PHYSICAL REVIE% A
VOLUME 49, NUMBER 5
Inversion of electron-water
elastic-scattering
data
A. Lun, Xue Jun Chen, * L. 3. Allen, and K. Amos
School of Physics, University of Melbourne, Parkville, Victoria 8052, Australia
(Received 22 October 1993)
Fixed-energy inverse scattering theory has been used to analyze the differential cross sections for
the elastic scattering of electrons from water molecules. Both semiclassical (WKB) aud fully quautal
inversion methods have been used with data taken in the energy range 100—1000 ev. Constrained
to be real, the local inversion potentials are found to be energy dependent, a dependence that
can be interpreted as the local equivalence of true nonlocality in the actual interaction. Further
improvement in fits to the data was found by allowing the interactions to be complex, reHecting the
role of coupling of the elastic to nonelastic channels.
PACS number(s):
I.
34.80.Bm, 03.65.Nk
INTRODUCTION
Direct procedures are the most common ones used to
analyze (fixed energy elastic) scattering data that one
obtains from beam experiments, whether those experiments involve nuclear, atomic, or molecular systems.
In the main those direct procedures are purely phenomenological with a parametric form chosen a priori
to be the (central, local) interaction between the colliding entities. Increasingly, however, those interactions (or
at least the real parts of them) have been defined by
folding some underlying pairwise microscopic interaction
with the density distribution of the quantal system involved. Whichever approach to the direct procedure is
used, there is a set of parameters that identify the scheme
and, invariably, their values are adjusted to give a best fit
to the measured data. That best fit is speci6ed, usually,
by finding a minimuxn chisquare (y2) fit to the data from
variations in the parameter space.
Inverse scattering methods [1] form an alternative procedural class with which to analyze the same data. %ith
inverse methods, the interaction between the colliding
pair is extracted from the data without a priori assumptions about the shape of the potential, although it may
belong to a certain class of potentials and the validity of
the dynamical equation of motion (the Schrodinger equation) is assumed. But the results are linked to the specific
method used and there is always a question of uniqueness.
However, in applications to date [2—5], the quality of fit
one can obtain with inverse scattering theory to (quality)
data is often such that the sensitivity of the potentials
obtained by inversion can be measured with respect to
the range and amount of input data. Those potentials
are specified hereafter as inversion potentials.
Of all of the methods for inversion of fixed energy
(cross-section) scattering data, those based upon a rational function representation of the underlying scatter-
Permanent address: Physics Department, Tsinghua
versity, Beijing, The People s Republic of China.
1050-2947/94/49(5)/3788(11)/$06. 00
Uni-
49
ing function, Si, (A), arguably are the most useful. With
such forms for the S function, solution of the inverse
problem with the Schrodinger equation is facilitated either by a semiclassical, WKB, procedure [6] (under conditions appropriate for use of that approximation) or by
a fully quantal scheme of the Lipperheide-Fiedeldey (LF)
type [7], methods which have been used extensively in recent years to analyze the elastic scattering cross sections
from the scattering of two nuclei [5,8]. The attendant fits
to measured data in those cases were usually an order of
magnitude better than any obtained by direct methods
of analysis. Semiclassical methods of inversion have also
been used with considerable success in studies of atomatom scattering [2]. In his review [2], Buck also noted the
possible use of rational function representations of the S
function. Likewise with an excellent fit to very accurate
data, electron-helium atom scattering has been inverted
to give the representative interaction [4). The quality of
the data in that case was such that an error analysis was
feasible and from which confidence limits at each radius
could be placed upon the extracted interaction.
In recent years, difI'erential cross sections for electron
scattering from small molecules have been measured and
the scattering viewed as a central field problem. Direct
methods have been used to analyze the data, so far with
varying success. For water molecules, Katase et al. [9]
form of the
have used both a purely phenomenological
(real) interaction potential (the sum of two Yukawa potentials) and a spherically averaged folded one in direct
solutions of the Schrodinger equations. Those analyses
gave quite good fits to the data taken with electron eIIergies in the range 100—1000 eV. Herein, we report the
results of new analyses of that data made using inverse
scattering theory. In the first instance, to ensure that
the scattering potentials so obtained are purely real, a
constraint is required that ensures the poles and zeros
of the required rational function representation of the S
functions are complex conjugate pairs. Excellent fits to
the data result, but the inversion potentials are then energy dependent. Energy dependence may be anticipated,
however, as such refIects nonlocal efIects in the actual
scattering processes. An estimate of potential scattering
3788
1994
The American Physical Society
INVERSION OF ELECTRON-WATER ELASTIC-SCATTERING DATA
49
nonlocality has been made assuming that it has FrahnLemmer form.
