au Hg
. . . 6s’ 5d6s2 5d26s2 5d36s2 5d46s2 5d55s2 5d66s2 5d76s* 5d96s 5dI06s 5dlu6s2
. . . 6s . . . 5d2 5d6s2 5d16s 5d46s 5d56s 5d66s 5dY 5d10 5dI06s
Connerade
336
the atom to produce an ion, the ordering on the nl levels should remain the same in the ion as it is in the atom. Table I, which of course is based on experimental information, shows clearly that the ordering for atoms and for ions is not always the same, i.e., that the periodic table for ions is not, in all cases, simply obtained from the one for atoms. There are four elements (V, Co, Ni, and La) from which it is impossible to remove one electron leaving the ion in its ground configuration. Quite clearly, if we accept the nl independent electron labels as meaningful, excitation of one electron cannot occur without a rearrangement of the system. In fact, it is pretty clear that the individual nl labels are not very meaningful in these cases and that the independent electron description is not satisfactory. Situations similar to the one just described can be created by inner-shell excitation: with an inner shell excited, barium resembles lanthanum as far as its outer electrons are concerned. Consequently, they seek to arrange themselves in a configuration involving 5d electrons, such as 5d6s or 5d2rather than remaining 6s2 as in the ground state. In other words, a simple, one-electron excitation scheme becomes impossible. Suppose we look at the resulting ion state (series limit) rather than at the full Rydberg series, for example, by using photoion spectroscopy rather than photoabsorption. We then find that the situation is much as shown in Fig. 2: each channel excited results in a single line, and there are essentially three possible situations. (a) The incident photon may only possess enough energy to excite one electron. This is optical spectroscopy, and is unlikely to yield much information on many-body excitation, since the energy is insufficient to drive Electron Affinities
Valence
Subvalence
~
Core
Photon Energy FIG. 2. (Schematic) The difference between spectra due to core excitation, subvalence excitation, and the excitation of optical electrons as it emerges from graphical solutions of the many-body Dyson equation. (See Connerade, 1984, and references therein for further details.)
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
337
cooperative modes. Only one line is observed, corresponding to the excitation of one particle. (b) The incident photon may possess enough energy to excite the deepest shells. This is X-ray spectroscopy and is, again, unlikely to provide much information on many-body effects. The reason here is that the initial state wavefunction is much smaller than that of the final state, so that the electron escapes very fast, leaving behind a well-characterised vacancy from which it is screened by all the intervening shells. Experimentally, one sees a single line, surrounded by a few satellites, which are a measure of the small breakdown of the quasi-particle approximation. (c) The incident electron may possess sufficient energy to excite a subvalence electron, but not enough to excite a deep shell. In this range, the spatial overlap between the vacancy and the excited-state wavefunction is greatest. Rearrangements can occur while the electron is escaping, and instead of a single line one then finds a jumble of features into which the quasiparticle line has fragmented. The preceding description is backed up both by experiments and by calculations, which all points to the fact that the vacuum ultraviolet and soft x-ray ranges of the electromagnetic spectrum are ideally matched to the study of manybody effects in atoms.
V. Giant Resonances The soft x-ray spectra of atoms, molecules, and solids corresponding to the transition elements and rare earths contain remarkably strong, isolated features that can be traced back right through the sequences (i.e., horizontally across the Periodic Table) past the onset of the long sequence to the corresponding rare gases, in which they appear as broad resonances in the continuum. These features possess many of the properties that have been listed previously as desirable or favourable to many-body excitation. By analogy with collective oscillations in atomic nuclei, they have been baptised “giant resonances,” and their study has been regarded as an ideal testing ground for many-body theories of the atom. The giant resonances can be understood (at least qualitatively) as arising from the presence of the short-range inner well in the effective potential of the excited electron, i.e., as a consequence of the non-Coulombic core. It is clear that, when a short-range well suddenly acquires a new bound state, it must come from somewhere. Quantum mechanics shows that it comes down from the continuum, where it exists as a resonance just before the short-range well becomes sufficiently binding to hold a new discrete state.
338
Connerade
Since the wavefunction of the excited electron can become resonantly localised in the inner region, the electron takes longer to escape than if it were promoted to a pure continuum state in a Coulombic potential, and it therefore can witness rearrangements within the atomic core. In fact, the spatial overlap between the initial- and final-state wavefunctions can be so large that the oscillator strength sum rule is exhausted within the broad feature, which therefore deserves the name giant resonance. When all the oscillator strength is gathered up into this resonance, no Rydberg series is observed in the same channel, i.e., the discrete part of the spectrum is quenched. When the sum rule is exhausted, this means that, for the 4d + ef resonance in La, for example, integrating over the feature yields an oscillator strength of 10, corresponding to the presence of 10 electrons in the subshell. In this sense, the whole shell is involved in the excitation. It is also the case that calculations of these features in Xe, Ba, and La become extremely sensitive to the mode of approximation used. In the many-body perturbation theory (MBF’T) applied to atoms (Kelly, 1964, 1966), an approach in which the many-body perturbation series is treated order by order and all the different interactions, represented by their Feynman graphs, are included for each order and evaluated on a complete independent particle Hartree-Fock basis of states, only a finite number of orders can be included for practical reasons. In the random phase approximation with exchange (or RPAE), a certain class of diagrams (the forward “bubbles” and their exchange counterparts) can be summed to all orders algebraically, but all diagrams cannot be included. It is also possible to perform calculations with suitably tailored Hartree-Fock potentials tuned specifically to obtain agreement in this excitation channel. All these many-body approaches lead to results significantly different from each other and very different from the configuration average approximation. The many-body theories are in broad agreement with experiment (see Fig. 3) in the critical example of Ba as well as in other nearby element such as Xe and La. Unfortunately, the situation is not simple and it is not the case that a glance at Fig. 3 will immediately allow one to distinguish which is the best theory, because there are other variables (partial cross sections, angular distributions of photoelectrons, photoion production rates, etc.) as well as many other atoms that exhibit many-body effects. Also, the experimental data in Fig. 3 were not obtained for free atoms, but for a solid in which structure due to reorganisation of the 6s2 electrons (see Table I) is not important. When an inner-shell vacancy is created in Ba I, many complex mixing effects involving the outer electrons occur and these are not included in the theory as of early 199 1 . There is, however, a purely experimental approach that can be used to estimate the importance of correlated motions in giant resonances, and this consists in studying the relative production rates of different stages of ionisation as one tunes
339
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
/ 100
110
120
130
140
150
w (eV) FIG. 3. Comparison between different many-body calculations (each model has its own acronym) and experimental data for the 4d + Ef giant resonance of Ba I. Note that the experimental data used by the theorists are in fact for polycrystalline Ba, which is preferred because the complex doubly excited structures that interfere with the giant resonance cannot be readily included in the theoretical model and are absent from data for the solid. The origin of the second bump on the experimental curve is unclear. It might be due to backscattering of the photoelectron in the solid. (a) The comparison published by Kelly er u / . (1982). (b) The comparison published by Amusia et a/. ( 1990a).
through a giant resonance. Figure 4 shows some examples taken from studies by Zimmermann ( 1989) in Berlin. What emerges from these studies is that the 4d --f Ef giant resonances behave differently at both ends of the 4f sequence. Near the beginning of the sequence (around Ce), ionisation is dominated by the production of triply charged ions,
Dv
I
It
h
$
n L KJ
v
b
30 40
o(4F)
n E
20
30
o(4F) o(4D) o(5P) 45s)
o
+ 0
v
b
30 10
0
90
100
110
120
130
energy [eV]
140
150
80
100
120
140
160
180
energy [eV]
FIG. 4. Photoion and partial yield cross sections for (a) Ce and (b) Dy after Zimmermann (1989).Note that, in (a) multiple ionisation dominates, whereas single ionisation dominates in (b), and that this is accompanied by a qualitative change in the shapes of the profiles from the typical form for a giant resonance to a Beutler-Fano autoionising profile. It is also interesting that all the partial cross sections are enhanced in the vicinity of the giant resonance.
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
34 1
but as the atomic number increases, the level of doubly charged ions becomes more important (at Dy), until, by the time one reaches Dy, the production of singly charged ions is the dominant process, as is to be expected in photoionisation if the independent-particle model is valid. An interesting and very suggestive change is apparent in these spectra, which had also been noticed by Radtke (1979) in earlier photoabsorption experiments. The characteristic asymmetric profile of the giant resonances (tall asymmetric features with a low energy minimum tied to the 4d thresholds and a long tail toward high energies), which can be parametrised by a simple model based on scattering theory (Connerade, 1988) occur around Ce but gradually turn beyond Dy into a typical autoionising profile (Fano, 1961) with a pronounced minimum adjacent to the resonance energy, characteristic of quasi-discrete state decaying into a photoionisation continuum. Thus, the experiment signals clearly that the cooperative effects are most significant toward Xe, and that, as the inner well becomes more binding with increasing atomic number and the Ef resonantly localised continuum state gradually turns into a 4f resonance, its cooperative nature dissipates and it becomes a more normal singly excited feature. In some ways, this picture is quite satisfying, but when one carefully examines the data for Ce (Fig. 4), one may wonder why, if multiple ionisation is to be the marker for cooperative behaviour, quadruple ionisation is not more significant. The answer seems to be that the outermost three electrons of Ce (with ionisation potentials of 5.5387, 10.85, and 20.198 eV, respectively) are fairly easy to shake off, that the fourth, at 36.758 eV, is significantly more difficult, but that shaking off any more than four requires one to break a closed shell open and is energetically inaccessible at primary photon energies of around 120 eV.
