You are on page 1of 7

18 Ion Beam Spectroscopy

Eric H. Pinnington
Department of Physics, University of Alberta, Edmonton, Alberta T6G 2J1, Canada

18.1 18.2

18.3

18.4 18.5

INTRODUCTION BEAM-FOIL SPECTROSCOPY 18.2.1 Spectroscopic Studies 18.2.2 Lifetime Measurements 18.2.3 Other Phenomena Studied Using the Beam-Foil Interaction BEAM-GAS SPECTROSCOPY 18.3.1 Recoil Ion Spectroscopy 18.3.2 Exchange Reactions BEAM LASER SPECTROSCOPY OTHER TECHNIQUES OF ION-BEAM SPECTROSCOPY

213 213 214 214 215 215 216 216 216 218

18.1

INTRODUCTION

18.2

BEAM-FOIL SPECTROSCOPY

A beam of ionized atoms has important advantages as a spectroscopic source. Unlike arcs, sparks and high temperature plasmas, the ions in a beam can be studied in an environment that is free of electric and magnetic elds, and relatively free of interparticle collisions [1]. Standard accelerator techniques from nuclear physics may be used to produce a well-collimated, mass-analyzed beam of ions having a low velocity spread. In principle, virtually any charge state of any element, isotopically pure if required, can be obtained, given the appropriate equipment. Finally, the well-dened velocity of the beam permits the study of processes evolving in time in terms of their spatial evolution along the beam. This is particularly important in the case of lifetime measurements, discussed in Sect. 18.2.2. An experiment can involve an ion energy as low as a few keV or as high as several GeV. However, in this article the individual applications of ion beam techniques in atomic physics will be classied by the means used (foil, gas, laser etc.) to excite the ions, since this better denes the type of research in each case. 213

A beam of ions passing through a thin (50200 nm) foil emerges in a range of ionization states, with the mean charge state increasing with the incident energy [2]. For example, an F+ beam at 0.5 MeV emerges from the foil with a mean charge of about +2e, while an Xe+ beam at 180 MeV emerges with a mean charge of about +29e. The outer electrons are distributed over many dierent states, most of which then decay to lower states by photon emission as the ions move away from the foil. Thin foils made from a light element, usually carbon, are used to minimize particle scattering and energy straggling. Since many dierent electronic states are populated in the beam-foil interaction, it is a highly nonselective excitation. This has advantages for spectroscopic studies (see Sect. 18.2.1), but causes cascade problems for lifetime measurements. The major disadvantage of the beam-foil light source is its low intensity, and the resulting limitation on the spectroscopic resolution that is attainable. Consequently, beam-foil spectroscopy is often carried out in conjunction with more classical, higher resolution photographic spectroscopy (see Sect. 42.1) for detailed analyses of atomic spectra. Because of the low intensity of

Atomic, Molecular, & Optical Physics Handbook, edited by G. W. F. Drake, AIP Press, New York c 1996.

214 / Atomic, Molecular, & Optical Physics Handbook the source, beam-foil spectroscopy is usually performed using photon-counting techniques. the spectrometer, it is actually possible to use a wide entrance slit without losing resolution, giving intensity gains of several orders of magnitude. Beam-foil spectroscopy has also been used to make precision wavelength measurements of resonance transitions in hydrogen-, helium- and lithium-like ions in order to derive values for the Lamb Shift (see Sect. 28.3.9) [10]. In addition to the refocusing procedure mentioned above, if the measurement is made using a scanning spectrometer, it is important to investigate the possible error caused by periodic imperfections in the grating drive screw [11]. This problem is avoided if the spectrometer is used in a nonscanning mode, such as by employing a microchannel plate [12] or other detector array. It is also preferable to use calibration lines emitted by the moving ions themselves since a spectrometer that is focused for the moving ions will be out of focus for a stationary calibration source. One approach here is to use dierent grating orders to superimpose reference lines from other spectral regions on the spectrum containing the transitions under study [13]. Examples of beam-foil precision wavelength measurements for He-like ions [14, 15] are compared with calculations [16, 17] in Table 18.2.

