Scanning Microscopy

Volume 1982 Number 1 1982 Article 1

1982

Cross Sections for Inelastic Scattering of by - Selected Topics Related to Electron Microscopy

Mitio Inokuti Argonne National Laboratory

Steven T. Manson Georgia State University

Follow this and additional works at: https://digitalcommons.usu.edu/electron

Part of the Biology Commons

Recommended Citation Inokuti, Mitio and Manson, Steven T. (1982) "Cross Sections for Inelastic Scattering of Electrons by Atoms - Selected Topics Related to Electron Microscopy," Scanning Electron Microscopy: Vol. 1982 : No. 1 , Article 1. Available at: https://digitalcommons.usu.edu/electron/vol1982/iss1/1

This Article is brought to you for free and open access by the Western Dairy Center at DigitalCommons@USU. It has been accepted for inclusion in Scanning Electron Microscopy by an authorized administrator of DigitalCommons@USU. For more information, please contact [email protected]. Electron Beam Interactions With Solids (Pp. 1-17) SEM, Inc., AMF O'Hare (Chicago), IL 60666, U.S.A.

CROSS SECTIONS FOR INELASTIC SCATTERING OF ELECTRONS BY ATOMS SELECTED TOPICS RELATED TO ELECTRON MICROSCOPY*

Mitio Inokuti Argonne National Laboratory, Argonne, Illinois 60439

and

Steven T. Manson Department of Physics and Astronomy Georgia State University, Atlanta, Georgia 30303

ABSTRACT 1. INTRODUCTION

We begin with a resume of the Bethe theory, which pro­ Throughout the present article, we consider incident elec­ vides a general framework for discussing the inelastic scatter­ trons at kinetic energies of interest to microscopy, i.e., be­ ing of fast electrons and leads to powerful criteria for judging tween a few tens of keV and several MeV . As to targets, we the reliability of cross-section data. The central notion of the consider first neutral atoms having lower atomic number s theory is the generalized oscillator strength as a function of (say, Z < 30), and later molecules and solids composed of both the energy transfer and the transfer, and is those atoms. For the majority of the inelastic of the only non-trivial factor in the inelastic-scattering cross sec­ electrons with atoms thus delimited, the Bethe theory (Bethe tion. Although the Bethe theory was initially conceived for 1930, 1932, 1933) is well justified, and provides a good free atoms, its basic ideas apply to solids, with suitable gen­ framework for general understanding and for numerical eralizations; in this respect, the notion of the dielectric evaluation of cross sections (Inokuti, 1971; lnokuti et al., response function is the most fundamental. Topic s selected 1978). for discussion include the generalized oscillator strengths for A necessary (though not sufficient) condition for the first the K-shell and L-shell ionization for all atoms with Z ~ 30, Born approximation, used in the Bethe theory, is that the evaluated by use of the Hartree-Slater potential. As a func­ mean orbital speed of the atomic electron pertinent to the in­ tion of the energy transfer, the generalized oscillator strength elastic be small compared to the incident electron most often shows a non-monotonic structure near the K-shell speed. For ionization (or excitation) of an inner shell by rela­ and L-shell thresholds, which has been interpreted as mani­ tivistic electrons, the condition means that the effective­ festations of electron-wave propagation through atomic charge number I seen by an atomic electron in that shell be fields. For and solids, there are additional struc­ substantially smaller than 137; 1 is somewhat smaller than tures due to the scattering of ejected electrons by the fields of the atomic number Z, and the condition is fulfilled for mod­ other atoms. erate Z, (say Z < 30). For the lowest incident energies of electrons we consider, the condition is satisfied only for lower Z. The condition discussed above literally applies to quantita­ tive discussion of cross sections. Even when the condition is not quite fulfilled, however, often results of the first Born approximation are useful; they may be good as qualitative guides and may be reliable to modest accuracy (perhaps with­ in a factor of two). This is especially the case for inelastic col­ *Work performed under the auspices of the U.S. Depart­ lisions resulting in an optically allowed transition and in ment of Energy and of the U.S. Army Research Office. small scattering angles; then, the impact parameter is large and thus the incident electron travels well outside the target . This recognition is readily verifiable in a variety of em­ pirical data, and is in effect expressible in a more rigorous Keywords: Bethe theory, inner shells, generalized oscillator theoretical form (Lassettre et al., 1969). strength, momentum transfer, cross section, inelastic scatter­ Figure 1 exemplifies differential cross sections plotted ing, non-hydrogenic, systematics, Bethe surface, extended against the scattering angle 0. The figure shows the cross sec­ x-ray absorption fine structures (EXAFS), hydrogenic ap­ tions for collisions of 25-keV electrons with neon, most of proximation. the data being taken from Geiger (1964). Notice that all cross sections are peaked at small ang les and that the plot is doub ly logarithmic. The elastic-collision cross section is virtua lly flat Contact: Mitio Inokuti, Argonne National Laboratory at small 0, because the interactions between the electron and Argonne, Illinois 60439 (312) 972-4186 the atom effective ly have a short range in this case. The potential for these interactions decreases with distance r as r-•

1 M. lnokuti and S.T . Manson

LIST OF SYMBOLS p ' Momentum of an electron after col­ lision The Bohr radius, lJ2 / me 2 = 0. 529 I 8 Radial wavefunction for a bound nf x 10-• cm state Impact parameter Radial wavefunction for a continu­ Electron um ff state The light speed in vacuum Q The recoil energy defined by Eq. (6) The constant in Eq. (I I) r Radial distance The dimensionless constant in Eq. !:_j Position vector of the j-th atomic ( 13) electron C ,I The normalization constant for a R The Rydberg energy , me• / 2l12 = continuum ff state, Eq . (30) 13.606 eV Differential cross section for energy­ R( €, f' , n, f, >.,K) The radial integral defined by Eq. transfer values between E and E + (20) dE T of an incident elec­ daR The Rutherford cross section tron df(K,E) / dE The density of the generalized oscil­ u The total effective potential defined laior strength per unit range of ener­ by Eq. (29) gy transfer E, for momentum trans­ V Speed of an electron fer l!K V (r) The potential of the atomic field of dT / dx Stopping power force seen by an electron dw Solid-angle element The potential in the hydrogenic ap­ e The charge on an electron proximation E Energy transfer from an incident Vo The constant in V Hyd(r), Eq. (27) electron to a target, viz., excitation X The variable defined by Eq. (12) energy of the target y Electron-exchange term in Eq. ( 15) Mean energy transfer per inelastic Y em(O,q,) Spherical harmonic function collision z Atomic number l, The rationalized Planck constant Zerr The constant in V Hyct-and argument Kr o( Dirac delta function K Wave vector transferred from an in­ Kinetic energy of an ejected electron cident electron to a target ( ll(S is Eigenenergy of the nf state momentum transfer) The complex dielectric res pon se The maximum value of the magni­ function, which depends on momen ­ tude of K tum transfer l1 ISand energy transfer The minimum value of the magni­ E tude of K Effective-charge number (effective k The wave number of an electron atomic number that characterizes an f The orbital angular-momentum atomic field seen by an inner-shell quantum number in an initial bound electron) state The atomic matrix element defined f' The orbital angular-momentum by Eq. (5) quantum number in a final continu­ e Scattering angle um state e Polar angle The magnetic quantum number e" Aperture angle The electron rest mass Cross section for energy-transfer Dipole matrix element squared for values between E and E + dE energy transfer E, i.e., a constant in Total inelastic-scattering cross sec­ Eq. (I I) a w1 tion Dipole matrix element squared for The total photoionization cross sec­ total inelastic scattering, i.e., a di­ tion of the K shell of molecular mensionless constant in Eq. (13) nitrogen, shown in Fig. I 7 n Principal quantum number ct> Azimuthal angle N The number density of atoms I O) The ground state of an atom p Momentum of an electron before I E) An excited state of an atom, at ex­ collision citation energy E

2 Cross Sections for Inelastic Scattering or more rapidly. The inelastic-collision cross sectio ns depend As a prelude to discussion on cross sections, we may note on 0 more stro ngly and behave as 0-2 over a moderate range here certain contrasts between the valence shell and the inner of 0, show ing that the interactions are of long range (due to shells. The valence shell has a linear dimension of the stan­ the instantaneous dipole moment associated with the atomic dard atomic size, i.e., of the order of the Bohr radius ao = transition). The potential for these interactions decre ases n2 / me 2 = 0.52918 x 10-s cm, and has binding energies of 2 2 with distance r as slowly as r- • At larger 0, the 0-dependence the order of the Rydberg energy R = me• /211 = 13.606 eV. of the inelastic-collision cross section is stronger; the onset of The electronic structure of the valence she ll is seriously influ­ the stro nger dependence is different for different atomic enced by subtle effects of many-electron correlations or of shells invo lved . It is the Bethe theory that enables one to see the atomic environment (i.e., chemical bonds and condensed­ precisely all these features of the cross sections and to phase formation). Conseq uentl y, experimental spectra of the und erstand how they come about. valence she ll are rich in general, and often contain keys for unraveling those subtle effects; theoretical calcu lation of the cross section for valence-shell excitation or ionization is com­ plicated in general and is often difficult in practice . By con­ 3 trast, an inner shell has a much sma ller linear dimension (of the order of a 0 / 1) and a much greater binding energy ( of the order of 12R). Many-electron correlations or atomic-environ­ ment effects influence the electronic structure of an inner shell only modestly. Thus, experimental spectra of the inner 2 shell. are governed roughly by the atomic number Z, and therefore often serve as a means of elemental chemical anal­ ysis. The simple picture of the inner-shell spectra is often taken for granted , but is in fact subject to a provision. The simple picture is right so long as an ejected electron is much 3 more energetic compared to the potential of its interaction s "O b with the ion core left behind and with the atomic environ­ "O---- N ment. Otherwise, the ejected electron is slow enough to "see" ' o 0 0 details of the potential, and gives rise to various observable 2 consequences in the inner-shell spectra. Much of the discu s­ 0, 0 sion in Section 3 will concern this topic. The present article is in effect a continuation of an earlier article (lnokuti, 1978) also written for the electron micro s­ -I I \ copist. For the reader of that article, the following will serve \ \ K-Shell Ionization as an update with an emphasis on newer findings on inner­ she ll ionization.