VI. Atomic Giant Resonances in Other Environments-Controlled Collapse and Instabilities of Valence As we have repeatedly stressed, one of the important criteria for giant resonances is that the spatial overlap between the initial and final states should be very large, so that the transition takes place well within the atom. This criterion was arrived at by our desire to favour collective processes as much as possible. There is, however, another consequence of the strong localisation of the excited states in giant resonances, which is the survival of the spectral features in a variety of different environments. If we consider a given atom A, then a molecule involving A or a solid composed of atoms A or containing an impurity A or indeed a surface with atoms A upon it can all be expected to
342
Connerade
exhibit a strong feature at the energy of the giant resonance, the spectral profile of which may exhibit distortions reflecting the change in the environment of the atom. Potentially, this is a powerful probe of different phases of matter, and is important also as a bridge between disciplines. Readers interested in this aspect of giant resonances will find the subject well covered in the proceedings of a summer school devoted to this subject (Connerade et al., 1987). In molecules, if we imagine atom A surrounded by a symmetric array of identical, electronegative halogen atoms (a halogen “cage”) that tend to draw electrons away from the nucleus (Dehmer, 1972), then some control over the balance between the inner and outer wells of the double-well potential can be achieved by altering the molecular environment (e.g., using different halogens). This can lead to changes in both width and energy of the giant resonances, and can even lead to sharper resonances than in A, as occurs for 4d + Ef excitation in LaF, (Connerade et al., 1980). Returning to the description of the Mayer-Fermi model of the lanthanide contraction, we may ponder whether the binding strength of the inner well could be altered in order to control the transfer of the electron from the outer to the inner well, and indeed whether an electron could remain poised, with part of its wavefunction in the inner and part in the outer well. This problem is referred to as conrrolfed collapse (Connerade, 1978b,c). In principle, many different external fields can be used for this purpose: laser excitation of outer electrons has been suggested and would enable the finest control, but laser ionisation (Lucatorto et al., 1981) and even the properties of the chemical environment (Maiste et al.. 1980a,b) have been experimented with to good effect. There is a single naturally occurring instance in which the nf electron possesses a bimodal wavefunction with a comparable probability of being found in either potential well. This situation has been analysed by Connerade and Mansfield (1975). (See also Section IX in this chapter.) Under these circumstances, all the oscillator strength is not concentrated into a single feature, and a Rydberg series re-emerges, but with a grossly distorted trend of the quantum defect, because recapitulation of the atomic orbitals breaks down. Connerade (1982) has developed quantum defect theory for a double-well potential by exploiting effec~ than p tive range theory, and has shown that the function coth(1 - p ) rather itself should be plotted against energy so that the threshold relation between the phase shift in the continuum and the quantum defect in the bound states can be generalised to take Levinson’s theorem into account for the short-range well. Difficult situations arise when the atomic orbital is on the verge of collapse or is only just collapsed. In such situations, term sensitivity can be large, and it is not even certain whether the self-consistent field method is entirely reliable or whether two viable solutions might not exist in some cases, as argued by Band and Fomichev (1980). If so, one needs to decide whether they both have physical validity (which is quite conceivable in connection with intermediate valence) or
343
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
whether they are artifacts of the numerical approach. This rather speculative question is discussed by Connerade and Karnatak (1989). Yet another kind of molecular giant resonance can be envisaged that does not rely simply on excitation in an atomic subshell, but involves excitation of electrons from valence shells. Thus, the molecules NZ,CO, NO, and 0, have been considered from this point of view by Lefebvre-Brion (1987), who plots the relationship between the width r and the resonance energies, and finds a behaviour similar (but not identical) to the universal curve for atoms (Connerade, 1984). As she emphasizes, the molecular nature of the resonances emerges from the fact that no resonance at all occurs in the united atom cases. It seems unlikely (see also the discussion in Section VII) that such excitations could lead to pronounced collective modes when the fields are not spherical. There have also been reports of giant resonances in large molecules (see, e.g., Brint er al., 1985) and the general question of whether such features might generally occur in even larger systems of appropriate symmetry (such as cyclopropane, cyclohexane, neopentane, or uranium hexafluoride) has been considered by Robin (1985, 1986). It is likely that each case should be examined separately, since the conditions relating to stability can vary widely and the mechanism involved can be purely molecular in some cases. In solids, giant resonances play an important role as they are often to be found in X-ray spectra and their existence underpins such useful practical tools as resonant photoemission. Even in such apparently unrelated areas as the photonstimulated desorption of hydrogen from surfaces (important to both hydrogen storage and catalysis), quasi-atomic giant resonances are to be found. Also in the condensed phase, an interesting situation exists for d f excitation in the lanthanides, where instabilities of valence may occur that can be probed directly by the excitation of giant resonances, since they relate to the number off electrons that are localised, which may depend on the environment. This is usually discussed in the context of Anderson localisation, but quasiatomic models have also been proposed in which the short-range inner part of the double-well potential, which survives in the solid, plays a crucial role. Experiments on clusters have been attempted, in which the instability of valence is tracked from the atom to the condensed matter limit by varying the size of the clusters. Readers interested in this subject may wish to read Connerade and Karnatak (1989). Finally, one should remark that many of the ideas involved in giant resonances are closely related to models used in the physics of high-density plasmas to describe the electron spectrum near the continuum. The problem here is to treat the energy levels of a disordered system. In one category of models, the central field approximation is used, with the ion represented as a point charge at the center of a spherical cavity in an otherwise uniform positive-charge distribution. A Thomas-Fermi model (More, 1982) or a model based on the Dirac equation
-
344
Connerade
followed by suitable averaging (Liberman, 1979) can be used. One finds narrow resonances near the base of the continuum (especially for large angular momentum t; the resonances at higher energy are broad and eventually become indistinguishable from the continuum. More elaborate models treat a small cluster of nearby atoms immersed in the plasma background by multiple scattering theory. The subject is reviewed by More (1982).
MI. Giant Resonances in Nuclei and in Atomic Clusters At this point, it is useful to consider the subject of giant resonances more generally, nor merely confining the discussion to electronic shells of atoms in diverse environments, since collective oscillations occur in a wide variety of situations, some of them indeed more favourable than single atoms to the excitation of collective modes. Historically, the earliest examples of giant resonances were found in nuclei, where the first use was made of dynamic collective models for the interpretation of resonances of sperical shells. The resonances were observed by Baldwin and Klaiber (1947, 1948) in photoabsorption experiments to study induced radioactivity using gamma rays of around 17 MeV, when cross sections about two orders of magnitude higher than expected were discovered. Giant resonances in nuclei are attributed to collective oscillations of the protons against the neutrons. Several modes are found, the most important ones being the giant electric dipole deformation, the giant quadrupole surface oscillation, and the giant monopole breathing mode. One also distinguishes between the protons and neutrons vibrating together (isoscalar mode) or in opposition (isovector mode). Absorption experiments usually detect giant dipole resonances which are the purest case of a giant resonance and will also be the situation of interest here. The subject of giant resonances in nuclei has a long history which was reviewed by Berman and Fultz (1 975). The nuclear dipole resonances exhibit single (for spherical nuclei) or double (for spheroidal nuclei) peaks depending on the forms of the shells and are seen even for very small nuclei such as 4He. It is also interesting that single-particle (or, in the language of many-body theory, excizonic) structures occur, giving sharper peaks that are superposed on the broader collective resonances and may interact with it. A closely similar situation for atoms has been considered by Connerade and Lane (1987) within the framework of K matrix theoryoriginally developed for the interpretation of nuclear resonances. The analogies between the fields are therefore very clear. The damping mechanisms of the collective resonances in nuclear physics have also been studied: they decay mainly through Landau damping and coupling to nuclear shape oscillations (Bertsch et al., 1983).