18.2.1

Spectroscopic Studies

The beam-foil interaction involves an eective shielding of the outer electrons from the core as the ion passes through the foil. As the electrons emerge from the foil, some are re-captured by the core into a statistical distribution of states. Since the probability that more than one electron will be captured in an excited state is high relative to other light sources, beam-foil spectroscopy is used extensively to study doubly- and multiply-excited states [3] (see Sect. 62.1.1). The beam-foil interaction also favors the production of high-L Rydberg states [4] (see Sect. 14.1). At low incident ion energies, electron capture can give a downstream beam containing signicant fractions of neutral and even negative ions. The rst observation of photon emission between bound states in a negative ion was achieved using beam-foil excitation [5]. A recent example of multiple excitation in a negative ion is the identication in lithium-like He in the 2p3 4S 0 state in which all three electrons are excited [6]. This state has also been identied in several other members of the lithium-like isoelectronic sequence. Lifetime measurements and spectroscopic analyses often go hand-in-hand. For example, if two (or more) transitions are assigned as having the same upper state, then the shape of their intensity decay curves (see Sect. 18.2.2) should be the same within the statistical uncertainties, provided that none of the transitions is blended with another line. The time-resolution inherent in the beam-foil source can also be used to advantage in the assignment of transitions from long-lived states, such as intercombination transitions [7] (see also Sect. 10.17). Here it is useful to record the beam-foil spectrum both close to the excitation region at the foil and relatively far downstream from the foil. Far in this context means that the short-lived levels have had time to decay. The intercombination (or other long-lived) transitions can then often be easily identied by their relatively strong intensity in the downstream (time-delayed) spectrum. The low intensity of the beam-foil light source inevitably results in spectra of rather modest resolution. The high velocity of the radiating ions can also increase the widths of the spectral lines (see Sect. 19.1 for Doppler line broadening). A useful trick here is to refocus the spectrometer for the moving light source by adjusting the grating-to-slit distance appropriately [8]. While this method works well over narrow spectral ranges, it requires that the spectrometer be refocused for each wavelength region. Furthermore, the linewidth is still determined by the width of the entrance slit. An ingenious suggestion by Bergkvist [9] showed that, by introducing an appropriate cylindrical lens immediately before

18.2.2

Lifetime Measurements

Measurements Using Multi-Exponential Analysis


Since the beam-foil source is highly nonselective, the population of a given level j can be both depleted by spontaneous emission to lower levels k, and replenished by cascade transitions from higher levels i. The radiative lifetime (or mean life) of level j is given by the reciprocal of the sum of the transition probabilities Ajk . The population of level j is given as a function of the time t after excitation at the foil by the sum of exponential terms Nj (t) = A exp(t/ ) +
i

Bi exp(t/i )

(18.1)

where the coecients Bi can be positive or negative, and the summation includes any contributions from levels that repopulate level j, either directly or via intermediate levels. An actual measurement consists of recording a decay curve; viz., the intensity of a transition from level j as a function of distance from the foil. The intensity is proportional to the level population. The distance x corresponds to an elapsed time of x/v, where v is the ion velocity. The measurement is usually made by stepping the foil upstream, rather than moving the observation region downstream. However, scanning the beam directly by an appropriate optical system is also possible [18]. In a typical beam-foil decay curve, the data points are spaced further apart in the tail of the curve in order to estimate

Atomic, Molecular, & Optical Physics Handbook, edited by G. W. F. Drake, AIP Press, New York c 1996.