-2 2. ELEMENTS OF THE BETHE THEORY

2. I Basics Suppose that an electron of speed v collides with an atom -3 -2 0 and excites it to a higher state, either discrete or continuum, log 8 at excitation energy E measured from the ground state. The 10 kinetic energy of the electron then will be reduced by E, Fig. I. Differential cross sections for collisions of 25-keV which may be called the energy loss (from the incident elec­ --- electrons with a neon atom. tron) or the energy transfer (to the target atom). The direc­ The horizontal axis represents the common loga­ tion of the electron motion may be deflected by angle 0, rithm of the scattering angle 0. The vertical axis wh ich is called the scattering angle. represents the logarithm of the cross section per unit The first point of Bethe is that the momentum transfer solid angle d a / d w measured in the squared Bohr 11~ = !? - !?' , where!? is the momentum before the collision radius ao2 = 0.280 x 10-•• cm 2 • The curve labeled and p' is the same after the collision , is the key variable for "Elastic" shows the elastic-scattering cross section. analyzing any collision of fast particles. The magnitude 11~ is The curve labeled "E 16.9 eV" represents the cross = readily calculable from 0, E, and v by use of elementary section for the excitation to the 2p53s state. The kinematics. For electron energies not negligible compared to curve labeled "E = 20 eV" represents the cross sec­ mc 2 511 keV, one must use relativistic kinematic s. tion for the excitation to 2p 54s and all higher states = The notion of the momentum transfer 11~ may be most combined. All the above are based on data given by Geiger (1964). The curve labeled "K-Shell Ionization" easily understandable when one relates it to the notion of the is based on the theoretical generalized oscillator impact parameter b used in classical mechanics. Indeed, the strength calculated by the present authors. The bro­ two notions are complementary in the sense that the relation ken straight lines are drawn to show the 0-2-depen­ dence of the cross sections, valid at a range of inter­ Kb= 1 (I) mediate 0 values.

3 M. lnokuti and S.T. Manson holds for the majority of collisions. In other words, collisions Here it is appropriate to recall the Rutherford formula, at large b are called soft or glancing, and result in small K; which applies to collisions of two free charged particles. collisions at small b are called hard or knock-on, and result Specifically for a collision of an electron with a free and sta­ in large K. Nevertheless, there is a fundamental distinction; tionary electron, the Rutherford cross section reads the momentum transfer is unambiguously defined in all cases, while the impact parameter is not a quantum-mecha­ (8) nical observable. [See Section 4.4 of lnokuti 1971 and Bohr 1948.] and Q represents in this hypothetical instance the kinetic For fixed v and E, the momentum transfer n~ may take a energy of the recoiled electron. Thus the meaning of the form range of values depending on 0. The smallest value of nK factor ! JJE(K) I' becomes clear; it represents the ratio of the occurs when 0 = 0, and is given by atomic cross section d a E to the Rutherford cross section . It is the only nontrivial factor in d a E in the sense that its evalu­ nK min E / v. (2) = ation by means of Eq . (5) presumes knowledge of atomic structure. To summarize, we owe to Bethe (then twenty -fo ur To derive this, one relates the change 6p in the electron years old) the crucial recognition that d a E factorizes into the momentum p with the change 6 T in the kinetic energy T as 6p = (dp / dT)6T, sets 6p = nKmin and 6T = E, and Rutherford cross section (which depends on the incident­ particle variables only) and the form factor (which depend s notes dT / dp = v. The derivation, as well as the result [Eq . on target properties but not explicitly on the incident speed v (2)], is correct in both relativistic and non-relativistic kine ­ or any other incident particle variable). For more detailed matics. Notice that nKm in for any inelastic collision is never commentary,see lnokuti (1971) and lnokuti (1978). vanishing, although it becomes smaller and smaller with in­ Finally, the generalized oscillator strength df(K,E) I dE per creasing v or with decreasing E. The largest value of nK for unit range of E is defined by fixed v occurs when 0 = 71", and is about twice the incident momentum, i.e., df(K,E) / dE = (E / Q) ! JJE(K) I'- (9)

(3) Equivalently, we may write where (3 = v / c and m is the electron rest mass. Thus, nK max df(K,E) / dE = (E / R)(Ka of' ! JJE(K) I'- (10) is in general large and increa ses without bounds as v - c. The second point of Bethe concerns the differential cross The equivalence of Eq. (10) with Eq. (9) is apparent as soon as 4 2 sect ion for energy transfer values between E and E + dE one recall s that Rao'= (me /2 n )(n 2 / me 2 ) ' = n 2 / 2m. The term "genera lized oscillator strength" is another inova­ da E = 4ao' (p '/ p)(Kaot• I JJE(K) l' 27rsin0d0, (4) tion of Bethe . As K - 0, it redu ces to the optical (dipole) oscillator stre ngth, which governs the light abso rption and where JJE(K) is an atomic matrix element practically all optical properties of the atom under considera­ tion. For the basics of photoabsorption by atoms, see Fano z and Coo per ( 1968), Man son (I 976), Manson (1977), Manson JJE(K) = (E I j:Jexp(i~•r) I 0) (5) (1978), Manson and Dill (1978), and Starace (1982) . taken between the excited state ( E I and the ground sta te 2.2 The Bethe Surface I 0), r_i being the position of the jth atomic electron. The The main object of study is the form factor I JJE( K) I' or quantity ! JJE( K) I' is called a form factor for inela stic scatter­ the generalized oscillator-strength density df (K,E) / dE as a ing, and ma y be taken as an even function of scalar K, so function of both nK and E . To make this point clear, lnokuti long as the target atoms or molecules are randomly oriented. (1971) used the term "Bethe surface ." Equations (4) and (5), as well as several equations to follow, We have already discu ssed the connection with the photo­ are written specifically for non-relativi stic speeds v, for the absorption, which corresponds to the limit (K - 0) we men­ sake of compact expression. Notice that d a E ha s the dimen­ tioned at the end of the last Subsection 2.1. sion area' energy, -, and 1JE( K) I' has the dimension ener­ I At larger E values, df(K,E) / dE substantially differs from gy-'. zero only when Q of Eq. (6) nearly equals E and far exceeds We may rewrite Eq. (4) to express d a E in terms of the an atomic-shell binding energy. Then, df(K,E) / dE shows a momentum tran sfe r nK or other related variables. For in­ marked peak at those values of Kand E which correspond to stance, one may introduce a variable with the energy dimen­ free-electron collision thus satisfying the relation Q = E. In­ sion, i.e., okuti (1971) called the peak the Bethe ridge, and emphasized its univer sal occurrence. Q (nK)2 / 2m (6) Figure 2 shows the Bethe surface for atomic hydrogen (lnokuti 1971). Figure 3 shows two examples that have been and write determined by experiment (Lahman-Bennani et al., 1979, 1980).