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
345
Thus, giant resonances in atomic nuclei occur at energies of around 20 MeV and in electronic shells of atoms, at energies of around 100 eV. The former are held together by a very short range potential, leading to a shell structure and resonances well accounted for in the shell model of Nilsson (1955), while the latter exist in a hybrid, double-valley potential which is short-range in the inner well but long-range at large radii. In the former case, different modes have been detected corresponding to departures from sphericity of the oscillating shell. This has not been found in atoms, for reasons that will emerge more clearly below. A completely new class of giant resonances has been discovered in atomic clusters. They occur in nearly spherical metallic clusters, typically those formed by alkali atoms, which are found to be particularly stable around the magic numbers 8, 18,20, 34,58,92, etc. This sequence of numbers is well accounted for (see, e.g., de Heer et al., 1987) within a spherical jellium model (Chou et al., 1984) or by using self-consistent pseudopotential local density functional calculations (Martins er al., 1985). an approach closely related to the very successful many-body methods developed by Zangwill(l987) for the calculation of atomic giant resonances. One can also start from a simple model such as the Woods-Saxon potential (Woods and Saxon, 1954), which Knight et al. (1984) used in their early work on shell closure in spherical clusters. Thus, stable closed shells of delocalised electrons have been shown to exist in certain clusters, reminiscent of the shells in nuclei and in atoms, but on a yet much larger scale. Just as the closed shells in nuclei or in atoms may exhibit a collective mode, so they do in near-spherical clusters (Hecht, 1979; Kreibig and Genzel, 1985; de Heer et al., 1987), and observations of the resulting giant resonances have been extended to free clusters (Selby et al., 1989) which may even be mass selected (Brbchignac et al., 1990). Experimentally, the giant resonances found by laser spectroscopy in mass selected alkali clusters are among the most impressive examples of collective oscillations yet discovered. Indeed, this is a very clean situation, not far short of the ideal as far as characterising collective resonances is concerned. The cross section rises to an extremely high value within a rather narrow resonance (see Fig. 5 ) which dominates the response, and detailed spectroscopy shows no sign of excitonic (single-particle) structures within the range of the resonance. Since the giant resonance is narrow, interference between different modes of oscillation can sometimes be observed, and is traceable to ellipsoidal modes when there is a slight distortion from spherical symmetry, a situation also encountered in nuclear physics. The work of Selby et al. (1989) shows that the closed-shell clusters exhibit a single plasmon resonance, while, for open-shell clusters, the resonance splits up into several peaks. This fact, added to the interference between the different spheroidal modes observed by Brbchignac et al. (1990) and to the general increase in spectral complexity with departures from the spheroidal model (which can be studied within a simple harmonic oscillator model by introducing at least
346
Connerade
FIG. 5. Giant collective resonances in the photoevaporation cross sections of the clusters (a) K; and (b) KL after Brbchignac er al. (1990). Note that, although the resonance frequency is not much affected by size, the cross section is enormously increased and is essentially the area presented by the cluster.
two oscillation frequencies, R,, and Q, along or perpendicular to an axis of symmetry; see the treatment by Clemenger, 1985) shows that the strong, isolated collective resonance is a property of spherical systems, a fact that will be relevant to the discussion of recent results in strong laser fields. (See Section X.) Clearly, this situation is very similar to the one in atoms. It is incidentally of some interest (but does not yet seem to have been investigated) that there exists a theoretical upper limit on the size of a convex spherical cluster composed of absolutely equivalent atoms. It is a general result of group theoretical geometry (see, e.g., Coxeter, 1961) that not more than 20 objects can be arranged on a sphere as vertices of a regular solid with all faces identical (the regular dodecahedron that is the largest of the five platonic solids). If one relaxes the requirement that all facets be identical, then an Archimedian polyhedron of 60 vertices (the 60-hedron or soccer ball) is the largest regular geometry possible (Coxeter, 1973). Above this number of atoms, the giant resonance might break up into modes indicative of the cluster acquiring a nonspherical shape, which would be an elegant way to demonstrate this upper limit experimentally if a
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
347
sufficiently good mass selection were possible. For very large clusters, there is of course the possibility of superstructures, e.g., a 60 sphere of 60 atom clusters, etc. A simple model for the collective resonances is readily constructed in terms of the Mie solution for an oscillating conducting sphere driven by an electromagnetic field (Mie, 1908). For free particles arranged in a sphere, the photoabsorption cross section has a simple damped resonance form,
where the resonance frequency w,, , also referred to as the classical Mie frequency or plasmon frequency, is the frequency of propagation of a surface plasma wave on the sphere of N free electrons, and the resonance arises from an oscillation of the spherical cloud against a fixed background of positive charges. In the case of clusters, the valence electrons oscillate against the background of the positive ions. The theory is based on using the Drude dielectric function, namely
where E is the complex dielectric constant and relies on the assumption that all of the available dipole oscillator strength is exhausted by the single resonance at w". This is known as the surface-plasmon pole approximation (Eckart, 1985). For conducting spherical clusters, the Mie theory is extended by setting w; = N e 2 / m , a , where (Y is a static polarizability (equal to the particle radius cubed in the bulk limit, to be discussed shortly) that needs to be measured or calculated in order to determine wo. (Near-spherical geometries can be taken into account in the manner of Clemenger (1985), and there are often two nearly degenerate frequencies corresponding to oscillations of a prolate or oblate spheroid.) The damping constant r is then the only free parameter, and has been found both to be small and not to vary much with cluster size. For a classical monovalent conducting sphere, (Y = Nr,3 where r, is the WignerSeitz radius. An estimate for clusters of Na yields a wavelength A, = 3650A for r, = 4 a.u. A basic assumption of the Mie (1908) model is that (a) the driving field should be spatially homogeneous over the sphere, i .e., that 2rkrs A
- -2
0.0
2 Q)
4-
Q
42
43
44
45
46
47
48
Energy (eV) FIG. 1 I . The 3p + 3d giant resonances in Cr, Mn, and Mn 11, which are isoelectronic with Cr and nevertheless possess very different forms of giant resonance because of interchannel interactions. (After Cooper er al.. 1989.)
scheme and results in a highly developed manifold of double excitations within the giant resonance. We thus uncover another aspect of cooperative effects in atoms. At the same time as they quench the independent electron spectrum (disappearance of Rydberg series within the dipole-allowed channels of the independent electron scheme), many-body effects can dramatically enhance double and multiple excitations, leading to an unexpected profusion of Rydberg series in the range where cooperative effects are expected to occur. Thus, although no actual cooperative oscillations occur, we witness a destruction of the outermost subshell structure of the atom which is due to a complete breakdown of the independent particle model.
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
36 1
Perhaps the best example of this behaviour is to be found in Ba I, where, according to the independent electron scheme, excitation from the 5p subshell should proceed as: 5p66s2'&+ 5p56s2(*P,,2,3,,)ns, nd converging on just two ionisation limits. This excitation scheme rests on the assumption that the 6s2 subshell is well behaved, i.e., that it is a passive spectator to the transition, so that the 5p5 electrons together form a single quasi particle defining just two states split by the spin-orbit interaction. However, because of the breakdown of independent electron characterisation discussed in Section IV, the quasi-particle approximation actually breaks down and, instead of just two limits, one observes a multitude of series converging on at least 14 states of the parent ion, which can be described as a thorough mixture of the configurations 5p56s2,5p56s5d, and 5p35d2. Formally, the transitions arise as two or even multielectron excitations, but the mixture in the photoabsorption spectrum of the atom is so complete that the transitions can only be approximately ordered into sequences, and the spacings only begin to resemble those of a Rydberg series in the immediate vicinity of the series limit. Situations where large spectral densities of interacting levels are observed occur frequently in nuclear physics and have been analysed in terms of the random matrix theory (Porter, 1965). Interestingly, it is this subject that led to the concepts of spectral rigidity and to the Wigner level distribution used as of early 1991 in discussions of chaos in high Rydberg states. However, the situation in both nuclear physics and the present example is different in two important respects. First, we are not under semiclassical conditions (very high values of the principal quantum number n). Second, we deal with a many-electron problem, i.e., irregular motion in the present example would occur simply because the number of dynamic variables is large. The first suggestion that the techniques of nuclear spectroscopy should be applied to the analysis of complex atomic spectra appears to have been made by Rosenzweig and Porter (1960) in an important paper that attempted to extract the form of the level distributions from existing knowledge of the optical spectra of the lanthanides. The paper of Rosenzweig and Porter (1960) was important because of the novelty of the idea and because it laid the conceptual foundation for a statistical approach to the study of atomic spectra. However, one must doubt the significance of the conclusions they were able to reach (i) because they worked with data from an energy range where independent electron models may still be valid, despite the complexity of the spectra and (ii) because the data were both incomplete and unreliable; specifically, the tabulations available to them had been com-
I I
627
Wavelength (A)
I 590
FIG. 12. A portion of the 5p spectrum of Ba, showing the high density of discrete transitions and the high spectral rigidity that results (upper plot: the accumulated density of states).