Ion Beam Spectroscopy / 215 the contribution of the long-lived cascade components in a reasonable total measurement time. Fitting the observed data to Eq. (18.1) to determine the lifetime of level j must be done with care (see also Sect. 8.2.2 for further discussion of curve-tting problems). Several computer routines have been developed, which attempt to determine the correct form of the tting function, i.e., how many exponential terms should be included, e.g., DISCRETE [19] and HOMER [20]. A useful technique here is to record the decay curve for each of the major cascade transitions into level j, and then include the lifetimes determined from analyses of these curves as xed parameters in tting the decay curve from level j to Eq. (18.1). An additional problem can arise in the measurement of very short lifetimes. These tend to be associated with short wavelength transitions for which no lenses are available to focus the beam on the spectrometer entrance slit. Consequently, the observation region extends along the beam. The eect of this averaging can be seen by integrating Eq. (18.1) from (t ) to (t + ). Since the rst data point that can be tted to (18.1) is at a time downstream, each coecient in (18.1) becomes multiplied by a factor [1 exp(2/ )], where is the lifetime for that component. Clearly, the shortlived components become very much weakened relative to the longer-lived cascade terms. Although this problem can be lessened at the cost of intensity by reducing the acceptance angle of the spectrometer, a preferable technique is to include the vignetting eect at the foil in the tting function itself. Routines to do this include FITB [21] and VNET [22], both of which permit the decay curve to start on the upstream side of the foil. Lifetimes in the picosecond regime have been reported using the beam-foil technique method [23]. concerns the resonance transition in Be-like ions, where the lifetimes of the 2s2p 1P o resonance level and the important cascade level, 2p2 1S, are almost equal. This problem has been studied by Engstrm et al. [30] o for the Be-like ions N IV and O V. Their results are summarized in Table 18.1, where they are compared with the weighted means of several previous beam-foil measurements and with two calculations [31, 32]. Two conclusions may be drawn from Table 18.1. First, the ANDC analyses give results in much closer accord with the calculations in this example. Second, consistency between several or many individual measurements of the same type (multi-exponential curve-tting of beam-foil data in this instance) does not guarantee that the results are not subject to signicant systematic error.

18.2.3

Other Phenomena Studied Using the Beam-Foil Interaction

Measurements Using Cascade Correction (ANDC) Techniques


A more rigorous approach to the cascade problem is provided by the ANDC method [24] discussed in Sect. 17.3.2. Various methods have been developed for applying the basic ANDC principle [2527]. The technique has been demonstrated [28, 29] to be very useful in cases where only a few cascades are important, as in atomic systems with a single active electron, (Nalike, Cu-like etc.) or with two active electrons, (Mg-like, Zn-like etc.), when it can yield reliable lifetimes with precisions of 5% or better. However, the technique tends to get results of rather lower precision when the cascade scheme is more complicated. The ANDC method is also useful when one or more strong cascades have lifetime components close to those of the primary decay, which can result in signicant errors if the analysis relies on only multi-exponential curve tting. One such example

Measurements of the velocity and scattering of the ions after passage through foils of known constitution and thickness have yielded data on electronic and nuclear stopping powers (see Sect. 87.2.1). Studies of the relative populations in the excited states of the ions have been used to model the ion-foil interaction itself [33]. The polarization of light emitted by ions after passage through a foil has been studied, both using linear polarization to study alignment of the atomic sublevels and circular polarization to study their orientation (see Sect. 12.2). Remarkably large degrees of circular polarization (up to 30%) have been observed using tilted foils, where foils have been tilted through large angles relative to the beam direction [34]. Of particular interest has been the study of quantum beats [35], in which the beam-foil decay curves show oscillations corresponding to the separation of close-lying ne- or hyperne-structure levels that are coherently excited as a result of the very short duration ( 1014 s) of the excitation process (see Sect. 17.3.4 for further details).

18.3

BEAM-GAS SPECTROSCOPY

While the use of a gas target in place of a foil has the obvious advantage that the gas cannot break, the loss of a tightly-localized excitation region means that the ne time-resolution of the beam-foil light source is largely lost. The beam-gas source has two main advantages, however. First, the passage of a beam of fast, highly-stripped ions through a neutral gas results in the production of recoil ions (see also Sect. 63.3) that are moving very slowly relative to the beam ions, thus avoiding the Doppler broadening problems mentioned in Sect. 18.2.1. Second, it is possible to study details of the interaction between the gas and the beam, such

Atomic, Molecular, & Optical Physics Handbook, edited by G. W. F. Drake, AIP Press, New York c 1996.