The potentials used in direct methods of analysis of
the scattering are purely real, as are those we obtained
with our 6rst study of inversion of the same data. Consequently &om all of those studies, the scattering phase
shifts are also purely real, the modulus of the S function
is then fixed to be unity and there is no reaction (absorption) cross section in the process. For energies in the hundreds of eV, it seems probable that some Bux of the beam
will be "lost" with the occurrence of nonelastic reaction
events. A complex central potential would be needed
then to describe the elastic scattering process as independent of any reaction channel couplings. Allowing the
pole-zero pairs description of the rational function form
of the S function required in the inversion procedures to
be unconstrained leads to such complex potentials and,
attendantly, a significant absorption cross section.
Following a brief review of the inverse scattering theory
and of the WKB and LF methods that we have used in
our analysis, the results of the calculations are discussed
in Sec. III.
ceed with inversion of the Schrodinger equation for fixed
energy scattering, one must interpolate and extrapolate
upon a discrete set of S-function values; however they
are obtained by fits to measured data.
There are several methods of solution of fixed energy
inverse scattering problems.
Herein we will consider
the applications of just two. They are the semiclassical (WKB) method [6] and a fully quantal method based
scheme [7,10,11].
upon a Lipperheide-Fiedeldey
A. The semiclassical (WEB) method
In this approach, with ro being the classical turning radius and K~(r) being the local momentum through the
interaction region, scattering phase shifts defined by
bg(t)
= (l+
SCATTERING THEORY
[Ki(r') —k] dr'
——kro+
-')
2
2
are used to specify the "classical" deQection function,
8(A)
II. ELEMENTS OF FIXED ENERGY INVERSE
3789
=2
—
d
dA
hg(A)
from which, via an Abel integral transformation,
find the quasipotential
one can
O(A)
Solutions of the Schrodinger equation with a central
local interaction describing the collision of two quantal
systems link to measured data via scattering amplitudes
that one extracts from the asymptotic forms of those solutions. In the partial wave expansion treatment of the
scattering, those scattering amplitudes, in turn, can be
de6ned in terms of S functions or in terms of phase shift
functions. In the center of mass kame, the scattering
amplitudes de6ne difFerential cross sections by
A2
4E1
vr
d
odo
—o.2
(
8(A)
I, o
v/2 —o2
The scattering potential is determined from that quasipotential via the Sabatier transformation by
VwKn(r)
=E
1
—e(
so long as there is a 1:1 correspondence
&om the transcendental equation
r ——ere
and are expressed in a partial wave expansion as
k
(2)
Therein S(A), with A being the angular momentum variable, are the S functions which relate to the phase shift
functions by
S(A)
= exp[2ib(A)]
.
The inverse scattering problem for fixed energy scattering then resolves to the following: Given the S function (equivalently the partial wave scattering amplitudes)
at a particular energy and as a function of the angular
momentum (A), find the central local potential which reproduces that S function.
To do so, however, the S function must be defined at
all (continuous) values of the angular momentum variable. But measured data are only sensitive to that S
function at the integer values of A —2. Thus to pro-
between
r and
cr
r-
This condition is valid if, for the actual potential,
E & V(r)+
1 dV
2
dr'
E,
a condition that E exceeds
b;& (the energy at which
"orbiting" occurs). Also, as the definition of the quasipotential has the limit
Q(o)
~
oo,
(10)
the transforms lead to r
ro and V —+ E in that limit.
The WKB method for basically attractive interactions is
valid only for radii in excess of the classical turning value
and/or the condition for orbiting.
The integral form of the quasipotential is solved easily if one has a rational function representation of S(A),
namely,
A. LUN, XUE JUN CHEN, L. J. ALLEN, AND K. AMOS
3790
the associated iterative scheme mentioned in
brief next, .
schemes are particularly
The Lipperheide-I'iedeldey
useful ways to go about determining that potential. In
the simplest of those schemes, one assumes that the fixed
energy S function for scattering can be represented by a
complex rational function form
generalize
as then one finds
(12)
The residual problem for use of the &KB procedure is
then to find the set of complex zero-pole pairs fn„, P„)
that "fit" S(A). One may also incorporate a reference
S function whenever scattering is dominated by aspects
not of primary interest. Coulomb scattering between two
nuclei is an example. The rational function form of the S
function ensures that the &KB potentials decrease as —,
for very large radii. Details are given in the Appendix.