4 Cross Sections for Inelastic Scattering

~')5 2 3 24

~ ~. co, (a) 2 0 i! i:

0 JO 50 90 ,20 150 '80 2!0 Et R

;os rJ2

' ~ 2 . ~Hl s ,< 16 ,L "1 I( t d Q•

(b) ( b) oa l J ,.I "~

'

Fig. 3. The Bethe surfaces of carbon dioxide (Fig. 3a) and ammonia (Fig. 3b), after Lahmam-Bennani et al. (1979), 1980), reproduced with permission by the Fig. 2. The Bethe surface for atomic hydrogen . authors and the publishers. --- The horizontal axes for E / R and In (Ka )2 define The base plane is defined by E/ R (the energy trans­ 0 fer measured in the Rydberg energy, 13.6 eV) and by the base plane. The vertical axis represents df(K),E) / In[ (Kao)2 (in the notation of the present article). d(E/R). The fourteen plates are placed E/R = 3/4, I The dashed curve indicates the Bethe ridge for 8/ 9, 1, 5/ 4, 3/2, 2, 3, 4, 5, 6, 7, 8, 9, and 10. The valence electrons. Notice the K-shell ionization con­ broken curve on the base plane shows the location tributions near the threshold energies of carbon, (E/ R) = (Kao)2 of the Bethe ridge, which is the main nitrogen , and oxygen atoms . feature for E/R i,> I; collisions represented by a point near the Bethe ridge occur as though the inci­ dent particle were to strike a free electron, the elec­ tron binding being of secondary importance. The optical region (Kao)2 ~ 1 is conspicuous only for The study of the Bethe surfac e is a rich subject with many small E/ R. Figure 2a shows the gradual spreading of applications and impli catio ns to diverse phenomena. Just to the Bethe ridge with decreasing E/ R, and eventually name several examples, we may start with sum rules, which its merger with the optical plateau at the region of usually mean theorems on the integrals involving df(K,E) / small (Kao)2 and E/R. Figure 2b shows in front a cut dE with respect to E (including sum s over discrete spectra), at at In[ (Kao)2I = - 4, i.e., a curve that closely ap­ fixed K. These sum rules (Section 3 .3 of lnokuti 1971) are proximates the photoabsorption cross section. This often usefu l as control on data. There are also theorems on figure is taken from Inokuti (1971). the int egra ls involving df(K,E) / dE with respect to K, at fixed

5 M. lnokuti and S.T. Manson E. (See Section IIE of Jnokuti et al., 1978, and Matsu zawa et al., 1979). If the Bethe surface is drawn on the plane with Cartesian axes representing E and In K, then the volume where X is the variable defined by Eq. (12). The term Y under the surface delimited by appropriate kinematic limits represents electron-exchange effects; it is a function of {3 and represents the stopping power of the target atom for any fast amounts to about ten percent of the other terms in the square charged particle (Bethe I 930, and Section 4.3 of lnokuti brackets. For precise treatment, see Bethe (1932, 1933) and 1971). The shape of the Bethe ridge is a renection of electron Section 4.3 of Inokuti (1971). More importantly, Eq. (15) binding in atoms, or more precisely, the electron momentum contains a single nontrivial property of the medium atom, distribution, and is connected with the Compton profile, i.e., i.e., the mean excitation energy I, which is defined in terms the spectral distribution of high-energy photons scattered by of the dipole oscillator-strength den sity as atomic electrons (Bonham and Wellenstein 1973, Wong et al., 1975, Barias et al., 1977, and Lahman-Bennani et al. In I = J InE[df(0,E) / dE]dE / J [df(0,E) / dE]dE (16) (1979, 1980)). Finally, the so-called (e,2e) measurements, i.e., coincidence measurements of scattered electrons and The integral in the denominator is equal to the atomic num­ ejected electrons resulting from collisions corresponding to ber Z according to the well-k nown su m rule . The mean ex­ the Bethe ridge, represent an area of many recent studies citation energies for various materials are an object of exten­ (McCarthy and Weigold 1976). sive studies, as seen in Inokuti and Turner (1978) and in Ino ­ kuti et al. (1981). The ratio 1/Z is a function of Z, exh ibit s 2.3 Integrated Cross Sections and Their Systematics minima and maxima renecting the atomic shell structure for lower Z, and tends to a limiting value of about 10 eV for high The third point of Bethe concerns the cross section in­ z. tegrated over all scattering angles. Starting with Eqs. (4)­ As an application of Eqs. (13) and (15), one may consider (7) and using general propertie s of the form factor or of the the mean energy tran sfer per inelastic collision, viz., generalized oscillator strength, one can show that

E av = a s/ a tot· (17) (11) This quantity is of great inter est to the electron microscopist where for co nsideration of the radiation damage of specimens and other related matters. Figure 5 shows the mean energy trans­ X ln[/:/2/ (1 - 132)] - {32, (12) = fer E av per inelastic collision, for 50-keY electrons (chained dash) and for at the same speed (solid line), as func­ and ME2 and CE are atomic properties derivable from tions of Z. As Inokuti (1978) pointed out, elec tron df(K,E) / dE. What is most important here is that the depen­ microscopists often use theoretical values of a and E av dence of a E on the electron speed v = {3c is analytically given 101 based on the Thomas-Fermi model of atoms . This model and is universal for all targets. treats all atomic electrons as a free-electron , disregards As a conse quence, the total inelastic-collision cross section the atomic shell structu re, and therefore naturally predicts a , i.e., the sum of all inelastic-collision cross sections, is 101 a tot and E av as smooth functions of Z. given by a formula of the same general structure, i.e., 2.4 Condensed Phases Bethe treated free atoms as target. Extension to free mole­ cules is formally stra ightforward. In the definition of the where M / and C are atomic properties that often allow 10 101 matrix element lJE(K) [Eq . (5)] molecular eigenfunctions acc urat e evaluation. (See Section 4.3 of lnokuti 1971 and lnokuti et al., 1967). An application of this result is now must be used, and the rotational and vibrational degrees of demonstrated in Fig. 4 and Fig. 5 (adapted from lnokuti et al., freedom must be accounted for. Despite the complications, 1981). Figure 4 show s a for 50-keV electron for all atoms the theory remains basically unchanged . For fuller discus­ 101 sion, Section 3.5 of Jnokuti (1971). with Z ~ 38. Notice the periodic variation with Z, due to the well-known structure of atoms . Extension to condensed phases began with the work of Fermi (1940), who pointed out what we call the den sity effect Another consequence of Eq. (I I) is the famous Bethe for­ mula for the stopping power dT / dx, i.e ., the mean energy on energy losses. For a relativistic particle traveling through loss per unit path length of a charged particle penetrating condensed matter, the relevant impact parameter may be­ through a medium . It may be evaluated as come so large that there are many medium atoms between the particle and a particular atom that becomes excited. To see this, recall Eqs. (1) and (2); the maximum impact of para­ dT / dx = N j E aE dE (14) meter is I / K min = nc / E, and becomes 2 x 137 ao = 145 A for a medium having the number density of atoms N, where for E = R. The medium atoms are instantaneously polarized the integration runs over the whole range of energetically by the electric field of the particle and tend to screen the par­ possible E values belonging to continuous and discrete spec­ ticle interactions with the atom that eventually receives tra. The integral is called the stopping cross section a,t, and energy. Fermi used a macroscopic description according to has the dimension area 2 energy. The Bethe formula for an electrodynamics, as summarized by Landau and Lifshitz electron may be written as (1960) .

6 Cross Sections for Inelastic Scattering

025 More detailed treatments were developed in the 1950's by many workers including Fano ( 1956), Ferrell (1957), and

0.20 Nozieres and Pines (1959). Special attention was paid to plas­ _/ mon excitations in metals, which then began to be studied through electron energy-loss measurements . Recent work on ~------0.15 plasmon excitation is reviewed by Raether (1980). The fol­ .., lowing paper by Powell (I 983) in the present Conference 0 .., 0.10 describes general features of electron inelastic scattering in b solids and relevani cross sections. Here we make only a few remarks. The three major points 0.05 of Bethe (i.e., the role of the momentum transfer, the factor­ ization of the cross section into the Rutherford factor and the 0.00 form factor, and the analytic structure of the integrated cross 0 10 20 30 40 section) all remain true for condensed phases . The form­ z factor idea is generalized, and it is customary to use the com­ plex dielectric response function E( If,E) for describing the Fig. 4. The total inelastic-scattering cross section a (mea­ 101 effects of electromagnetic perturbation associated with sured in A 2 ) for 50-keV electrons, plotted against angular frequency E/11 and propagation vector K. The func­ atomic number Z. tion may be interpreted also as the Fourier transform of the The solid line shows the result of calculations that electron density fluctuation in the medium . For charged­ incorporate relativistic kinematics for the incident particle interactions, the quantity E Im [ - I / E (If ,E)] plays electron. For comparison, the dotted line shows the the role of the generalized oscillator strength df(K,E) / dE. results of calculations that disregard relativistic kine­ The use of the complex function E(If ,E) entails studie s on matics. The figure is taken from lnokuti et al. (1981). the analytic properties, especially on the integral relat ions between the real and imaginary parts, called Kramers-Kronig dispersion relations . Thorough exploitation of the these rela­ tions has been carried out for several instances, e.g., metallic aluminum (Shiles et al., 1980), but only for data at K = 0.