216 / Atomic, Molecular, & Optical Physics Handbook as the charge-exchange reactions, as they occur rather than merely observing their consequences. Dierences between the emission spectrum produced by the beamgas interaction and that produced by, say, interaction of the same ion beam with a foil, can give important clues concerning resonance processes for charge exchange or excitation exchange. Such experiments have experienced a resurgence recently with the advent of the ECR ion source [36], by which it is possible to obtain relatively high beam currents of highly-charged ions. (see Sect. 61.2.2). Recent work has shifted from measurements of the total cross section for electron capture in a collision, to more detailed studies of the individual capture channels [47]. Such studies provide much more stringent tests of theoretical models of the ion-atom charge transfer processes. They often involve such techniques as spectroscopy of the optical radiation or of the Auger electrons (see Sect. 25.1.1) emitted by the ions after electron exchange. This topic is covered in Chaps. 49, 62 and 63. A further example is the study by Prior et al. [48] of the angular distribution of Auger electrons emitted by doubly-excited states formed in hydrogen-like projectile ions with an energy of 40 keV, following doubleelectron capture from target helium atoms. They found that signicant alignment of the magnetic substates of the projectile ions can result from electron capture. Such anisotropies in the Auger electron emission demonstrate the danger of using single-angle measurements to determine cross sections.

18.3.1

Recoil Ion Spectroscopy

When a fast beam of highly-charged ions passes through a neutral gas, it leaves a trail of ionized gas atoms in its wake. These ions recoil from the interaction region with relatively low velocities (v/c = 104 105 ). Furthermore, the recoil velocities tend to be restricted to a narrow range of angles approximately perpendicular to the beam direction. Hence, observation of the radiation emitted by the recoil ions transverse to their motion; i.e., along a direction parallel to the beam, gives a very low Doppler width, D / being typically 106 , the limit imposed by the thermal motion of the target gas atoms. Recoil ion spectroscopy is therefore a useful procedure for precision wavelength measurements highly-stripped ions, such as He-like Ar16+ recoil ions produced by a beam of 2 GeV U70+ ions [37]. Some examples of wavelength measurements using recoil ions [38, 39] are compared with calculations [16, 17] in Table 18.2. The energies and charge states of the recoil ions may be determined using standard time-of-ight techniques [43], and detecting the recoil ions in coincidence with their progenitor projectile ions yields information on the dependence of the recoil energy on the details of the ionizing collision [44]. Detecting optical emission between Rydberg states of the ions (see Sect. 14.1) in coincidence with either the projectile or recoil ions can also give information on the role played by autoionization (see Chap. 25) by an ion between its formation in the collision and its arrival at the detector, as in the measurements by Martin et al. using Kr16+ ions with an energy of 360 keV [45]. Lifetimes of metastable states in highly-stripped recoil ions can also be measured by forming a secondary beam from the recoil ions. A position-sensitive detector is then used to record in-ight x-ray emission in coincidence with the charge state of the radiating ion [46].

18.4

BEAM LASER SPECTROSCOPY

18.3.2

Exchange Reactions

Measurements made on the projectile ions, rather than the recoiling target ions, also yield useful information about electron-capture processes. The strength of such a process is described in terms of its cross section

As for the beam-foil source, excitation of a beam of fast ions by a transverse tunable laser produces the localized excitation required for high temporal resolution. Now, however, the excitation is highly selective, permitting the population of just a single level. The laserinduced-uorescence (LIF) signal is therefore described by a single exponential decay, for which an exact analysis with rigorous error bounds is possible. The obvious restriction to levels that can be accessed by electric dipole (E1) transitions from the ground or metastable levels may be overcome by combining laser excitation with a nonselective mode of excitation, as has been done in beam-gas-laser [49] or beam-foil-laser [50] measurements. Discussion of these precision lifetime measurements is given in Chap. 17. Here the discussion is restricted to precision atomic spectroscopy. Some recent examples of precision beam-laser wavelength measurements [40, 42] are compared with calculation [16, 41] in Table 18.2. In addition, a recent high-precision measurement of the spin-forbidden 1s2s 1S0 1s2p 3S1 interval in N5+ [51] yields 986.321(7) cm1 , in reasonable agreement with the theoretical value 986.58(30) cm1 [16]. A major aim in beam-laser spectroscopy is to minimize the width of the LIF signal. The instrumental linewidth of the laser itself can be reduced to below 1 kHz, so that the width of the LIF signal is usually dominated by the velocity spread of the ions and by the divergences of the ion and laser beams. The eects of beam divergence can be minimized by using the collinear geometry, in which the ion and laser beams are parallel. If the angle between the ion and laser beams is , the Doppler-shifted laser frequency, as measured in the ions rest frame, is fL (1 cos ), where fL is the laser

Atomic, Molecular, & Optical Physics Handbook, edited by G. W. F. Drake, AIP Press, New York c 1996.