B. Fully
quantal inversion
S(A)
The total scattering potential
be obtained by iteration as
y((kr) +
f
~
k
—l(t+ 1) ) y((kr) = U(r)y((kr) .
2
~
where, with Vp = 0, the increment function for each additional pole-zero pair of the N set defining the S function
is given in terms of the Jost solutions kom the preceding
iterate of the potential by
2i
r
Therein,
Iz (r)
I(+)(
From this form, the scattering potential U(r) can be defined in terms of the free Jost solutions that satisfy
——
A2
and have asymptotic
f (+) (
f(+)(„)
r2
dr2
()
(14)
properties
~ikv
) r —+oo
(15)
f
wherein
are free
by inversion then
pairs of (complex)
tering function as
Jost functions. The potential obtained
is given by the sum over % pole-zero
angular momenta that define the scatper Eq. (11) and has the form
rdr
where the function
(r
v.
X~I
and the Wronskian
is a solution of
P K
P
=
T
function x is defined by
(18)
This development can be extended [11] to include reference potentials in the basic Ricatti equations and to
[V(r) = Vjy(r)] can then
(20)
The objective is to invert the scattering data to define
the potential in the radial Schrodinger equation,
d2
=
dr
(1(
are logarithmic
)
~
)(r) + I(+)(r)
derivatives,
(eg~fp(+) (r)„)
f (+)(
(22)
)
with f&(+) (r) being the Jost solutions of the potential
[V„ i(r)] that asymptote as e+'"", respectively. Again,
one can use a reference S function in the procedure. The
reference potential associated with that S function then
specifies Vp. As with the &KB scheme, the fully quantal
one leads to potentials that decrease as —, at very large
radii [7]. That behavior, and the behavior of the quantal
inverse potentials as r m 0, is discussed in the Appendix
as well.
III. RESUITS
AND DISCUSSION
The 200—1000 eV data are fitted extremely well when
rational forms for the S function with two pole-zero pairs
of complex (and conjugate) angular momentum values
are used. The specific values of those poles and zeros are listed in Table I. All are given to eight decimal
places stressing the importance of many digit accuracy
for inverse scattering calculations. Indeed, in the (fully
quantal) study of the extensive data set from 350 MeV
0- 0 heavy-ion scattering [8], eight digit acnuclear
curacy was essential to give the extremely good fit to
the elastic cross-section data. At each energy for this
electron-molecule scattering, the parameter values were
defined by a y minimization fit to the actual data. With
minor variations, the resulting parameter values show a
monotonic variation with incident energy. Essentially,
the first pole o. i moves further away from both the real
and imaginary (A plane) axes. The pole n2, on the other
hand, moves slightly towards the real angular momentum
VERSgQN Q F EI,ECTRQ N-WA TER EI-ASTCC-SCATT
49
S functions s that
data an
are
h
real
1000
700
500
400
300
real
imaginary
imaginary
e5 3.285 75604 0.53465170 0.76808099
916014 0.78895677
4 3339 0 .62 915601 0.78895677
0 2.8164
4
628877 0.71376029
69
1.416
0.67646893
1.24892894 2
~
200
eC
200
(quantal)
200
(WEB)
2.47321338
8 3..36016047 1.
1 00000000 0.69847163
2.54989600 2.91580747
7 0.81507489
t betw~~n the
.
-
h the
equatio»»]ues
Energy
(eV)
3791
are shown a t
easur d cross se
calculation
Speclfi
y
er
d
pootentia s
version
DATA
yield those
in-
of
s
They are -nn excel en
.
values of the relevant
functions o
of ratlonaal form
.
tliat were
t to the j.nv er sion
e when
taann at the P ropriate (t + 2 values of the angular
momentum variable.
Th e real
. re local pootentials tha t we have o btajned &o
lnvel sionnofthe 200 500 eV data +om lectron scatte
tering
. 1- 2 . They
ev
e were evaluated
ol--zero pair
d th
~
section
f»g-
~
11
g
o
'
it movess slightly
s
away
dependence re
nd at the same time
tim
. That e
h
t
c
'
eV,
dh
1
d in
'
.