140.0 3. SELECTED TOPICS I 120.0 3.1 Generalized Oscillator Strengths for Inner Shells of Atoms: Methods of Calculations 100.0 Earlier calculations on the generalized oscillator strengths of inner shells and related quantities were based on the $' ao.o 11) hydrogeni c approximation (Walske 1952, Walske 1956, and ~ 80.0 Khandelwal and Merzbacher 1966, to name just three exam­ ra:l 11/ ples). In this scheme , one uses for one-electron eigenfunc­ 40.0 tions for both the initial state and the final state hydrog enic I/ function s, but accounts for the screening by other electrons 20.0 by means of a suitable effective nuclear charge by an adjust­ ment of the energy scale to fit the experimental ioniza tion 0.0 threshold. Then, the generalized oscillator strength s may be 0 5 10 15 20 25 30 35 40 readily evaluated analytically. Yet, the procedure is intrinsic­ z ally unrealistic for values of the energy transfer E compara­ ble with the ionization threshold; this deficiency is serious Fig. 5. The mean energy transfer E AV per inelastic collision because much of the strength lies precisely at those E values (measured in eV), plotted against atomic number Z. for small and moderate K values. The general features of the curve Z-dependence are the same for all charged particles incident with suffi­ Manson ( 1972a, 1972b) initiated more realistic calcula­ ciently high speed. The solid line is for a at tions , within the one-electron orbital picture. In this picture , one approximates the ground state I0) of the whole atom by 91.8 MeV (viz., {3 = 0.4127). The chained dash line is a suitably antisymmetrized product of one-electron orbitals for an electron at 50-keV (viz., the same {3).The stop­ of the form r· 1 P y(r) Yem(0,¢), where (r, 0, ct>)are the ping power for an electron is smaller than that for a 0 proton of the same speed when one accounts for the spherical coordinates of an atomic electron , P nr (r) is the exchange of the primary electron with a secondary radial function with the principal quantum number n and the electron in close collisions. The dotted line shows a orbital angular-momentum quantum number e,Y em( 0, ¢ ) is limiting E AV for any extremely relativistic charged the spherical harmonic , and mis the magnetic quantum num­ particle. The figure is based on the data given by ln­ ber. At the same time, one approximates the excited state IE) okuti et al. (1981). in the continuum by r· 1 P ,r· (r) Y r· m<0, ct>),where P ,r- (r) is M. lnokuti and S.T. Manson the radial function representing an electron ejected with 112 d 2 P .r n2 f(f+ 1) kinetic energy E and angular momentum f'. For the ioniza­ --- + [ E-V (r) - --- ] P,r = O, (25) tion of the nf shell having the binding energy I nl' , E is related 2m dr 2 2m r2 to the energy transfer E by behaves as r 1'+ 1 for small r, and is to be normalized as (18)

For the transition of an electron from the nf subshell to the ionized state, one may write the atomic matrix element Notice that the same potential V(r) is used in Eq. (25) as in squared as Eq. (21); the use of the same V(r) not only simplifies the cal­ culation, but also guarantees its internal self-consistency. (2f'+l)l:x(2A+l) I ( · e'" e) I' 0 00 The contrast of the modern calculation with the hydro­ genie approximation is seen in the choice of the potential [ R (E,f' ,n,f,A,K) )2, (19) Y(r) . The hydrogenic approximation amounts to using where { is the Wigner 3j symbol, the sum over the ~ ~ i ) (27) index " runs from - I f- f' I to f+ f' in steps of 2, and R (E, f' ,n,f,A,K) is the radial matrix element defined by where Zeff and V O are two adjustable parameters. This poten­ tial V Hyct (r) satisfies neither of the two limiting forms, Eqs. R( E,f', n,f,A,K) P,r ,(r)h(Kr)Pn r (r)dr, (20) = J; (23) and (24). Therefore, the hydrogenic approximation gives no realistic behavior of radial wavefunctions for r = 0 or for h (Kr) being the sp herical Bessel function of the Ath order. r- oo, nor trustworthy results for properties depending upon the behavior of the wavefunctions at large or small r. An In Eq . (19), the radial matrix element is the only quantity example of such properties is the generalized oscillator that depends on the dynamics of atomic electron, all the strength near a threshold energy ( E = 0), which crucially other factors being geometric, i.e., dependent only upon depends upon the radial wavefunctions at larger r, as we shall angular-momentum quantum number s. fully document in Subsection 3.2. The most crucial part of the calculation is the determina­ Many properties of the realistic potential V(r) have been tion of the radial functions P's. The function P nr (r) for a extensively studied (Rau and Fano 1968); their consequences bound state with a discrete eigenenergy E111, < 0 sat isfie s the to radial wavefunctions P's have been elucidated in great radial Schrodinger equation detail (Fano et al. 1976, Manson 1976, Manson I 977, Man­ son 1978, Manson and Dill 1978), especia lly in connection with the optical oscillator strength spec tra, i.e., df(K,E) / dE 112 d2 P nr n2 f(f+ I) at the limit K - 0. Calculations by Manson (1972a , 1972b) - -- + [ Enr - Y(r) - -- - ] P nr = 0, (21) 2m dr 2 2mr 2 and their extensions (Manson and Inokuti, 1980) are based on extensive experience with work on the optical osci llator behaves as r r+ 1 for small r, and vanishes rapidly for larger so strength. Manson and Inokuti (1980) have calculated th e that it may be normalized as spectra of the generalized oscillator strengths for the ioni za­ tion of the K-shell and the L-shell of all atoms for Z ~ 30, (22) but have not published the results comprehensively. In the following section (Section 3.2), some of the resu lts will be discussed. In Eq. (21), V(r) is the potential of the field of force seen by Leap man et al. ( 1980) and Rez and Leapman (I 981) also the electron, and the field is due to all the other atomic elec­ reported similar calculations on the K-, L-, and M-shell gen­ trons and the nucleus. For a neutral atom of atomic number eralized oscillator strengths and related quantities for a selec­ Z, the general limiting behavior is tion of atoms, based on virtually the same method as that of Manson and Inokuti (1980). McGuire (1977, 1979) carried V(r) = -Ze 2 / r for r = 0, (23) out similar work as well. But his method contains an addi­ and tional mathematical approximation; he divides the full r-range (0 < r < ex,) into several interval s, in each of which the V(r) = -e 2 /r for r - ex,. (24) potential V(r) is approximated by a Coulomb potential 2 - Zie / r with a suitable effective charge number Zi. This In many calculations including Manson's (1972a, 1972b), allows one to write down the solution in that interval as a V(r) is determined through a version of self-consistent field linear combination of regular and irregular Coulomb func­ theories, called the Hartree-Slater method. tions and then to determine the coefficients of the linear The function P ,r (r) for a final, continuum state with combination by requirement of smooth connection of the energy E > 0 satisfies the same equation radial wavefunctions. This procedure may very well be more efficient than the straightforward numerical solution of Eqs. ( 17) and (21 ), done by Leap man et al. (I 980) and by Manson

8 Cross Sections for Inelastic Scattering and Inokuti (1980), but has a definite possibility of generat­ 6.0 ing spurious results. 5.0 Finally, calculations more accurate than the Hartree-Slater (a) E = 1.57 keV potential field method are indeed possible, for example, by 4.0 the use of the Hartree-Fock method, the random-phase ap­ proximation, or the method of configuration mixing . Yet, as ~ 3.0 ~ far as the properties of a deep inner-shell are concerned, ::.:: 2.0 more accura te calculations are unlikely to alter drastically the results of the Hartree-Slater calc ulation s. The reason for this _gtlO 1.0 expectation comes from the well-known notion of the per­ 0.0 turbation theory; the possible corrections to the Hartree­ -1.0 Slater calcu lation s must arise from the perturbative contribu­ tions from virtual excited states, but these states are located - 2.0 at very high excitation energies when one deals with a deep - 3.0 f inner-shell state. ---5 - 4 - 3 - 2 -1 0 log 0 3.2 Generalized Oscillator Strengths for Inner Shells of Atoms: Results 6.0 The most suitable way to show the data of the generalized 5.0 oscillator strength is the plot first given by Miller and Platz­ (b) E = 127 eV man (1957). By use of Eqs. (4), (7), (9), and (10), one can 4.0 readily see that it is suitab le to plot df(K,E) / dE at fixed E as 3.0 2 '½, a function of ln[(Kao) ]. Equivalently, one may plot ~ o-•l 11E(K) 2 as a function of In Q. Then, the area under the .__,.:..::: 2.0 curve represents the integrated cross section over a range of tlO 1.0 the momentum transfer nK (or over a range of the scattering _g angle). To show this point precisely, we may rewrite Eq. (7) 0.0 as - LO

- 2.0 R df(K,E) da E = ---- d [In (Kao) 2] . (28) - 3.0 mv2 E dE ---5 - 4 - 3 - 2 -1 0 log e For fixed incident electron speed v and fixed energy loss E, the momentum transfer is uniquely calculab le through kine­ 2 matics. Therefore, the Miller-Platzman plot is a graphical Fig. 6. The relation between 0 and (Kao) • representation of the angular distribution of inelastically --- Suppose that the incident electron has kinetic scattered electrons at fixed v and E . Yet, it takes some time energy T and momentum p, then T + mc 2 = [(cp) 2 and experience for anyone to become fully familiar with the + (mc 2 ) 2 J Y,. Suppose that the scattered electron has relation between 0 and (Kao )2. As an aid to this end, we pre­ kinetic energy T ' and momentum p', then T' + mc2 sent here Fig. 6, which shows the relation for energy-transfer ==[ (cp ' )2 + (mc 2 ) 2 ] v,. The energy loss Eis defined values corresponding to the K-shell threshold (E = 1.57 keV) by T ' = T - E. The squared momentum transfer is and to the 2s-subshell threshold (E = 127 eV) of aluminum . given by (Kb) 2 = p 2 + p ' 2 - 2pp ' cos0. These rela­ The figure illustrates several points. First, (Kao)2 varies over tions enable one to calculate (Kao)2 for a given set of a wide range with varying 0. Second, the range of the varia­ T, E, and 0. Here we show log,o (Kao)2 as a function 2 tion in (Kao ) becomes greater and greater with increasing in­ of log , 0, for fixed T and E. Figure 6a shows the rela­ 2 0 cident speed v. Third, (Kao) depends weakly on 0 for suf­ tion for E = 1.57 keV, corresponding to the 2 ficiently small 0, but becomes roughly proportional to 0 at threshold for K-shell ionization of aluminum. Figure large 0; the transition between the two kinds of dependence 6b shows the same relation for E = 127 eV, corres­ occurs at smaller and sma ller 0 with increasing v. ponding to the threshold for the 2s-subshell ioniza­ Several of the following figures -are examples of the Miller­ tion of aluminum. In each case, five curves refer to Platzman plot showing the results of calculations by Manson different kinetic energies T of the incident electron: and Inokuti (1980) for selected atoms. These figures also 10 keV (-- - ), 50 keV (------), 100 keV (----), show the corresponding results of the hydrogenic approxima­ 500 keV (- · - ·), and 1 MeV (------). For tion. Manson and Inokuti have actually calc ulated and plot­ the lowest T (10 keV), the curve for each E starts ted the genera lized oscillator-strength density df(K,E) / d highest at small 0 and ends lowest at large 0. For (E/ R) for the ionization from the ls, 2s, and 2p orbits of all higher and higher T, the curve covers wider and atoms through Zn (Z = 30). We shall respond to any reason­ wider ranges. Comparison of Fig. 6a and Fig. 6b able request for providing any of the numerical or graphical readily indicates that, for fixed T and for the same 0 data we have at hand. interval, the range of (Ka 0)2 values is wider for small­ er E (i.e., in Fig. 6b).