Ion Beam Spectroscopy / 217

Table 18.1. Beam-foil lifetime measurements by Engstrm et al. [30] for the 2s2p 1 P o level in Be-Like N3+ and O4+ o ions. Units are ns. Experiment Ion N3+ O4+
a b c d e

Theory ANDC 0.425 0.015 0.338 0.015 a 0.429 0.348 b 0.421 0.343

Multi-Expl. Fitting 0.50 0.03 (0.35 0.03c ) 0.39 0.02 (0.29 0.02c )

Other Expt. 0.463 0.01d 0.402 0.014e

Two-valence-electron model-potential calculation [31]. Intermediate-coupling multiconguration calculation [32]. Lifetime measured for the 2p2 1S cascade level. Weighted mean of 4 earlier experiments quoted in Ref. [30]. Weighted mean of 5 earlier experiments quoted in Ref. [30].

Table 18.2. Examples of fast-beam measurements of vac for the 1s2s 3S 1s2p 3P o transitions in He-like Ions. This Table compares experiment and theory for some low-Z, intermediate-Z and high-Z members of the He-like isoelectronic sequence. The theoretical uncertainties (shown in brackets) reect the leading terms omitted in the calculations by Drake [16]. Since these terms vary as Z 4 , a lower experimental precision is required at higher values of Z to provide a useful test of the calculation. Thus the recoil-ion measurements at intermediate-Z and the beam-foil measurements at high-Z provide useful comparisons with theory, even though their precision is much lower than that of the beam-laser experiments, which can only be performed for relatively low Z. Units are nm. Ion Be
2+

Technique Beam-Laser Beam-Laser Beam-Laser Beam-Laser Beam-Laser Beam-Laser Recoil-Ion Recoil-Ion Recoil-Ion Recoil-Ion Beam-Foil Beam-Foil

J J 10 11 12 10 11 12 10 12 10 12 12 12

Experiment 372.236904(5) 372.397129(4)a 372.190681(4)a 282.53674(10)d 282.66614(10)d 282.24601(7)d 127.774(4)e 124.8104(10)e 66.136(4)g 55.994(10)g 22.627(4)h 20.665(8)i
a

Theory 372.23625b, 372.23657(42)c 372.39704b, 372.39733(42)c 372.19082b, 372.19112(42)c 282.53616(64)c 282.66633(64)c 282.24652(64)c 127.7696(20)c, 127.7740f 124.8103(19)c, 124.8100f 66.1437(57)c, 66.1573f 55.9983(41)c, 55.9986f 22.6290(37)c, 22.6291f 20.6755(36)c

B3+

Ne8+

Ar16+

Ni26+ Cu27+
a Scholl

et al. [40]. et al. [41]. c Drake [16]. d Dineen et al. [42]. e Hallett et al. [38]. f Plante et al. [17]. g Beyer et al. [39]. h Zacarias et al. [14]. i Buchet et al. [15].
b Accad

Atomic, Molecular, & Optical Physics Handbook, edited by G. W. F. Drake, AIP Press, New York c 1996.