363
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
0.9
0.5
0
i
2
3
Relative spacings x
FIG.13. Statistical distribution of the level spacings for the sample of Fig. 12, showing the tendency toward a Wigner distribution of states, characteristic of a chaotic system.
piled from different experiments and contained levels of many different J values, which had to be “binned” by J value since levels of different J do not interact. In such complex spectra, the procedure was fraught with uncertainty. The 5p excitation spectrum of Ba I seems to provide a much more satisfactory example (Connerade et al., 1990). First and foremost, the data provide a manifold of high spectral density, whose spectral rigidity is high as can be seen from Fig. 12, all the levels having the same J value, since they are all reached in photoabsorption from the IS, ground state. Second, the distribution of levels does indeed signal a trend toward the Wigner law (see Fig. 13) which favours equal spacings between levels. Thus, there is clear evidence that cooperative effects in atoms can lead to a breakdown of the shell structure in many-electron atoms, and that this breakdown is associated with the emergence of irregular or chaotic motion, as hinted at in the Introduction (Section I). Furthermore, the homologous sequence Ca, Sr, and Ba has been studied, and, for the corresponding spectra of the lighter alkaline earths, although similar energy degeneracies exist, one finds that fairly regular, interacting Rydberg series can still be recognised, becoming more complex (more departures from the independent electron scheme) in Sr than in Ca. Thus, the breakdown of the external shell structure due to cooperative excitations in Ba is clearly associated with the larger number of electrons than are present in the other alkaline earths.
364
Connerade
XII. Conclusion We have presented the main ideas and the experimental background associated with the study of cooperative effects in atomic physics. The field is an active area of research. Many of our conclusions are therefore provisional, and the development of new experimental techniques, in particular intense lasers and also mass-selected beams of clusters, is expected to have a major impact on the subject over the next few years.
REFERENCES Amusia, M. Ya., Dolmatov, V. K., and Ivanov, V. K. (1974). Sov. Phys. JETP 39,752. Amusia, M. Ya., Cherepkov, N. A,, Janev, R. K., and ZivanoviC, D. J. (1976). Electronic and Aromic Collisions COBIC I, 242. Amusia, M. Ya., Ivanov, V. K., andchernysheva, L. V. (1981). J. Phys. B A t . Mol. Phys. 14, L19. Amusia, M. Ya., Dolmatov, V. K., and Ivanov, V. K. (1978). Lett. J . Tech. Phys. 4, 1305. Amusia, M. Ya., Chernysheva, L. V., Gribakin, G. F., and Tsemekhman, K. L. (1990a). J. Phys. B At. Mol. Opt. Phys. 23, 393. Amusia, M. Ya., Dolmatov, V. K., and Mansurov, M. M. (19Wb). J. Phys. B At. Mol. Phys. 23, L491. Amusia, M . Ya., Gribakin, G. F., Ivanov, V. K., and Chernysheva, L. V. (1990~).J . Phys. B At. Mol. Opt. Phys. 23,385. Amusia, M. Ya., Ivanov, V. K., Cherepkov, N. A., and Chernysheva, L. V. (1971). Sov. Phys. JETP 60,160. Armstrong, J. A,, Esherick, P., and Wynne, J. J. (1977). Phys. Rev. 15, 180. Baldwin, G. C., and Klaiber, G. S. (1947). Phys. Rev. 71,3. Baldwin, G. C., and Klaiber, G. S. (1948). Phys. Rev. 73, 1156. Band, 1. M., and Fomichev, V. I. (1980). Phys. Lett. A75, 178. Berman, B. L., and Fultz, S. C. (1975). Rev. Mod. Phys. 47,713. Berry, R. S. (1986). In “The Lesson of Quantum Theory” (J. de Boer, E. Dal, and 0. Ulfbeck, eds.), p. 241. Elsevier Science Publishers, Amsterdam. Bertsch, G. F., Bortignon, P. F., and Broglia, R. A. (1983). Rev. Mod. Phys. 55, 287. B o b , N. (1913). Phil. Mug. 26, 1,476, 857. Boyer, K . , and Rhodes, C. K. (1985). Phys. Rev. Len. 54, 1490. Boyer, K., Gibson, G., Jara, H.,Luk, T. S., Mclntyre, I. A., McPherson, A., Rosman, R., Rhodes, C. K., and Solem, J. C. (1990). In “Atoms in Strong Fields’’ (C. A. Nicolaides, C. W. Clark, and M. H. Nayfeh, eds.), p. 283. NATO AS1 Series B (Physics), Vol. 212. Plenum, New York. Brack, M. (1989). Phys. Rev. (condensed matter) 39,3533. Brandt, W., and Lundqvist, S. (1967). J. Quanr. Spectrosc. Rudiur. Transfer 7,411. Bdchignac, C. (1990). Private communication. Brkchignac, C., Cahuzac, Ph., Carlier, F., and Leygner, J. (1990). Chemical Physics Letters 164, 433. Brint, P., Connerade, J. P., Mayhew, C., and Sommer, K. (1985). J . Chem. Soc. Furuday Trans. 81, 1643.
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
365
Brout, R., and Carruthers, P. (1963). “Lectures on the Many Electron Problem,” Interscience Monographs and Texts in Physics and Astronomy. John Wiley, New York. Bruhn, R., Sonntag, B., and Wolff, M. W. (1978). Phys. Letr. 69A, 9. Burnett, K. (1989). J . Modern Optics 36,925. Chou, M. Y., Cleland, A., and Cohen, M. L. (1984). Solid State Commun. 52, 645. Clemenger, K. (1985). Phys. Rev. B (condensed matter) 32, 1359. Codling. K., and Frasinski, L. J. (1990). “Atoms in Strong Fields,” C. A. Nicolaides, C. W. Clark, and M. H. Nayfeh, eds. p. 513, NATO AS1 Series B, Vol. 212, Plenum, New York. Connerade, J. P. (1978a). Contemp. Phys. 19,415. Connerade, J. P. (1978b). J . Phys. B At. Mol. Phys. 11, L381. Connerade, J. P. (1978~).J. Phys. B At. Mol. Phys. 11, L409. Connerade, J. P. (1982). J . Phys. B A f . Mol. Phys. 15, L886. Connerade, J. P. (1983). J . Phys. BAr. Mol. Phys. 16, L257. Connerade, J. P. (1984). J. Phys. BAr. Mol. Phys. 17, L165. Connerade, J. P. (1988). In “Fundamental Processes of Atomic Dynamics” (J. S. Briggs, er al., eds.), p. 571. NATO AS1 Series B (Physics), Vol. 181. Plenum, New York. Connerade, J. P., and Dietz, K. (1987). Comments on Afomic and Molecular Physics XIX, 283. Connerade, J. P., and Karnatak, R. C. (1989). Comments on Aromic and Molecular Physics XXIV, I . Connerade, J. P., and Lane, A. M. (1987). J. Phys. B Ar. Mol. Phys. 20, L181. Connerade, J. P., and Lane, A. M. (1989). Reporfs on Progress in Physics 51, 1439. Connerade, J. P., and Mansfield, M. W. D. (1975). Proc. Roy. Soc. (London) A 348,235. Connerade, J. P., Mansfield, M. W. D., and Martin, M. A. P. (1976). Proc. Roy. Sac. (London) A 350,405. Connerade, J. P., Pantelouris, M., Baig, M. A., Martin, M. A. P.,and Cukier, M. (1980). J. Phys. B 13, L357. Connerade, J . P., Esteva, J. M., and Karnatak, R. C. (eds.) (1987). “Giant Resonances in Atoms, Molecules and Solids.” NATO AS1 Series B (Physics), Vol. 151. Plenum, New York. Connerade, J. P., Ma, H., Shen, N., and Stavrakas, T. A. (1988). J. Phys. B. At. Mol. Phys. 21, L24 I . Connerade, J. P., Baig, M. A., and Sweeney, M.(1990). J . Phys. B 23,713. Cooper, J. W., Clark, C. W., Cromer, C. L., Lucatorto, T. B., Sonntag, B. F., Kennedy, E. T., and Costello, J. T. (1989). Phys. Rev. A39, 6074. Coxeter, H. S. M. (1961). In “Introduction to Geometry.” John Wiley, New York. Coxeter, H. S . M. (1973). In “Regular Polytopes.” Reprinted. Dover, New York. Davis, L. C., and Feldcamp, L. A. (1978). Phys. Rev. A 17,2012. de Heer, W. A,, Knight, W. D., Chou, M. Y., and Cohen, M. L. (1987). Solid Sfate Physics 40,93. Dehmer, J. L. (1972). J. Chem. Phys. 56,4496. Ekardt (1985). Phys. Rev B 31,6360. Fano, U. (1961). Phys. Rev. 124, 1866. Fano, U. (1975). In “Photoionisation and Other Probes of Many-Electron Interactions” (F. J. Wuilleumier, ed.), p. 11. NATO AS1 (Series B) Plenum, Fano, U. (1976). ”Photoionisation and Other Probes of Many Electron Interactions,” F. Wuilleumier, ed., Plenum Press, New York. Fermi, E. (1928). In “Quantentheorie und Chemie.” (H. Falkenhagen, ed.). Hinzel-Verlag,Leipzig. Gailitis, M. (1963). Sov. Phys. JETP 17, 1328. Garvin, L. J., Brown, E. R., Carter, S. L., and Kelly, H. P. (1973). J. Phys. B Ar. Mol. Phys. 16, L269. Goeppert-Mayer, M. (1941). Phys. Rev. 60, 184.