9 M. Inokuti and S.T. Manson

Figures 7-11 concern aluminum. Let us discuss each of Fig. 7. The generalized oscillator strength for the ionization them in turn. Figure 7 shows results for the ionization of the --- of the 2s-subshell of aluminum. 2s-subshell, which has the binding energy 8 = 127 eV ac­ The horizontal axis represents In I (Kao )2 I and the cording to Shirley et al. (1977). Each curve represents the vertical axis represents the density df(K,E) / d(E/ R) density df(K,E) I d(E / R) of the generalized oscillator of the generalized oscillator strength per unit range strength per unit range of E/ R = (f + 8) / R, where f is the of E/R = (f + B) / R, where f is the kinetic energy of kinetic energy of an ejected electron. Figure 7a shows results an ejected electron, and Bis the binding energy of the for the lowest f values, i.e., d R = 0, 0.5, 1.0, 1.5, and 2.0. shell, or the threshold for the ionization from that In particular, the solid curves represent Hartree-Slater shell. In this case, B = 127 eV, according to Shirley results, and the chained curves hydrogenic-approximation et al. (1977). Figure 7a shows values for d R = 0, results. In either case, the curve lying highest at smallest 0.5, 1.0, 1.5, 2.0. Figure 7b shows values for d R = (Kao )2 corresponds to d R = 0; the curve lying next highest 3, 4, 5, 6, 8. Figure 7c shows values for d R = 10, at sma llest (Kao) 2 corresponds to d R = 0.5, and so on. In 15, 20, 25, 30. The three sets of d R values are stan­ other words, the optical limit df(0,E) / d (E/ R) is mono­ dard and common to many of the figures to follow. tonically decreasing with f, as is always the case for the In all plots (including the figures to follow), the solid hydrogenic-approximation. curves represent Hartree-Slater results, and the Nevertheless, Fig. 7a illustrates a sharp difference of the chained curves hydrogenic-approximation results. In Hartree -Slater results from the hydrogenic-approximation Fig. 7a, the curve lying highest at the smallest 2 results. First of all, the magnitude at (Kao )2 - 0 is less than In I(Kao) ) values corresponds to f / R = 0, the curve half the hydrogenic-approximation value at d R = 0. Sec­ lying next highest at the smallest In ( (Kao )2) values to ond, the same magnitude stays virtually constant for the d R = 0.5, and so forth. range 0.5 ~ d R ~ 2.0, while the hydrogenic-approximation value decreases steadi ly with increasing d R. Finally, in the same f range, the Hartree-Slater results show the gradua l The peculiar behavior of the 2p-generalized oscillator development of the maximum at a high (Kao )2 value, i.e., the strengt h at low f and low K, in sharp disagreement from the emergence of the Bethe ridge. hydrogenic-approximation results, is attributable to the In Fig. 7b, one begins to observe the approach to the phenomenon of the delayed maximum in photoabsorption hydrogenic behavior. Here, the so lid curves represent the cross sections. Hartree-Slater results for dR = 3, 4, 5, 6, and 8, in the Detailed interpretation of the delayed maximum in the 2p order of decreasing height at low (Kao )2; the chained curves ionization of aluminum was given by Manson (I 972b). Brief­ represent the hydrogenic-approximation results for the same ly, this arises from the f-dependence of the ct-continuum final d R values also in the order of decreasi ng height. Through­ state, which is the dominant contributor to the gene rali zed out the f range of Fig. 7b, the Hartree-Slater results indicate oscillator strength. [See Fig. 3 of Manson (1972b)]. At f/ R the full development of the Bethe ridge . In contrast, the = 0 and small d R, the centrifu gal potential keeps the hydrogenic-approximation results begin to show the Bethe ct-continuum wave out of sma ll distances at which the initial ridge only belatedly with increasing f. 2p wavefunction has appreciable magnitudes, and therefore In Fig. 7c, one sees the virtua l agreement with the hydro­ the radial matrix element [Eq . (20)] must become small. At genic-approximat ion . The curves show results for d R = 10, higher d R, the ct-continuum wave begins to reach the sma ll 15, 20, 25, and 30 in the order of decreasing height; this ap­ distances and to attain appreciable overlap with the initial 2p plies to both the Hartree-Slater results (shown by the solid state. According to Eq. (25), the total effective potential curves) and the hydrogenic-approximation results. Figure 8 concerns the ionization of the 2p-subshell, which U(r) = V(r) + n2 f(f + l) / 2mr 2 (29) has the binding energy 8 = 8 I ev. Figure Sa shows results for the same set of the lowest f va lues, d R = 0, 0.5, 1.0, 1.5, with f = 2 determines the d wave . As Fig. 9 shows, U (r) gives and 2.0 . The chained curves are based on the hydrogenic­ rise to a shallow, but wide-ranged attractive region. As a approximation, and show the monotonic decrease of the result, properties of the d-wave change markedly between generalized oscillator strength with increasing L By sharp d R = 0 and d R = 2.0 (see Fig. 10). For example, the contrast, the Hartree-Slater val ues (shown by the solid phase o (with respect to the Coulomb wave) increases con­ curv es) are increasing with increasing f; the lowest solid curve siderably with L [See Fig. 3 of Manson (1969).) The change shows the Hartree-Slater value for d R = 0, which is less of ohere is only about 0. 7 radians, and the situation is differ­ than one-tenth the corresponding hydrogenic-approximation ent from a typical resonance, which is associated with a value. Figure Sb shows results for d R = 3, 4, 5, 6, and 8. change of o by almost 1r and implies the presence of a quasi­ Both the Hartree-Slater results (shown by the solid curves) bound state. Nevertheless, there is significant f-dependence and the hydrogenic-approximation results (shown by the of some d-wave properties, most notably the d-wave ampli­ chained curves) are decreasing with increasing f. Figure Sc tude, which is defined as follows. According to Eq. (25), the shows the close approach to the hydrogenic behavior, real­ continuum wavefunction P ,e (r) behaves near r = 0 as ized for d R = 10, 15, 20, 25, and 30; the highest curve cor­ responds to dR = 10, the next highest to d R = 15, and so (30) on. where C,e is a number, depending on f and e,to be deter­ mined so that the normalization relation, Eq. (26), is satis-

10 Cross Sections for Inelastic Scattering

5.0 .-r--,-,.....,...--,-,.....,...--,-,.....,...--,-,-,---,-..,.--,-~,-~--,-,..---,-r-n (a) Al 2s ------(a) Al 2p ------" ' 4 .0 -- ·------.."\ o2 --- ....__,\ '---- ..__,w ------'\'--: 3.0 ------"Cl " '----,....._ w ..__,~ 20 -"Cl LO

0.00 L...~...... L~~--'-~~..1...... ~_,_,~~--'-~~-'--'-_.,_-' - B - 4 -2 0 2 4 0) - 10 -8 - 4 - 2 0 2 4 ln (Ka/ In (Ka/

1.5 (b) Al 2s ------" (b) Al 2p -----..... ,....._ 2 0.09 ~ ' \ '---- ~ LO ------...... \ w .'- \ ~ ~ '--.. 0 .06 '------"-,\ ~ ~ - ...... '- ~ ..__,~ ~ -_-_-_-_-_--_-_-_-_------" ;;:;- 0.5 -"Cl -0 0.03

0 .00 0.0 - LO - 8 -8 - 4 -2 0 2 4 - 10 - 8 - 6 - 4 - 2 0 2 4 ln (Ka/ ln (Ka/

0.10 0.5

(c) Al 2s (c) Al 2p 0.06 0.4 2 2 '----w '----w 0.06 0.3 ~ ~ '-....,....._ :::::: ------C.:J w 0 .04 ~ 02 -.....,~ ------·- ____, "Cl -"Cl ------0 .02 0.1 ------·------0.00 0.0 - lO -8 - 6 - 4 -2 0 2 4 - 10 - 8 -8 -4 - 2 0 2 4 ln (Ka/ ln (Ka/ Fig. 8. The generalized oscillator strength for the ionization fied. The determination requires knowledge of the unnor­ of the 2p-subshell of aluminum. malized wavefunction over the entire r domain . Analysis The threshold energy B is 81 eV according to Shir­ shows that IC,e I2 is the major factor that determines the ley et al. (1977). The standard sets of d R values are t-dependence of the matrix element over a small f interval. It used for Fig. Sa, Fig. Sb, and Fig. Sc. Most of the should also be noted that IC,e 12 is virtually the same as the captions to Fig. 7 apply here. As an exception, in Fig. notion of the density of states, often used by solid-state Sa only, the Hartree-Slater results (shown by solid theorists . curves) are increasing with increasing t; in other Figure 11 shows results for the ionization of the K-shell (ls words, the lowest-lying solid curve corresponds to shell) of aluminum. For simplicity, we include in Fig. I la re­ d R = 0, the next lowest curve to d R = 0.5, and so sults for two values of ejected-electron energy, i.e., d R = 0 forth.