218 / Atomic, Molecular, & Optical Physics Handbook frequency and = v/c is assumed to be much smaller than unity. The range in frequency resulting from a divergence is thus fL sin , and tends to zero as tends to zero. One disadvantage of the collinear geometry is that, if the laser is brought into resonance with an atomic transition by adjusting the ion velocity and/or the laser frequency, the LIF signal is produced over the entire overlap region between the ion and laser beams. The resonance can be restricted to a desired region by setting the ion velocity to be slightly o-resonance. The velocity is then locally adjusted for resonance with the laser by passing the ions through a Faraday cage electrode to which a variable voltage is applied [52]. In the collinear geometry, the width of the resonance signal is usually dominated by the spread in the ion velocity c. The width resulting from a given may be diminished using kinematic compression. In terms of the ion energy E and the ion mass M , the Doppler width of the LIF signal is given by fD = fL = fL E/(2M c2 E)1/2 (18.2) low relative velocity between the ions and electrons. This permits measurements at the low energies of importance in studies of Rydberg state formation and in some astrophysical applications. The recent advent of storage rings has taken this concept one step further, as the stored ions may be cooled using lasers to longitudinal temperatures of a few kelvins [53]. (Laser cooling is discussed in Chap. 73.) As well as permitting very precise studies of electron-ion recombination and laser-stimulated recombination, it is also possible to measure very long lifetimes [54] and to search for condensed structures of ions similar to those that have been found in ion traps (see Chap. 73). Finally, in a very dierent type of experiment, the ionsurface interaction is studied using a well-collimated ion beam at grazing incidence on a clean, at surface. Such experiments have revealed very large atomic orientation [55]. This orientation can also be passed on to the nuclei of the atoms via the hyperne interaction, thus providing a source of oriented nuclei [56].

and thus decreases as E is increased [52]. A more signicant improvement in frequency resolution is made possible by including rf resonance in a laser doubleresonance experiment. Here, the ions are brought into resonance with an o-resonance laser using two separate Faraday cage electrodes. The rst resonance depletes the population in the ion state from which excitation occurs, thus weakening the second resonance signal. An rf eld is then applied to the ions between the two electrodes. Tuning the frequency of this eld over the region that corresponds to ne- or hyperne structure intervals in the ion can repopulate the state from which laser excitation occurs, thus re-establishing the second LIF resonance signal. The width of the resonance signal is now determined by the transit-time broadening that results from the nite time spent by the ions in the rf-eld (see also Sect. 74.6.1). For example, in the experiments of Sen et al. [52] with a beam of 131 Eu+ ions at 1.35 keV, the width of the laser-rf double resonance signal is 59 kHz, compared with a width of 45 MHz obtained using a single LIF resonance.

REFERENCES
1. I. Martinson, Rep. Prog. Phys. 52, 157 (1989). 2. K. Shima, T. Mikumo, and H. Tawara, At. Data Nucl. Data Tables 34, 357 (1986). 3. H. G. Berry, Phys. Scr. 12, 5 (1975); T. Andersen and S. Mannervik, Comments At. Mol. Phys. 16, 185 (1985). 4. F. G. Serpa and A. E. Livingston, Phys. Rev. A 43, 6447 (1991). 5. C. F. Bunge, Phys. Rev. Lett. 44, 1450 (1980). 6. E. J. Knystautas, Phys. Rev. Lett. 69, 2635 (1992); E. Trbert, P. H. Heckmann, J. Doerfert, and J. Granzow, a J. Phys. B 25, L353 (1993). 7. E. Trbert, Phys. Scr. 48, 699 (1993). a 8. J. O. Stoner and J. A. Leavitt, Opt. Acta 20, 435 (1973). 9. K.-E. Bergkvist, in Beam-Foil Spectroscopy, edited by I. A. Sellin and D. J. Pegg (Plenum Press, New York, 1976), p. 719. 10. J. Dsesquelles, Nucl. Instrum. Meth. Phys. Res. B 31, e 30 (1988); R. D. Deslattes, Phys. Scr. T 46, 9 (1993). 11. A. E. Livingston and S. J. Hinterlong, Appl. Opt. 20, 1727 (1981); R. DeSerio, ibid. 20, 1781 (1981). 12. J. A. Demarest and R. L. Watson, Nucl. Instrum. Meth. Phys. Res. B 24/25, 296 (1981). 13. R. DeSerio, H. G. Berry, R. L. Brooks, J. Hardis, A. E. Livingston, and S. J. Hinterlong, Phys. Rev. A 24, 1872 (1981); J. D. Silver, A. F. McClelland, J. M. Laming, S. D. Rosner, G. C. Chandler, D. D. Dietrich, and P. O. Egan, ibid. 36, 1515 (1987). 14. A. S. Zacarias, A. E. Livingston, Y. N. Lu, R. F. Ward, H. G. Berry, and R. W. Dunford, Nucl. Instrum. Meth. Phys. Res. B 31, 41 (1988). 15. J. P. Buchet, M. C. Buchet-Poulizac, A. Denis, J. Dse esquelles, M. Druetta, J. P. Grandin, X. Husson, D. Lecler, and H. F. Beyer, Nucl. Instrum. Meth. Phys. Res. B 9, 645 (1985). 16. G. W. F. Drake, Can. J. Phys. 66, 586 (1988).