366
Connerade
Griffin, D. C., Andrew, K. L., and Cowan, R. D. ( 1969). Phvs. Rev. 177,62. Griffin, D. C., Cowan, R. D., and Andrew, K. L. (1971). Phvs. Rev. A3, 1233. Griffin, D. C., Cowan, R. D., and Pindzola, M. S. (1987). In “Giant Resonances in Atoms, Molecules and Solids” J. P. Connerade er a l . . eds. Plenum, New York. Hecht, J. (1979). J . Appl. Phys. 59, 7186. Herrick, D. R . (1978). Phys. Rev. 17, 1. Herrick, D. R., and Sinanoglu, 0. (1975). P h y . Rev. A l l , 97. Herzberg, G. (1935). “Atomic Spectra and Atomic Structure,’’ p. 148. Dover, New York. Kelly, H. P. (1964). Phys. Rev. 144, 39. Kelly, H. P. (1966). P h y . Rev. 136,293. Kelly, H. P., Carter, S. L., and Norum, B. E. (1982). Phvs. Rev. 25, 2052. Knight, W. D., Clemenger, K., de Heer. W. A,, Saunden, W. A., Chou, M. Y., and Cohen, M. L. (1984). Phvs. Rev. Lett. 52, 2141. Kramers, H. A. (1923). Zeitfiir Phys. 13, 312. Krause, M. 0..Carlson, T. A., and Fahlman, A. (1984). Phvs. Rev. A30, 1316. Kreibig, U . , and Genzel, L. (1983). Sur$ Sci. 156, 678. Lambropoulos, P. (1985). Phys. Rev. Lett. 55, 2141. Landau, L. D., and Lifshitz, E. M. (1977). “Quantum Mechanics,” 3rd ed. (translation), pp, 2, 3, 21. Pergammon London, New York, Paris. Langmuir, I. (1921). Phvs. Rev. 17, 339. Lefebvre-Brion, H. (1987). f n “Giant Resonances in Atoms, Molecules and Solids” (J. P. Connerade, J. M. Estera, and R. C. Karnatak)., p. 301. Plenum, New York. L’Huillier, A,, and Wendin, G. (1988). J . Phys. B At. Mol. Opt. Phys. 21, L247. Liberman, D. (1979). Phys. Rev. B20, 4981. Lin, C. D. (1974). Phys. Rev. A10, 1986. Lucatorto, T. I., McIlrath, T., Sugar, J.. and Younger, S. M. (1981). Phys. Rev. Lett. 47, 1124. Luk, T. S., Johann, U . , Egger, H., F’ummer, H., and Rhodes, C. K. (1985). Phys. Rev. A32, 214. Maiste, A. A., Ruus, R. E., Ruchas, S. A,, Karaziya, R. I., and Elango, M. A. (1980). Sov. Phvs. a JETP 51,474. Sov. Phys. JETP52,844b. Mansfield, M. W. D. (1977). Proc. Roy. Sac. Lond. A358,253. Martins, J. L., Buttet. J., and Car, R., (1985). Phys. Rev. B (condensed matter) 31, 1804. Mie, G. (1908). Ann. Phys. (Leipzig) 25, 377. More, R. M. (1982). J. Q u m t . Spec. and Radiat. Transfer. 21, 345. Mulliken, R. S. (1964). J. Am. Chem. SOC.86,3183. Nikitin, S. I., and Ostrovsky, V. N. (1976). J. Phys. B At. Mol. Phys. 9, 3141. Nikitin, S. I . , and Ostrovsky, V. N. (1978). J. Phys. B At. Mol. Phys. 11, 1681. Nilsson, S. G. (1955). Mat. Phys. Medd. K . Dan. Vidensk. Selsk. 29, no. 16, I. Nuroh, K., Scott, M. J., and Zaremba, E. (1982). Phys. Rev. Lett. 49, 862. Porter, C. E. (1965). “Statistical Theories of Spectra: Fluctuations.” Academic Press, New York. Radtke, E. (1979). J. Phys. B A t . Mol. Phys. 12, L71. Robin, M. B. (1985). Chem. Phys. Lett. 19, 33. Also in “Higher Excited States of Polyatomic Molecules,” Vol. 3; Academic Press, New York (1986). Rosenzweig, N., and Porter, C. E. (1960). Phys. Rev. 120, 1698. Seaton, M. J. (1966a). Proc. Phys. SOC. 88, 801. Seaton, M. J. (1966b). Proc. Phys. SOC. 88,815. Selby, K., Vollmer, M., Masui, 1.. Kresin, V., de Heer, W. A,, and Knight, W. D. (1989). Phvs. Rev. B (condensed matter) 40, 5417. Sinanoglu, 0. (1960). J . Chem. Phys. 33, 1212. Sinanoglu, 0. (1961). Proc. Roy. SOC.(London) A260, 379. Starace, A. F. (1976). In “Photoionisation and Other Probes of Many-Electron Interactions” (F. J. Wuilleumier, ed.), p. 395. NATO AS1 Series B (Physics), Vol. 18. Plenum, New York.
COOPERATIVE EFFECTS IN ATOMIC PHYSICS
367
Stoner, E. C. (1924). Phil. Mug. 48,719. Sugiura, M. Y. (1927). Journal de Physique ez le Radium Serie V1, Tome VIII, 8. Szoke, A.. and Rhodes, C. K . (1986). Phys. Rev. Lett. 56,720. Wannier, G. H. (1953). Phys. Rev. 90, 817. Woods, R. D., and Saxon, D. S. (1954). Phys. Rev. 95,517. Zangwill, A. (1987). In “Giant Resonances in Atoms, Molecules and Solids” (J. P. Connerade, J. M. Esteva, and R. C. Karnatak, eds.), p. 321. Plenum, New York. Zimmermann, P. ( 1989). Commenrs on Atomic undMofeculur Phvsics XXIII, 45.
This Page Intentionally Left Blank
Index
A Above-threshold Ionization, 34, 38, 41, 95,101 Adiabaticity parameter, 35, 38, 39,43 Antimatter, 283, 284 Antiscreening, 244, 299 a-Particle. See Helium. Apparent electron excitation cross section, See Electron excitation. Atomic physics, 325 Aufbau principle, 329, 334 Autoionisation. 333
B Bethe rule, 97, 100 Binomial probabilities, 252, 264, 270 Branching ratio, 4
C Cascades, 4. 9, 26,29 Classical limit, 246, 250 Closed shells, 329 Clusters, 344 Collective processes, 341, 349, 356 Collisional redistribution, 114- 116 Collision-induced coherences, I 14- 174 cross-relaxation, 149- 152 dipole-dipole interaction, 152- 155 dressed states, 165
369
four-wave mixing, 129- 139 Hanle effect, 124- 127 open two-level system, 127 in probe absorption, 139- 144 relationships in second- and third-order responses, 145 resonances in fifth-order nonlinearities, 147149 responses in spontaneous processes, 155160 signal changes, 160- 163 Collision-induced resonances. See CollisionInduced Coherences Computational methods, 250 Controlled collapse, 34 1 Cooperative effects, 325 Correlated emission, 172 Correlation, 314 dynamic, 222 general, 218, 221, 226, 230, 251,266, 312, 314 scattering, 222,265 static, 222, 266 two-center, 244,299, 312 Correlations, 328 Coulomb interference, 184 Coulomb potentials, 223 Cross relaxation, 149- I52
370
Index D
Differential electron excitation cross section. See Electron excitation. Dipole-dipole interaction, 116, 152, 155 Direct electron excitation cross sections. See Electron excitation. Double excitations, 327 Dressed states, 116, 158- 160, 164- 166 Dynamic stark effect, 35,43
E Eight-wave mixing, 114, 148. See also Wave mixing. Einstein A Coefficient-A, 4 Electron Beam Current-I, 3 , 4 Electron capture double, 289 with excitation, 297 with ionization, 292 multiple, 288 single, 239 special features, 239 Electron excitation, 5, 8, 10 apparent cross section-Q,, , 4 , 9, 28 differential cross section, 2, 12, 28 direct cross sections, 4, 6, 28, and rare-gas atoms, 12-27 Energy balance in MuCF, 178, 208 Energy conservation and non-conservation, 234,240,294 Exclusive probabilities, 224, 254 Excitation with capture, 297 double, 287 with ionization, 291 multiple, 286 single, 265
F Faraday cup, 3,25 Fifth-order nonlinearities, 147 Fluctuation-induced resonances, 169- 171 Fluctuations, 114, 156, 167, 169, 170-172 Fluorescence and ionization spectroscopy, 1 16, 124, 172 Fokker-Planck method, 259 Four-wave mixing, 129-139, 152, 161- 165, 168, 172. See also Wave mixing. G Giant resonances, 337 g-Hartree, 352-354
H Hanle effect, 124 Hanle resonance, 121, 125, 133, 135, 169 Hartree-Fock, 326 Helium, 180- 183 muonic, 184, 187 nuclear masses, I80 High electromagnetic fields, 355 Hydrogen, 178-211 helium impurity, 207 nuclear masses, 180 plasma, 197, 204 stopping power, 197, 202 target for MuCF, 178 I Inclusive probabilities, 224, 254 Independent electron approximation, 226, 229, 252,270,274,288, 293, 312 Independent particle approximation. 326 Inelastic collisions, 117, 121, 131, 138, 157, 159 Innershell, 327, 336 Intershell and intersubshell coupling, 358 Ionization with capture, 292 double, 268,273 with excitation, 291 multiple, 270 probability, 36,45,47,52,61-63, 87.94 single, 250 lonization process, nonlinearity power of, 47, 55.60, 103
K Kinematics, 240, 295 L Lanthanide contraction, 334 Laser-induced fluorescence (LIF), 2, 28, 118 ofHe, 13, 15 of Ne, 16-19 Line mixing, 150, 152 LS coupling, 8, 1 I
M Manybody effects, 325 Many-body perturbation theory (MBPT), 237, 281, 338. See also Perturbation theory. Many-body problem, 217,312,266 Many-electron problem formulation, 219, 232
37 1