.
fo the 200
1 are energy
dependence is
ner
ta
the top sec-
are
tials can be
tion of that figure the cro
1
}i
10
ya
se equlva lent (ene gy
o1U 1o
nt Sc o
functional trans or
E,
10
10
2p
being
t hee local
onloca i y
interac
t ) loca o
+ Wioc(r, E
E) —E
vj(r,
E) = 0
ma
n
(23)
form and
b4
10
-2
10
30
0
60
(degj
~
.5
~
---.
100
~
200
o---~ XO
500
1000
=
V)
C
U
O
0
i
a
eV
eV
—50
eV
0eV
0
FI
I
2
de
p
displayed are shown
t—
6(~
I
ntial
o s
s for elas t iee electron
diverse energ'
Th ( l)
0.2
0.3
0.4
Radius (a. u. j
I
of the 200 —500 e V sc
otentialsIs obtained
o
bby WKB inversion
A. LUN, XUE JUN CHEN, L. J. ALLEN, AND K. AMOS
3792
h2
2p
v~ y v(r) —z
z} =
p(r,
f J(r, r')p(r', z}dr
(24)
being the nonlocal one, phase equivalence only requires
that the two solutions equate asymptotically. It is useful
to consider a Frahn-Lemmer type of nonlocal interaction,
vlZ. )
J(r, r') = F(B) v(p):- F(r) v(p),
(25)
R and p are 2{r+r'{ and {r —r'{, respectively, and
the range of the nonlocality is small. Then a Taylor series
expansion allows the nonlocal Schrodinger equation to be
mapped into the local form provided
where
Q(r, E) = T(r)
((()(r,
E),
(26)
where
dWi„(r, E)
dE
Note that the wave functions need
cally (phase equivalence) and that
any energy independent nonlocal
ther, taking the nonlocality to be
v(p)
= (~vrcr)
exp
identifies [13] the Frahn-Lemmer
2h2
2
double Yukawa function is shown by the dashed curve.
That double Yukawa form
—7.00'
—1.87m
—+ 0. 18
F d, i(r) = —0.15
(30)
is a good representation for radii in excess of about 0.2
a. u. , below which, however, the &KB process is less reliable. But, it is also the case that the data are rather
insensitive to the actual properties of the potential in
that radial region.
The local potentials obtained by inversion are essentially independent of energy for energies in excess of 500
eV. Thus the term V (r) in the nonlocal Schrodinger equation is specified. Also the 1000 eV data not only are well
fitted by the inversion scheme but also we have found
thereby, a real potential that is very much like the double
Yukawa form used in the original study [9]. The comparison is shown in the bottom section of Fig. 4. The inversion potential is displayed therein by the solid curve while
the phenomenological one is shown by the dashed curve.
There is very little to distinguish between them. But the
only match asymptotithis relation is true for
interaction [12]. Furof Gaussian form
p2
(——
function
dE
10
(28)
~
de„(r) E)
po2
(27)
49
F as
(29)
Shown in Fig. 3 is the nonlocal Frahn-Lemmer function we have obtained &om the set of inversion potentials
at 350 eV. The result, specified by using a four-point Lagrange derivative formula with the 200, 300, 400, and
500 eV potentials, is shown as the continuous curve in
the diagram and the units are 2h", . For comparison a
1.5
1.0
0.5
ersion
ble Yukawa
—10
—02—
—15
—0.4—
0.5
1.0
1.5
eV elec1000
FIG. 4. The results of analyses of the
0.5
Radius
{a.u. )
FIG. 3. The Frahn-Lemmer nonlocality function deduced
from the 200—500 eV inversion potentials.
The derivatives
involved were estimated at 350 eV.
tron-water scattering cross section. The calculated cross sections were made using the WKB inversion potential (solid
behavior at large
curve) and with modifications to have
radii. The (real) phase shifts that each potential yields are
shown in the middle section and the inversion potential is
compared with a double Yukawa model form in the bottom
segment.
~
INVERSION OF ELECTRON-WATER ELASTIC-SCATTERING DATA
49
two interactions do have quite diferent asymptotic behavior. The inversion potential eventually decreases as
—,. But the data are not sensitive to the asymptotic form
of these interactions. This is evident &om the top and
middle sections of the figure. The calculated cross sections are compared with the data in the top section and
those calculations were made using the inversion potential as a base but with its variation from a cutofF radius
R,„q altered to have 4 character. The solid curve is the
unaltered result R,„q
oo, the long dashed curve is the
result found using R,„q —1.5 a.u. , and the short dashed
curve was found when R, „& —1.0 a.u. Clearly the resultant cross sections are indistinguishable except at the
extreme forward scattering angles. The associated phase
shift values are shown by (continuous) curves in the middle section. All three cases dier at each integer l value
by but a few percent &om each other.