18.5

OTHER TECHNIQUES OF ION-BEAM SPECTROSCOPY

While ion beams nd uses in many other applications, possibly the three main areas involve merged beams, storage rings and studies of the ion-surface interaction at grazing incidence. Merged beam experiments usually study recombination processes involving electrons and atomic or molecular ions (see Sect. 52.1 regarding recombination processes). The advantage of using merged electron and ion beams is that the time development of the processes may be studied spatially, while maintaining a

Atomic, Molecular, & Optical Physics Handbook, edited by G. W. F. Drake, AIP Press, New York c 1996.

Ion Beam Spectroscopy / 219


17. D. R. Plante, W. R. Johnson, and J. Sapirstein, Phys. Rev. A 49, 3519 (1994). 18. E. H. Pinnington, P. Weinberg, W. Verfuss, and H. O. Lutz, Z. Phys. A 281, 325 (1977). 19. S. W. Provencher, J. Chem. Phys. 64, 2772 (1976). 20. D. J. G. Irwin and A. E. Livingston, Comput. Phys. Commun. 7, 95 (1974). 21. P. H. Heckmann, E. Trbert, H. Winter, F. Hannebauer, a H. H. Bukow, and H. von Buttlar, Phys. Lett. A 57, 126 (1976). 22. E. H. Pinnington, W. Ansbacher, J. A. Kernahan, Z.-Q. Ge, and A. S. Inamdar, Nucl. Instrum. Meth. Phys. Res. B 31, 206 (1988). 23. C. L. Cocke, in Beam-Foil Spectroscopy, edited by I. A. Sellin and D. J. Pegg (Plenum Press, New York, 1976), p. 283; L. Engstrm and P. Bengtsson, Phys. Scr. 43, 480 o (1991). 24. L. J. Curtis, H. G. Berry, and J. Bromander, Phys. Scr. 2, 216 (1970); Phys. Lett. A 34, 169 (1971). vspace*1pt 25. E. H. Pinnington, R. N. Gosselin, J. A. ONeill, J. A. Kernahan, K. E. Donnelly, and R. L. Brooks, Phys. Scr. 20, 151 (1979). vspace*1pt 26. L. Engstrm, Nucl. Instrum. Meth. Phys. Res. 202, 369 o (1982). vspace*1pt 27. W. Ansbacher, E. H. Pinnington, J. A. Kernahan, and R. N. Gosselin, Can. J. Phys. 64, 1365 (1986). vspace*1pt 28. E. H. Pinnington, Nucl. Instrum. Meth. Phys. Res. B 9, 549 (1985). vspace*1pt 29. L. Engstrm, Phys. Scr. 40, 17 (1989). vspace*1pt o 30. L. Engstrm, B. Denne, J. O. Ekberg, K. W. Jones, o C. Jupn, U. Litzn, W. T. Meng, A.Trigueiros, and I. e e Martinson, Phys. Scr. 24, 551 (1981). vspace*1pt 31. C. Laughlin, E. R. Constantinides, and G. A. Victor, J. Phys. B 11, 2243 (1978). vspace*1pt 32. H. Nussbaumer and P. J. Storey, Astron. Astrophys. 74, 244 (1979). 33. T. berg and O. Goscinski, Phys. Rev. A 24, 801 (1981); A C. Jupn, B. Denne, J. O. Ekberg, L. Engstrm, U. e o Litzn, I. Martinson, W. Tai-Meng, A. Trigueiros, and e E. Veje, ibid. A 26, 2468 (1982). 