INDEX general, 217, 312, 266 two-center, 244, 299, 312 Mayer-Fermi Model, 335 Mean Fields, 352 Mie Sphere, 347 Molecular Scattering, 247, 302 MuCF. See Muon-catalyzedfusion. Multicharge Ions, 34, 39, 101 Multielectron excitations, 361 Multifrequency laser radiation, 51, 60, 65 statistical factor of, 52, 59, 89, 96 Multiphoton cross section, 34- 1 I 1 polarization dependence of, 97, 100 Multiphoton excitations, 357 Multiphoton ionization, 34- 11 I direct, 34-36, 44 measuring methods, 45-72 measuring procedures, 73-89 measuring results, 90- 101 resonance, 45 probability of, 36 Muomolecular, 179, 189- 196, 200 Auger transitions, 187, 191- 192 Born-Oppenheimer approximation, 200 electro-muomolecularcomplexes, 190- 192 formation rate, 192, 195 geometry (radii), 194 nuclear effects, 185, 193 QED corrections, 193- 194 resonant formation, 192, 195 spectra, 189-191 Muon, 178-181, 208 Muon-catalyzed fusion (MUCF), 177-21 1 catalytic cycle, 186- 198 cycling rate. 179-181, 186-188 density dependence, 196, 203, 206-209 energy balance, 178, 208 fusion, 179- 187, 196- 199 muon loss, 180-181, 187 muon sticking, 186, 199-209 Q-value. 181-184 regeneration, 201 -204 temperature dependence, 195, 198, 209 transfer dCL---)tp,187 x-rays after fusion, 204-206
N Nonlinear susceptibilities, 116-1 18, 148 Non-Rydberg spectroscopy, 333 Nuclear dr fusion, 183- 186 branching ratio. 199 parameters, 183, 196
reaction products, 182- 184 R-matrix approach, 196 resonance detuning, 185 Nuclei, 345
0 One-electron approximation, 95 Open two-level systems, 127, 139 Optical emission cross section-Q., , 3-5 Optogalvanic spectroscopy, 128
P Perturbation theory, 232, 244, 250. See also Many-body perturbation theory. Phase conjugation geometry, 161, 171 Photoionization, 268, 278, 281 Plasmons, 348 Pressure-induced coherence, 121 Pressure-induced resonance, 166 Probabilities, 221, 224,252-254.264 Projectile electrons, 244, 299
Q Quantum jumps, I60 Quasiparticle, 335
R Raman scattering, 122, 157 Random phase approximation, 339 Ratios of cross sections, 268, 277, 280, 285, 288,291, 296, 301 Runge-Lens vector, 330 S Saturated absorption, 114- 118, 145 Saturation-inducedextra resonances, 172 Scaling laws, 264 Shake, 265-268, 279-280, 288, 296 Aberg’s hodel, 279 simple, 276 Single frequency laser radiation, 46, 53, 65 Six-wave mixing, 114- 116, 147, 152. See atso Wave mixing. Spatial coherence effects, 173 Stabilisation, 333, 357 Stark shift, 101, 108 Statistical energy deposition, 255, 272 Subharmonics, 148
T Thomas singularities, 240, 294 Time-dependent Hartree-Fock (TDHF). 25 1
372
Index
Time ordering, 234 Transfer. See Electron capture. Transient response, 167 Transition probability, 5 Trapped states, 157 Two-photon absorption, 143, 152 Two-photon maser transition, 172 Two-photon resonance, 152- 154 Two-sided diagrams, 122, I32 Two-time correlation, 116, 157
W Wall collisions, 174 Wave mixing eight-wave, 114, 148 four-wave, 129-139, 152, 161-165, 168, 172 six-wave, 114- 116, 147, 152. Wigner distribution, 363 Woods-Saxon potential, 348
U
X X-Ray spectroscopy, 337
V
Zi terms, 236,281
Ultraviolet, 337
Y-2 Vacuum field Rabi splittings, 174 Velocity-changing collisions, 116, 160, 163
Contents of Previous Volumes
Volume 1 Molecular Orbital Theory of the Spin Properties of Conjugated Molecules, G . G. Hall and A. T. Amos Electron Affinities of Atoms and Molecules, B . L. Moiseiwirsch
The Measurement of the Photoionization Cross Sections of the Atomic Gases, James A. R. Samson The Theory of Electron-Atom Collisions, R. Peterkop and V. Veldre Experimental Studies of Excitation in Collisions between Atomic and Ionic Systems, F. J. de Heer
Atomic Rearrangement Collisions, B. H . Bransden Mass Spectrometry of Free Radicals, S. N. The Production of Rotational and Vibrational Foner Transitions in Encounters between Molecules, K. Takayanagi Volume 3 The Study of Intermolecular Potentials with Molecular Beams at Thermal Energies, H.PAU~Y The Quanta1 Calculation of Photoionization Cross Sections, A. L. Stewart and J . P. Toennies High-Intensity and High-Energy Molecular Radiofrequency Spectroscopy of Stored Ions I: Storage, H.C. Dehmelt Beams, J. B. Anderson, R. P. Andres, and J . B. Fenn Optical Pumping Methods in Atomic Spectroscopy, B. Budick
Volume 2
Energy Transfer in Organic Molecular Crystals: A Survey of Experiments, H. C . Wolf
The Calculation of van der Waals Interactions, A. Dalgarno and W. D. Davison
Atomic and Molecular Scattering from Solid Surfaces, Robert E. Stickney
Thermal Diffusion in Gases, E. A. Mason, R. J . Munn, and Francis J . Smith Spectroscopy in the Vacuum Ultraviolet, W. R. S.Carton
Quantum Mechanics in Gas Crystal-Surface van der Waals Scattering, E. Chanoch Beder Reactive Collisions between Gas and Surface Atoms, Henry Wise and Bernard J . Wood 373
374
Contents of Previous Volumes
Volume 4
The Calculation of Atomic Transition Probabilities, R. J. S. Crosslev
H. S. W. Massey-A Sixtieth Birthday Tribute, E. H. S. Burhop Electronic Eigenenergies of the Hydrogen Molecular Ion, D. R. Bates and R. H. G. Reid Applications of Quantum Theory to the Viscosity of Dilute Gases, R. A. Buckingham and E. Gal Positrons and Positronium in Gases, P. A . Fraser Classical Theory of Atomic Scattering, A. Burgess and I . C. Percival Born Expansions, A . R. Holt and B . L. Moiseiwirsch Resonances in Electron Scattering by Atoms and Molecules, P. G.Burke Relativistic Inner Shell Ionizations, C. B. 0. Mohr Recent Measurements on Charge Transfer, J . B . Hasted Measurements of Electron Excitation Functions, D. W. 0. Heddle and R. C . W. Keesing Some New Experimental Methods in Collision Physics, R. F. Srebbings Atomic Collision Processes in Gaseous Nebulae, M. J . Seaton Collisions in the Ionosphere, A. Dalgarno The Direct Study of Ionization in Space, R. L. F. Boyd
Tables of One- and Two-Particle Coefficients of Fractional Parentage for Configurations skss‘ppq,C. D . H . Chisholm, A . Dalgarno. and F. R. Innes
Volume 5 Flowing Afterglow Measurements of IonNeutral Reactions, E. E. Ferguson. F. C. Fehsenfeld, and A. L. Schmeltekopf Experiments with Merging Beams, Roy H. Ne,ynaber Radiofrequency Spectroscopy of Stored Ions 11: Spectroscopy, H. C. Dehmelr
Relativistic 2-Dependent Corrections to Atomic Energy Levels, Holly Thomis Doyle
Volume 6 Dissociative Recombination, J . N. Bardsley and M. A. Biondi Analysis of the Velocity Field in Plasmas from the Doppler Broadening of Spectral Emission Lines, A . S. Kaufman The Rotational Excitation of Molecules by Slow Electrons, Kazuo Takayanagi and Yukikaru Itikawa The Diffusion of Atoms and Molecules, E. A . Mason and T. R. Marrero Theory and Application of Sturmian Functions, Manuel Rotenberg Use of Classical Mechanics in the Treatment of Collisions between Massive Systems, D. R. Bates and A. E. Kingston
Volume 7 Physics of the Hydrogen Master, C. Audoin. J . P. Schermann, and P. Criver Molecular Wave Functions: Calculation and Use in Atomic and Molecular Processes, J . C. Browne Localized Molecular Orbitals, Hare/ Weinstein, Ruben Pauncz. and Maurice Cohen General Theory of Spin-Coupled Wave Functions for Atoms and Molecules, J. Gerratt Diabatic States o f Molecules-Quasi-Stationary Electronic States, Thomas F. 0’Malle.y
The Spectra of Molecular Solids, 0. Schnepp
Selection Rules within Atomic Shells, B . R. Judd
The Meaning of Collision Broadening of Spectral Lines: The Classical Oscillator Analog, A. Ben-Reuven
Green’s Function Technique in Atomic and Molecular Physics, Gy. Csanak. H. S. Tavlor. and Robert Yaris
CONTENTS OF PREVIOUS VOLUMES A Review of Pseudo-Potentials with Emphasis on Their Application to Liquid Metals, Nathan Wiser and A. J . Greenfield
315
A Review of Jovian Ionospheric Chemistry, Wesley T. Huntress, Jr.