11 M . Inokuti and S.T. Man son

Fig. 9. The total effective potential for a d electron (f = 2) 12 emerging from the aluminum atom. (a) Al d effective potential The potential U is defined by Eq. (29) of the text, 1.0 and its value measured in R is here plotted against distance r measured in ao. Figure 9a shows the Har­ 0.8 tree-Slater potential over the wide range of r. Note a:: the bump around r / ao = 3.3. Even though the top 0.6 of the hump is below zero energy, the potential gives ~ rise to the marked difference in the behavior of the 0.4 radial functions for d R = 0 and d R = 2, shown in 02 Fig. 10. Figure 9b shows the Hartree-Slater potential (plotted as the solid curve) at small r. It also shows 0.0 the potential used in the hydrogenic approximation (plotted as the chained curve). Note the logarithmic --02 vertical scale. The two potentials are similar in the 0 2 4 8 10 12 small region r / ao < 0.3 (except at the close vicinity of the nucleus not shown here), but differ greatly for ®6 .0 large r; in particular, the hydrogenic potential stays positive throughout, and approach VO/ R = 6.82 5.0 [See Eq. (27).J (b) Al d effective potential 4.0 El R = 0.5. The hydrogenic-appro ximation result s (shown in o2 the chained curves) are decreasing with e: the upper curve '-._ 3.0 ::::> refers to d R = 0, and the lower one to El R = 0.5. The .__, 2.0 Hartree-Slater results (shown in the so lid curves) are opposite 2 t>() in the order: the lower curve refers to d R = 0, and the up­ _g 1.0 per one to d R = 0.5. Here again, the non-monotonic be­ ------·-- 0.0 havior in e is attributable to that of IC"' !2 • Before concludi ng this Sub section, we point out the gen­ - 1.0 erality of many observations we made above. First of all, the non-hydrogenic behavior of the generalized oscillator is seen - 2.0 0 0.4 in all atoms we studied, whenever the kinetic energy e of an 12 L6 eje cted electron is sma ll or comparable to the atomic poten­ tial in the relevant spatia l region . Second, the delayed max­ imum is common to many instances of p - d transitions, where the final d-wave is governed by the effective potential 2.0 U having a well-a nd-hump structure. Actually, the case of C/l ct-waves for aluminum is not the most clearcut. The ct-wave C: potential for chlorine, argon, or potassium shows a much g more pronounced hump, and the delayed maximum is much >Ju 1.0 C: more prominent (see Fig. 12). Third, the near-threshold ;:j structure of the kind we saw in the K-shell ionization of -(2) aluminum is common to most atoms (Manson and lnokuti > 0.0 ~ 1980, and Holland et al. 1978). To illustrate the common oc­ ): \ I cure nce of the non-hydrogenic behavior, Figs. 13-15 show \ / ·.___/ selected results. -~ -0 -LO We sho uld also note that there are many other ways for a::~ showing data than the Miller-Platzman plot (which is the most fundamental). For inst ance, one co uld show the gen­ era lized oscillator strength as a function of e, either at fixed K or at fixed 0. We may call the result a spectral plot. Figure 16 is an example. A spec tral plot of the differential cross section da E / dw at a fixed 0 is often called an electron energy loss Fig. IO. Radial wavefunctions for the 2p state (solid curve) spectrum. Often one sees in the literature an energy loss spec­ and the d-continuum states for aluminum. trum of a slightl y different kind, i.e ., The chained-dot curve (- · - ·) shows the d­ A continuum wave-function for d R = 0, and the .1a E = J~ (daE / dw) dw, (31) chained-dash curve ( - - - -) the same for El R = 2. The plotted values correspond to 1r - Y, P ,r (r) A plotted as a function of E, for a fixed aperture ang le 0. Ino- with f = 2 in the notation of the text. The solid kuti (1978) gave so me general remarks on thi s quantity . curve shows the normalized 2p bound-state. Notice Leap man et al. (1980) and Re z and Leapman ( 1981) have that the d-continuum wavefunction for d R = 2 presented exten sive results on this quantity of frequent has a much greater overlap with the 2p state than refe rence in electron microscopy. for d R = 0.

12 Cross Sections for Inelastic Scattering

0.025 ------=-- -""" 0.020 ~ ~ 0015 w (a) Al ts w " 0.015 " ~ --0 "w w 0.010 0.010 " (c) Al ls ~ ...... ,~.... --0 -0 0.006 0.000

0.000 0.000 - 10 - 8 ----6 - 4 - 2 0 2 4 - 10 -8 ----6 -4 - 2 0 2 4 ln (Ka,/ In (Ka/ 0.025 Fig. 11. The generalized oscillator strength for the ioniza- tion of the Is-shell (or K-shell) of aluminum. ------The threshold energy Bis 1.57 keV, according to 0020 ------Shirley et al. (1977). Most of the captions to Fig. 7 ~ apply here. An exception is Fig. 1la. In a deviation w " O.D15 from the standard set, we show here results for d R ~ (b) Al ls = 0 and e / R = 0.5 only. The Hartree-Slater results "w are shown by solid curves; the lower one corres- ~ 0.0!0 ponds to e / R = 0 and the upper one to e = 0.5 ...... ,..... -0 The hydrogenic-approximation results are shown by chained curves; the upper one corresponds to O.CXXi el R = 0, and the lower one to d R = 0.5. Another exception is Fig. llb, in which the upper curves 0.000 (both solid and chained) correspond to e / R = 3 - 10 ---8 - 6 - 4 -2 0 2 4 and the lower curves to el R = 8. In (Ka/ 12 ------LO

0.8 0.8 Cl d effective potential 06 0.4

::i:: 0.4 0.0 :::i " 02 ---04

0 0 ---0.8 - 10 - 8 -6 - 4 - 2 0 2 4 ln (Ka./

- 12 Fig. 13. The generalized oscillator strength for the 2p-sub­ 0 2 4 8 10 12 shell of chlorine. The threshold energy B is 210 eV according to Shirley et al. (1977). The standard set of the five lowest e values (d R = 0, 0.5, 1, 1.5, 2) is used. Fig. 12. The total effective potential for a d electron emerg­ The Hartree-Slater results (shown as solid curves) ing from the chlorine atom. are much lower than the hydrogenic-approximation The potential U is defined by Eq. (29) of the text, results (shown as the chained curves), owing to the and its value measured in R is plotted against dis­ effects of the effective potential for the d-continu­ tance r measured in ao. The hump around r/ ao = um final states (shown in Fig. 12). Indeed, the 2.5 stands out to positive energy, and causes a pro­ Hartree-Slater results are lowest for dR = 0, high­ minent delayed maximum for the 2p - d transi­ est for 0.5, and then decreasing for larger d R, tion. while the hydrogenic-approximation results are decreasing steadily with increasing e.