34. H. G. Berry, L. J. Curtis, D. G. Ellis, and R. M. Schectman, Phys. Rev. Lett. 32, 751 (1974); S. Huldt, L. J. Curtis, B. Denne, L. Engstrm, K. Ishii, and I. Martinson, o Phys. Lett. A 66, 103 (1978). 35. H. J. Andr, Phys. Rev. Lett. 25, 325 (1970); H. G. a Berry, J. L. Subtil, E. H. Pinnington, H. J. Andr, W. a Wittmann, and A. Gaupp, Phys. Rev. A 7, 1609 (1973). 36. R. Geller, App. Phys. Lett. 16, 401 (1970); C. M. Lyneis and T. A. Antaya, Rev. Sci. Instrum. 61, 221 (1990). 37. J. M. Laming and J. D. Silver, Phys. Lett. A 123, 395 (1987). 38. W. A. Hallett, D. D. Dietrich, and J. D. Silver, Phys. Rev. A 47, 1130 (1993). 39. H. F. Beyer, F. Folkmann, and K.-H. Schartner, Z. Phys. D 1, 65 (1986). 40. T. J. Scholl, R. Cameron, S. D. Rosner, L. Zhang, R. A. Holt, C. J. Sansonetti, and J. D. Gillasby, Phys. Rev. Lett. 71, 2188 (1993). 41. Y. Accad, C. L. Pekeris, and B. Schi, Phys. Rev. A 4, 516 (1971). 42. T. P. Dineen, N. Berrah-Mansour, H. G. Berry, L. Young, and R. C. Pardo, Phys. Rev. Lett. 66, 2859 (1991). 43. G. P. Grandin, D. Hennecart, X. Hussin, D. Lecler, I. Lesteven-Vaisse, and D. Lis, Europhys. Lett. 6, 683 (1988). 44. J. C. Levin, R. T. Short, C.-S. O, H. Cederquist, S. B. Elston, J. P. Gibbons, I. A. Sellin, and H. SchmidtBcking, Phys. Rev. A 36, 1649 (1987). o 45. S. Martin. A. Denis, Y. Ouerdane, and M. Carr, Phys. e Lett. A 165, 441 (1992). 46. G. Hubricht and E. Trbert, Phys. Scr. 39, 581 (1989). a 47. M. Barat, Nucl. Instrum. Meth. Phys. Res. B 9, 364 (1985). 48. M. H. Prior, R. A. Holt, D. Schneider, K. L. Randall, and R. Hutton, Phys. Rev. A 48, 1964 (1993). 49. D. Schutze-Hagenest, H. Harde, W. Brand, and W. Demtrder, Z. Phys. A 282, 149 (1977); H. Schmoranzer o and U. Volz, Phys. Scr. T47, 42 (1993). 50. Y. Baudinet-Robinet, P.-D. Dumont, H.-P. Garnir, and A. El. Himdy, Phys. Rev. A 40, 6321 (1989). 51. E. G. Myers, J. K. Thompson, E. P. Gavathas, N. R. Clausen, J. D. Silver, and D. J. H. Howie, Phys. Rev. Lett. 75, 3637 (1995). 52. A. Sen, W. J. Childs, and L. S. Goodman, Nucl. Instrum. Meth. Phys. Res. B 31, 324 (1988). 53. S. Schrder et al., Phys. Rev. Lett. 64, 2901 (1990); J. o S. Hangst, M. Kristensen, J. S. Nielsen, O. Poulsen J. P. Schier, and P. Shi, ibid. 67, 1238 (1991). 54. T. Andersen, L. H. Andersen, P. Balling, H. K. Haugen, P. Hvelplund, W. W. Smith, and K. Taulbjerg, Phys. Rev. A 47, 890 (1993). 55. H. J. Andr, R. Frhling, H. J. Plhn, and J. D. Silver, a o o Phys. Rev. Lett. 37, 1212 (1976). 56. H. J. Andr, R. Frhling, H. J. Plhn, H. Winter, and W. a o o Wittmann, J. Phys. (Paris) 40, C1-275 (1979).

Atomic, Molecular, & Optical Physics Handbook, edited by G. W. F. Drake, AIP Press, New York c 1996.

You might also like