Volume 11 Volume 8 Interstellar Molecules: Their Formation and Destruction, D. McNally Monte Carlo Trajectory Calculations of Atomic and Molecular Excitation in Thermal Systems, James C. Keck Nonrelativistic Off-Shell Two-Body Coulomb Amplitudes, Joseph C. Y. Chen and Augustine C. Chen Photoionization with Molecular Beams, R. B. Cairns, Hatstead Harrison. and R. I . Schoen
The Theory of Collisions between Charged Particles and Highly Excited Atoms, I . C. f e r cival and D. Richards Electron Impact Excitation of Positive Ions, M. J . Seaton The R-Matrix Theory of Atomic Process, P. G. Burke and W. D. Robb Role of Energy in Reactive Molecular Scattering: An Information-Theoretic Approach, R. B. Bernstein and R. D. Levine Inner Shell Ionization by Incident Nuclei, Johannes M. Hansteen
The Auger Effect, E. H. S. Burhop and W. N. Asaad
Stark Broadening, Hans R. Griem
Volume 9
Chemiluminescence in Gases, M. F. Golde and B . A. Thrush
Correlation in Excited States of Atoms, A . W. Weiss
Volume 12
The Calculation of Electron-Atom Excitation Cross Sections, M . R. H. Rudge
Nonadiabatic Transitions between Ionic and Covalent States, R. K. Janev
Collision-Induced Transitions between Rotational Levels, Takeshi Oka
Recent Progress in the Theory of Atomic Isotope Shift, J . Bauche and R . J . Champeau
The Differential Cross Section of Low-Energy Electron-Atom Collisions, D. Andrick
Topics on Multiphoton Processes in Atoms, f . Lambropoulos
Molecular Beam Electric Resonance Spectroscopy, Jens C. Zorn and Thomas C. English
Optical Pumping of Molecules, M . Brover, G. Goudedard, J . C. Lehmann. and J . Vigui
Atomic and Molecular Processes in the Martian Atmosphere, Michael B. McElrov
Highly Ionized Ions, Ivan A. Sellin
Volume 10
Ion Chemistry in the D Region, George C. Reid
Relativistic Effects in the Many-Electron Atom, Lloyd Armstrong. Jr. and Serge Feneuille
Volume 13
The First Born Approximation, K. L. Bell and A. E. Kingston
Time-of-Flight Scattering Spectroscopy, Wilhelm Raith
Photoelectron Spectroscopy, W. C. Price
Atomic and Molecular Polarizabilities-A Review of Recent Advances, Thomas M. Miller and Benjamin Bederson
Dye Lasers in Atomic Spectroscopy, W. Lunge, J . Luther. and A . Steudel
Study of Collisions by Laser Spectroscopy, Paul R. Berman
Recent Progress in the Classification of the Spectra of Highly Ionized Atoms, B. C. Fawcett
Collision Experiments with Laser-Excited Atoms in Crossed Beams, I . K Hertel and W. Stoll
376
Contents of Previous Volumes
Scattering Studies of Rotational and Vibrational Excitation of Molecules, Manfred Faubel and J . Peter Toennies
Ion-Atom Charge Transfer Collisions at Low Energies, J. B. Hasted Aspects of Recombination, D. R. Bates
Low-Energy Electron Scattering by Complex Atoms: Theory and Calculations, R. K. Nesbet
The Theory of Fast Heavy Particle Collisions, B. H. Bransden
Microwave Transitions of Interstellar Atoms and Molecules, W.B. Somerville
Volume 14 Resonances in Electron Atom and Molecule Scattering, D. E. Golden The Accurate Calculation of Atomic Properties by Numerical Methods, Brian C. Webster. Michael J . Jamieson, and Ronald F. Srewart (e, 2e) Collisions, Erich Weigold and Ian E . McCarthy Forbidden Transitions in One- and Two-Electron Atoms, Richard Marrus and Peter J . Mohr Semiclassical Effects in Heavy-Particle Collisions, M. S. Child Atomic Physics Tests of the Basic Concepts in Quantum Mechanics, Francis M. Pipkin Quasi-Molecular Interference Effects in IonAtom Collisions, S. V. Bobashev Rydberg Atoms, S. A. Edelstein and T. F. Gallagher UV and X-Ray Spectroscopy in Astrophysics, A. K. Dupree
Volume 15 Negative Ions, H. S. W. Massey Atomic Physics from Atmospheric and Astrophysical Studies, A. Dalgarno Collisions of Highly Excited Atoms, R. F. Stebbings Theoretical Aspects of Positron Collisions in Gases, J . W. Humberston Experimental Aspects of Positron Collisions in Gases, T. C. Grifith Reactive Scattering: Recent Advances in Theory and Experiment, Richard B. Bernsrein
Atomic Collision Processes in Controlled Thermonuclear Fusion Research, H . B. Gilbody Inner-Shell Ionization, E. H. S. Burhop Excitation of Atoms by Electron Impact, D. W. 0.Heddle Coherence and Correlation in Atomic Collisions, H. Kleinpoppen Theory of Low Energy Electron-Molecule Collisions, P. G. Burke
Volume 16 Atomic Hartree-Fock Theory, M. Cohen and R. P. McEachran Experiments and Model Calculations to Determine Interatomic Potentials, R. Diiren Sources of Polarized Electrons, R. J . Celorta andD. T.Pierce Theory of Atomic Processes in Strong Resonant Electromagnetic Fields, S. Swain Spectroscopy of Laser-Praduced Plasmas, M. H. Key and R. J . Hutcheon Relativistic Effects in Atomic Collisions Theory, B. L.Moiseiwitsch Parity Nonconservation in Atoms: Status of Theory and Experiment, E. N. Forrson andL. Wilets
Volume 17 Collective Effects in Photoionization of Atoms, M. Ya. Amusia Nonadiabatic Charge Transfer, D. S. F. Crothers Atomic Rydberg States, Serge Feneuille and Pierre Jacquinot Superfluorescence, M. F. H. Schuurmans, Q.H . F. Vrehen, D. Polder, and H . M. Gibbs Applications of Resonance Ionization Spectroscopy in Atomic and Molecular Physics, M. G.