13 M. Inokut i and S.T. Man son om Fig. 14. The generalized oscillator strength for the 2s-sub ­ 006 shell of chlorine...... The threshold energy B is 278 eV accordin g to a:: 0.05 "--.. Shirley et al. (1977) . The standard set of the lowe st E (.:J values (d R = 0, 0.5 , 1, 1.5, 2) is used. The Har­ ~ 0.0-4 "--.. tree-Slater result s (shown as the solid curve s) ar e in­ crea sing with increa sing E, and indicate the emer g­ ~ 0.03 .__,,::ii.... ence of the Bethe ridge , while the hydrogenic-ap­ -0 0.02 Cl 2s proximation results (shown as the chained cur ves) indicate no sign of it. 001

0.00 0.40 - 8 - 8 - 4 - 2 0 - lO 2 4 ® ln (Ka/ 0.35 (a) Na 2s ...... \ a::: 0.30 I 0.40 "--.. I (.:J (Ka/ = 0.000 1 .__,, 0.25 \ 0.35 "'\ (a) Na 2s -0 "--.. \ ------\ 0.20 ...... 0.30 -..... \ £ \ a::: ::ii "--.. ------.."' \ \ .__,, 0.15 \ (.:J 0.25 \ .__,, "' \\ --0 '\ -0 ------'\ 0.10 "--.. 0.2() "- "' -..... ~ 0.06 ' :--.... .__,,::ii OJ 5 --- --0 0.00 0.10 0 5 10 15 20 25 30 c/ R 0.05

0.00 0.16 - 10 - 8 - 6 - 4 - 2 0 2 4 ,,.....____ ln (Ka/ \ \ (b) ~a 2s o? 0 12 \ ------"--.. ~ -- --, c...l (Ka/ = 4.00 \ ~ "--.. 0.06 0.12 ~ ...... ::ii --0 0.0-4 0.06

0.00 ~~~~~~~~~~~~~._..~~-L-~---J 0.0-4 0 5 15 20 25 30 (b) Na 2s t:/ R

0.00 Fig. 16. The generalized oscillator strength for the 2s-sub­ - 10 - 8 - 6 - 4 - 2 0 2 shell of sodium , plotted as a function of ejected ln (Ka/ electron energy at fixed momentum tran sfer. Basically the same data shown in Fig. 15 are plot­ Fig. 15. The generalized oscillator strength for the 2s-sub­ ted , but in a different way. The horizontal axis here shell of sodium. represent s d R, i.e. , the ejected-electron energ y The binding energy B is 71 eV according to Shir­ measured in R, at a fixed value of (Kao)2. In all ley et al. (1977). Figure 15a shows results for the cases, the solid curve shows the Hartree-Slater five lowest E / R values of the standard set. The Har­ results, and the chained curve the hydrogenic-ap­ 2 tree-Slater results (shown as solid curves) are de­ proximation result s. Figure 16a represents (Ka 0 ) = creasing with increasing E, while the hydrogenic ­ 10-•, i.e., the optical limit. Figure 16b represent s approximation results are decreasing with increas­ (Kao)2 = 4, showing an approach to the hydrogen­ ing E. Figure 15b shows the results for the five inter­ ic-approximation results and the emergence of the mediate E / R values of the standard set. Bethe ridge.

14 Cross Sections for Inelastic Scattering

3.3 Molecular and Solid-State Effects In solid-state contexts, effects analogous to the shape resonance are often called XANES (x-ray absorption near­ In many instances we have seen much evidence for the role edge structure), and have been the subject of many recent of atomic fields in governing the motion of an ejected elec­ studies. Examples of theoretical studies include Durham et tron, especially when its energy is low. For ·a , the al. (1981) and Durham et al. (1982). field of force seen by an ejected electron is in general non­ One sees another feature that distinguishes the molecule sphe rical because of the molecular geometry. This is so even from the atom. That is to say, there are undulations at higher though the electronic structure of deep inner shells is affected energies E ~ IO R in all the four symmetry classes. (However, only modestly by the molecular binding, as seen by the chem­ the undulations in the 71"u and 71"g symmetry classes are t?o ical shift s of core binding energies.[See Shirley et al. (1977) small to be seen in Fig. 17. ) These undulations were recogniz­ and Car lson (1975).] ed much earlier, and are known as EXAFS (extended x-ray Molecular effects on the optical oscillator strength, i.e., absorption fine structures).[See Teo and Joy (1981) for exam­ df(K,E) / dE at K - 0, have been recognized both experimen­ ple.) Briefly, the undulations result from constructive and tally and theoretically, and much of the understanding here destructive interference of ejected-electron waves with those should be pertinent to the generalized oscillator strength at scattered from different atoms. The variation with energy is finite momentum transfer. (However, there has been no ex­ tensive calculation specifically for molecule s.) roughly represented by sin(kD), where k = (d R) ½ /ao is the Figure 17 illustrates a dramatic example of molecular ef­ wave number of the ejected electron and D is the internuclear fects. This figure shows the photoionization cross section distance . Because of the origin, it is easy to see that the EX­ (the same as df(K,E) / dE at K - 0, apart from a univer sal AFS is in principle univer sal to all mole cules and solids, even consta nt) for the K-shell of molecular nitrog en, as calculated though the size of the undulations is often small and the pat­ by Dehmer and Dill (1976). The calculation is based upon a tern is more complicated for cases involving many and dif­ single-e lectron picture (like the calculation on atoms we saw ferent atoms. Indeed the notion of the EXAFS is so well­ in Subsections 3.1 and 3.2), and upon a potential that is mani­ known and prevailing in solid-state physics that Holland et festly non-spherical. Becau se of the molecular geometry, one al. (1978) adapted an EXAFS theory to interpret the near­ must distinguish four final -state classes designated by sym- threshold structure of ato mic K-shell spect ra . The shape resonance and the EXAFS are two well recog­ bols u u 11" 11" as opposed to the single class (the p state) g' LI' g' U' nized molecular effects. In reality , there are further effects for an atom. Whereas three of the classes show a smooth be­ cont ributing to the near-edge structure of inner-shell spectra havior for lower energies E of ejec ted electrons, the uu sym­ of molecules and solids. Ind eed, the electron energy loss metry gives rise to a sharp peak at about d R = 1.2, and spectrum obta ined experimenta lly by Wight et al. (1972173), causes a marked difference from the atom ic case. According shown as Fig. 18, indicates peaks additional to the shape­ to Dehmer and Dill (1976), the origin of the peak is a shape resonance peak. [Not e that the energy loss spectrum in the resonance, i.e. , the appearance of a quasi-bound state in the forward scatteri ng is roughly E·-' df(0,E) I dE, plotted as a molecular potential field. Roughly speaking, an electron in function of E and is thus distorted from the optical oscil­ the r; u state at that energy is temporarily trapped by the field lator-s trength spectrum. See Sec. 3.1 of lnokuti (1971).] before escaping out eventually. The shape resonance is a gen­ Some of the additional peaks are attributed to effects beyond eral occurrence for many molecules. Indeed, a review article the sing le-electron picture, e.g., simu ltaneou s excitation of by Dehmer and Dill (1979) shows many other examples. anot her electron along with the inner-shell ionization . In concl usion, we may reiterate that there has been no cal­ culation of the generalized oscillator strength of inner shells specifically including molecular or solid-state effects . Yet, the developments described above suggest that there is al­ ready enough groundwork for such a calculation, in both concepts and techniques and that the time may be ripe for a major undertaking.

10° Fig. 17. Photoionization cross section of the K-shell of :a- molecular nitrogen. ~ The figure is reproduced here with permission >- from Dehmer and Dill (1976). The horizontal axis 0 >- represents the kinetic energy of an ejected electron b measured in Rydbergs, and corresponds to d R in 10-1 the text of the present paper. The vertical axis represents the photoionization cross section uTOT 18 measured in Mb = 10- cm 2 , which is related to the density of the optical oscillator strength by r;TOT = 2 2 2 47r (e / nc) a0 df(O,E) / d(E / R) = 8.067 x 18 df(O,E) / d (E/ R) x 10- cm 2 • Each of the solid 10·2L- -- --~-----~ ------: 0.1 I 10 100 curves shows contributions from the indicated class PHOTOELECTRONENERGY(RYDBERG S) of final-state symmetry. The dashed curve shows twice the photoionization cross section of the K­ shell of atomic nitrogen.

15 M. lnokuti and S.T. Manson A atoms and molecules by use of keV electron scatter ing . Int. 1.0 J. Quantum Chem. Symp. No. 7, 377-394. V> Ca rlson TA . (1975). Photoelectron and Auger Spectroscopy, C: Plenum Pre ss, New York. ::, >,... Dehmer JL and Dill D. (1976). Molecular effects on the ...0 inner -shell photoabsorption. K-shell spectrum of N2. J . :c... Chem. Phys . 65, 5327-5334. ..e 0.5 Dehmer JL and Dill D. (1979) . The continuum multiple­ >- )'~ scattering approach to electron-molecule scatterin g and I- molecular photoionization, in: Electron-Molecule and CJ) z Photon-Molecule Collisions, T . Rescigno, V. McKoy, and B. w Schneider (eds.), Plenum Pre ss, New York , 225-265. I- z Durham PJ, Pendry JB, and Hodges CH. (1981). XANES: 0 Determination of bond angles and multi-atom co rrelati ons in 400 order and disordered systems . Solid State Commun. 38, 410 420 430 159-162. ENERGYLOSS (eV} Durham PJ, Pendry JB, and Hodges CH. (1982). Calcula­ Fig. 18. Electron energy-loss spectrum of molecular nitro­ tion of x-ray absorption near-edge struct ure. Comput. Phys. gen in the neighborhood of the K-shell threshold Comm., 25, 193-205. (409.9 eV). The figure is reproduced here with permission Fano U. (1956). Atomic theory of electro-ma gnetic int erac­ from Wight et al. (1972/73). The vertical axis tion s in den se materials . Phys. Rev. 103, 1202-1218. represents the intensity of electrons scattered into Fano U and Cooper JW. (1968). Spect ral distribution of the forward direction, the incident energy being 2.5 atom ic oscillator strength . Rev. Mod . Phys. 40, 441-507. ke V. The strongest peak (labeled as A) corresponds to the shape resonance discussed by Dehmer and Fano U, Theodosiou CE, and Dehmer JL. (1976). Electron­ Dill (1976). The other peaks (labeled as B, C, D, E, optical prop ert ies of atomic fields. Rev. Mod. Phys. 48, F, and G) are attributable to other phenomena that 49-68. are outside the scope of the present article. Fermi E. (1940). The ionization loss of energy in and ACKNOWLEDGEMENTS condensed materials. Phy s. Rev. 57, 485-493. Ferrell RA. ( 1957). Character istic energy loss of electrons We thank C.E . Brion, J.L. Dehmer, J. Geiger, and H.F. passing through metal foils: II. Dispersion relation and short Wellenstein for their generous permission and assistance for wavelength cutoff for plasma oscillations. Phys. Rev. 107, reproducing here figures or data from their publications. 450-462.