CONTENTS OF PREVIOUS VOLUMES Payne. C. H . Chen, G. S. Hurst, and G. W. Foltz
Inner-Shell Vacancy Production in Ion-Atom Collisions, C . D. Lin and Patrick Richard Atomic Processes in the Sun, P. L. Dufton and A. E. Kingston
Volume 18
377
The Reduced Potential Curve Method for Diatomic Molecules and Its Applications, F. JenE The Vibrational Excitation of Molecules by Electron Impact, D. G. Thompson Vibrational and Rotational Excitation in Molecular Collisions, Manfred Faubel Spin Polarization of Atomic and Molecular Photoelectrons, N. A. Cherepkov
Theory of Electron-Atom Scattering in a Radiation Field, Leonard Rosenberg Positron-Gas Scattering Experiments, Talbert S. Stein and Walter E. Kauppila Nonresonant Multiphoton Ionization of Atoms, J . Morellec. D. Normand. and G. Petite Classical and Semiclassical Methods in Inelastic Heavy-Particle Collisions, A. S. Dickinson and D. Richards Recent Computational Developments in the Use of Complex Scaling in Resonance Phenomena, B. R. Junker Direct Excitation in Atomic Collisions: Studies of Quasi-One-Electron Systems, N. Anderson and S. E. Nielsen
Volume 20 Ion-Ion Recombination in an Ambient Gas, D. R. Bates Atomic Charges within Molecules, G. G. Hall Experimental Studies on Cluster Ions, T. D. Mark and A. W. Castleman. Jr. Nuclear Reaction Effects on Atomic Inner-Shell Ionization, W.E. Meyerhof and J.-F. Chemin Numerical Calculations on Electron-Impact Ionization, Christopher Eottcher Electron and Ion Mobilities, Gordon R. Freeman and David A. Armstrong
Model Potentials in Atomic Structure, A. Hibbert
On the Problem of Extreme UV and X-Ray Lasers, I. I. Sobel'man and A. V. Vinogradov
Recent Developments in the Theory of Electron Scattering by Highly Polar Molecules, D. W. Norcross and L. A. Collins
Radiative Properties of Rydberg States in Resonant Cavities, S. Haroche and J . M . Raimond
Quantum Electrodynamic Effects in FewElectron Atomic Systems, G. W.F. Drake
Rydberg Atoms: High-Resolution Spectroscopy and Radiation Interaction-Rydberg Molecules, J . A. C. Gallas, G. Leuchs, H . Walther, and H. Figger
Volume 19 Volume 21 Electron Capture in Collisions of Hydrogen Atoms with Fully Stripped Ions, B. H . Bransden andR. K. Janev Interactions of Simple Ion-Atom Systems, J . T. Park
Subnatural Linewidths in Atomic Spectroscopy, Dennis P. O'Erien, Pierre Meystre, and Herbert Walther Molecular Applications of Quantum Defect Theory, Chris H . Greene and Ch. Jungen
High-Resolution Spectroscopy of Stored Ions, D. J. Wineland, Wayne M. llano, and R. S. Van Dyck, Jr.
Theory of Dielectronic Recombination, Yukap Hahn
Spin-Dependent Phenomena in Inelastic Electron-Atom Collisions, K. Blum and H. Kleinpoppen
Recent Developments in Semiclassical Floquet Theories for Intense-Field Multiphoton Rocesses, Shih-I Chu
378
Contents of Previous Volumes
Scattering in Strong Magnetic Fields, M. R. C. McDowell and M. Zarcone Pressure Ionization, Resonances, and the Continuity of Bound and Free States, R. M , More
Volume 24 The Selected Ion Flow Tube (SIFT): Studies of Ion-Neutral Reactions, D. Smith and N. G. Adams Near-Threshold Electron-Molecule Scattering, Michael A . Morrison
Volume 22 Positronium-Its Formation and Interaction with Simple Systems, J. W. Humbersron Experimental Aspects of Positron and Positronium Physics, T. C. Grijith Doubly Excited States, Including New Classification Schemes, C. D.Lin Measurements of Charge Transfer and Ionization in Collisions Involving Hydrogen Atoms, H. B. Gilbody
Angular Correlation in Multiphoton Ionization of Atoms, S. J . Smith and G. Leuchs Optical Pumping and Spin Exchange in Gas Cells, R. J. Knize, 2. Wu. and W. Happer Correlations in Electron-Atom Scattering, A. Crowe
Volume 25
Electron-Ion and Ion-Ion Collisions with Intersecting Beams, K. Dolder and B. Peart
Alexander Dalgarno: Life and Personality, David R. Bates and George A . Victor
Electron Capture by Simple Ions, Edward Pollack and Yukap Hahn
Alexander Dalgarno: Contributions to Atomic and Molecular Physics, Neal Lane
Relativistic Heavy-Ion- Atom Collisions, R. Anholr and Harvey Gould
Alexander Dalgarno: Contributions to Aeronomy, Michael B. McElroy
Continued-Fraction Methods in Atomic Physics, S. Swain
Alexander Dalgarno: Contributions to Astrophysics, David A. Williams Dipole Polarizability Measurements, Thomas M. Miller and Benjamin Bederson
Volume 23
Flow Tube Studies of Ion-Molecule Reactions, Eldon Ferguson
Vacuum Ultraviolet Laser Spectroscopy of Small Molecules, C. R. Vidal
Differential Scattering in He-He and He+-He Collisions at KeV Energies, R. F. Srebbings
Foundations of the Relativistic Theory of Atomic and Molecular Structure, Ian P. Grant and Harry M. Quiney
Atomic Excitation in Dense Plasmas, Jon C. Weisheit Pressure Broadening and Laser-Induced Spectral Line Shapes, Kenneth M. Sando and Shih-I Chu
Point-Charge Models for Molecules Derived from Least-Squares Fitting of the Electric Potential, D. E. Williams and Ji-Min Yan Transition Arrays in the Spectra of Ionized Atoms, J . Bauche, C. Bauche-Arnoult. and M . Klapisch Photoionization and Collisional Ionization of Excited Atoms Using Synchrotron and Laser Radiation, F. J . Wuilleumier. D. L. Ederer, and J . L. Picqui
Model-Potential Methods, G. Laughlin and G. A . Victor Z-Expansion Methods, M. Cohen Schwinger Variational Methods, Deborah Kay Watson Fine-Structure Transitions in Proton-Ion Collisions, R. H . G. Reid
CONTENTS OF PREVIOUS VOLUMES Electron Impact Excitation, R. J . W. Henry and A. E . Kingston Recent Advances in the Numerical Calculation of Ionization Amplitudes, Christopher Bottcher The Numerical Solution of the Equations of Molecular Scattering, A. C. Allison High Energy Charge Transfer, B. H. Bransden and D. P. Dewangan Relativistic Random-Phase W. R. Johnson
Approximation,
Relativistic Sturmian and Finite Basis Set Methods in Atomic Physics, G. W. F. Drake and S.P. Goldman Dissociation Dynamics of Polyatomic Molecules, T. Uzer Photodissociation Processes in Diatomic Molecules of Astrophysical Interest, Kate P. Kirby and Ewine F. van Dishoeck The Abundances and Excitation of Interstellar Molecules, John H.Black
379
Volume 27 Negative Ions: Structure and Spectra, David R. Bates Electron Polarization Phenomena in ElectronAtom Collisions, Joachim Kessler Electron-Atom Scattering, I . E. McCarthv and E. Weigold Electron-Atom Ionization, I. E. McCarthy and E. Weigold Role of Autoionizing States in Multiphoton Ionization of Complex Atoms, V. I. Lengyel and M . I . Havsak Multiphoton Ionization of Atomic Hydrogen Using Perturbation Theory, E. Karule
Volume 28 The Theory of Fast Ion-Atom Collisions, J . S. Briggs and J . H. Macek Some Recent Developments in the Fundamental Theory of Light, Peter W. Milonni and Surendra Singh
Volume 26
Squeezed States of the Radiation Field, Khalid Zaheer and M. Suhail Zubairy
Comparisons of Positrons and Electron Scattering by Gases, Walter E. Kauppila and Talbert S. Stein
Cavity Quantum Electrodynamics, E. A. Hinds
Electron Capture at Relativistic Energies, B. L. Moiseiwirsch The Low-Energy, Heavy Particle Collisions-A Close-Coupling Treatment, Mineo Kirnura and Neal F. Lane Vibronic Phenomena in Collisions of Atomic and Molecular Species, V. Sidis Associative Ionization: Experiments, Potentials, and Dynamics, John Weiner, Franqoise Masnou-Sweeuws, and Annick Giusri-Suzor
Volume 29 Studies of Electron Excitation of Rare-Gas Atoms into and out of Metastable Levels Using Optical and Laser Techniques, Chun C. Lin and L. W. Anderson Cross Sections for Direct Multiphoton Ionization of Atoms, M. V. Ammosov. N. B. Delone. M. Yu. Ivanov, I . I. Bondar, and A. V. Masalov Collision-Induced Coherences in Optical Physics, G. S. Aganval
On the /3 Decay of I8’Re: An lnterface of Atomic and Nuclear Physics and Cosmochronology, Zonghau Chen, Leonard Rosenberg, and Larry Spruch
Muon-Catalyzed Fusion, Johann Rufelski and Helga E. Rafelski
Progress in Low Pressure Mercury-Rare Gas Discharge Research, J. Maya and R. Lagushenko
Multiple Electron Excitation, Ionization, and Transfer in High-Velocity Atomic and Molecular Collisions, J. H. McGuire
Cooperative Effects in Atomic Physics, J . P. Connerade
This Page Intentionally Left Blank