REFERENCES Geiger J . (1964). Streuung von 25 keV-Elektronen an Gasen. II. Streuung an Neon, Argon, Krypto n und Xenon. Z. Phys. Barias AD, Ruekner W, Wellenstein HF. (1977). High ener­ 177, 138-145. gy electron impact spectroscopy measurement on the Comp­ ton effect. Phil. Mag. 36, 201-207. Holland BW , Pendry JB, Pettifer RF, and Bordas J. (1978). Atomic origin of structure in EXAFS expe riment s. J. Phys. Bethe H . (1930). Zur Theorie des Durchgangs schneller Kor­ C. 11, 633-642 . puskularstrahlen durch Materie. Ann. Phy s. (Leipzig) 5, 325-400. Inokuti M. (1971) . Inelastic collisions of fast charged par­ ticles with atoms and molecules-The Bethe theory revisited. Bethe H. (1932). Bremsformel fi.ir Elektronen relativistischer Rev . Mod. Phys. 43, 297-347. Geschwindigkeiten. Phys. 76, 293-299. lnokuti M. (1978). Electron-scattering cross sect ion s per­ Bethe H . (1933). Quantenmechanik der Ein- und Zwei-Elek­ tinent to electron microscopy . Ultramicroscopy 3, 423-42 7. tronenprobleme. Chapter 3 in Handbuch der Phy sik, (H . Geiger and K. Scheel, eds.), Verlag von Julius Springer, Jnokuti M, Dehmer JL, Baer T, and Han son JD . (1981). Berlin, Vol. 24/1, 273-560. Oscillator-strength moments, stopping power s, and total inelastic-scattering cross sections of all atoms through stron­ Bohr N. (1948) . The penetration of atomic particles through tium . Phys . Rev. A. 23, 95-109 . matter. K. Dan . Videns. Selsk., Mat. Fys. Medd. 18, No. 8, Inokuti M, Itikawa Y, and Turner JE . (1978). Addenda: In­ 1-144. elastic collisions of fast charged particles with atoms and Bonham RA and Wellenstein HF . (1973). Measurements of molecules- The Bethe theory revisited [Rev. Mod. Phy s. 43, quantities related to the charge and momentum densities of 297 (1971)]. Rev. Mod. Phy s. 50, 23-35.

16 Cross Sections for Inelastic Scattering

Inokuti M, Kim Y.-K., and Platzman RL. (1967). Total cross Manson ST and Inokuti M. (1980). Near-threshold structure sections for inela stic scattering of charged particles by atoms in the atomic K-shell spectra for ionization by photons or and molecules. I. A sum rule for the Bethe cross sections and fast charged particles . J . Phys. 8 . 13, L323-L326 . its application to the helium atom. Phy s. Rev. 164, 55-61. Matsuzawa M, Mitsuoka, and Inokuti M. (1979) . Integral s Inokuti M and Turner JE. (1978). Mean excitation energies of the squared atomic form factor over the momentum trans­ for stopping power as derived from oscillator-strength dis­ fer. J . Phy s. 8. 12, 3033-3046. tributions, in: Proceedings of the Sixth Symposium on McCarthy IE and Weigold E. (1976). (e,2e) Spectroscopy. Microdosimetry, Brussels, 1978, J. Booz and H.G . Ebert Phys. Rep. 27, 275-371. (eds.), Harwood Academic Publishers, London, 675-687. McGuire EJ. (1977). Scaled electron ionization cross sections Khandelwal GS and Merzbacher E. (1966). Stopping power in the Born approximation. Phys. Rev. A. 16, 73-79. of M electrons. Phys. Rev. 144, 349-352. McGuire EJ. (1979). Scaled electron ionization cross sect ion s Lahman-Bennani A, Duguet A, and Wellenstein HF . (1979). in the Born approximations for atoms with 55 :':i Z :':i 102. Bethe surface and Co mpton profile for CO2 obtained by use Phys. Rev . A . 20, 445-456. of 35-keV incident electrons. Chem. Phys. Lett. 60, 405-410. Miller WF and Platzman RL. (1957). On the theory of the in­ Lahman-Bennani A, Duguet A, Wellenstein HF, and elastic scattering of electrons by helium atoms. Proc. Phys. Ronault M. (1980). Bethe surface and Compton profile of Soc. (London), 70, 299-303. NH 3 obtained by 35-keV electron impact. J. Chem. Phys . 72, 6398-6408. Nozieres P and Pines D. (I 959). Electron interaction in solids. Characteristic energy loss spectrum. Phys . Rev. 113, Landau LD and Lifshitz EM. (1960). Electrodynamics of 1254-1267. Continuous Media, translated by J .8. Sykes and J .S. Bell, Pergamon, London. See Chapter s IX and XII in particular. Powell JC. (1983). Inelastic scattering of electrons in so lids, in: Electron Beam Interactions, SEM, Inc ., AMF O'Hare, Lassettre EN, Skerbele A, and Dillon MA. (1969). General­ IL, 19-31. 1 1 ized oscillator strength for I S - 2 P transition of helium. Raether H. (1980). Excitation of Plasmons and Interband Theory of limiting oscillator strength. J. Chem. Phys. SO, Transitions by Electrons, Springer-Verlag, Berlin. 1829-1839. Rau ARP and Fano U. (1968). Atomic potential wells and Leapman RD, Rez P, and Mayers DF. (1980) . K-, L-, and the periodic table. Phys. Rev . 167, 7-10. M-shell generalized oscillator and strengths and ionization cross sections for fast electron collisions. J . Chem. Phy s. 72, Rez P . and Leapman RD . (1981). Core loss shape and cross 1232-1243. section calcu lations, in: Analytical Electron Microscopy- 1981, R.H . Geiss (ed .), San Franc isco Press, San Francisco, Manson ST. (1969). Dependence of the phase shift on energy 181-186. and atomic number for by atomic fields . Phys. Rev. 182, 97-103. Shiles E, Sasaki T, Inokuti M, and Smith DY . (1980). Self­ consistency and sum-rule tests in the Kramers-Kronig anal­ Manson ST. (1972a) . Theoretical study of generalized oscil­ ysis of optical data: Applications to aluminum. Phys. Rev. lator strengths in atoms: Comparison with experiment and 8. 22, 1612-1628. other calculations. Phys. Rev. A. S, 668-677. Shirley DA, Martin RL, Kowalczyk SP, Mcfeely FR, and Manson ST. (1972b). Inelastic collisions of fast charged par­ Ley L. ( 1977). Core-electron binding energies of the first ticles with atoms: Ionization of the aluminum L-shell. Phys. thirty elements. Phys. Rev. 8 . 15, 544-552. Rev. A. 6, 1013-1024. Starace AF. (1982). Theory of Atomic Photoionization, in : Manson ST. (I 976). Atomic photoelectron spectroscopy, Handbuch der Physic, S. Fli.igge (ed.), Springer, Berlin, Vol. Part I, in: Advances in Electronics and Electron Physics, L. 31, 1-121. Marton (ed.), Academic Press, New York, Vol. 41, 73-111. Teo BK and Joy C. (eds.) (1981). EXAFS Spectroscopy, Manson ST. ( 1977). Atomic photoelectron spectroscopy, Plenum Press, New York. Part II, in: Advances in Electronics and Electron Physics, L. Marton (ed.), Academic Press, New York, Vol. 44, 1-32. Walske MC (I 952). The stopp ing power of K-electrons. Phys. Rev. 88, 1283-1289. Manson ST. (1978). The calculation of photoionization cross sections; An atomic view, in: Topics in Applied Physics, Vol. Walske MC. (1956). Stopping power of L-electrons. Phys . 26, Photoemission in Solids, I. General Principles, M. Car­ Rev. 101, 940-944. dona and L. Ley (eds.), Springer-Verlag, Berlin, 135-163. Wight GR, Brion CE, and van der Wiel MJ. (1972/73). K­ Manson ST and Dill D. (I 978) . The photoionization of shell energy loss spectra of 2.5 keV electrons in N2 and CO. atoms: Cross sections and photoelectron angular distribu­ J . Electron Spectrosc. Rel. Phenom. 1, 457-469. tions, in: Electron Spectroscopy, Th eory, Techniques, and Wong TC, Lee JS, Wellenstein HF and Bonham RA. (1975) . Applications, C.R. Brundle and A.O. Baker (eds.), Academ­ Experimental definition of valence shell and cumulative-shell ic Press, New York, Vol. 2, 157-195. Compton profiles from 25-keV electron-impact studies on N2, Ne, and Ar. Phy s. Rev . A. 12, 1846-1858 .

17