The inversion potentials also give total elastic and momentum transfer cross sections for this scattering. The
values at each energy are listed in Table II in units of
10
cm and compare closely with those evaluated by
Katase et al [9]. .The energy variations of both calculated results are very similar to the values obtained by
inversion: slightly larger for the elastic cross sections and
slightly smaller for the momentum transfer ones.
All of the results so far presented were obtained by
using WKB inversion. At low energies that scheme may
be less accurate, in which case a fully quantal inversion
scheme should be considered. As a first example, we analyzed the 200 eV electron scattering cross sections &om
water and to demonstrate that the procedure is not specific to a single molecule, &om methane as well. The
data &om the latter reaction were also more numerous.
The two conjugate pole-zero pair values of the fitted S
function used in the fully quantal scheme and for the
scattering &om methane are listed at the bottom of Table I. Note that we restricted the search to have poles
with Re(A)
1.0 to ensure that the low radial behavior of the inversion potential did not oscillate. Details
are given in the Appendix. Using the resultant potential in the Schrodinger equation, and so determining the
scattering cross section, gave results that are in excellent agreement with the observations (as were those for
scattering of 200 eV electrons from water) and that is displayed in the top section of Fig. 5. In this case we also
made a WKB inversion and found, using this potential, a
similarly good representation of the data. The pole-zero
o~ 100
N
b 0
—
10
~
)
TABLE II. The elastic and momentum transfer cross sections (units 10
cm ) for electron scattering from H20.
~el
= &total
&mom
Energy
Katase
This
Katase
This
(eV)
et al.
0.548
work
et al.
work
0.0515
0.0930
0.156
0.208
0.296
0.464
0.0504
0.0924
0.1540
0.2044
0.2828
0.4201
700
500
400
300
0.819
0.608
0.893
1.04
1.32
1.56
1.103
1.437
1.649
2QO
2.11
2.229
100Q
3793
~
T 'T
30
60
0
90
.
I
4 T
120
(deg)
-10
Radius (a. u. )
PIG. 5. The cross section for 200 eV electrons elastically
scattered from CH4 compared with the result found with the
fully quantal (real) inversion potential. That potential is compared with the &KB result in the bottom segment of this
figure.
pair values of the rational S function for this calculation
were not restricted and are also given in the bottom section of Table I. The fully quantal and WKB inversion
potentials are shown by the solid and dashed curves in
the bottom segment of this diagram. They are also very
similar and indicate that the simpler to use WKB inversion scheme can be entertained, at least to this energy,
for electron scattering &om small simple molecules.
The 100 eV data &om water cannot be as well represented by either WKB or fully quantal inversion calculations so long as the constraint upon the pole-zero pairs to
be complex conjugate (implying that the potential is real)
with the restriction to having just two pole-zero pairs is
maintained. In any event, there remains the question of
Hux loss. Nonelastic events will occur and unless each reaction channel is considered along with the elastic one in
a complete coupled channels approach, the events leaving
the target in an excited configuration act as a source of
Hux loss &om the beam. In such a case, the local interaction with which elastic scattering is described by a mean
field theory becomes complex, the imaginary component
being absorptive in character. As a consequence, an absorption cross section exists and the total scattering and
elastic scattering cross sections no longer coincide.
Both inversion schemes considered herein can be used
to obtain complex interactions. This occurs by using rational form S functions in which the conjugate constraint
A. LUN, XUE JUN CHEN, L. J. ALLEN, AND K. AMOS
3794
TABLE III. The poles and zeros of the rational form S functions that fit cross-section data from
electron scattering from H20, and which are associated with complex (inversion) potentials.
Energy
real
(eV)
100
1.20712235
3.02150092
1
2
1
2
500
imaginary
0.97212999
0.33417144
1.33001364
0.39985415
2.53465613
2. 16370124
upon the pole-zero pairs has been removed &om the data
fitting process. By so doing, two pole-zero pair S functions again give excellent fits to data. We consider the
100 and 500 eV cross sections in this way. The fitted
parameter values in those cases are shown in Table III.
They were obtained by starting with random values and
allowing a y minimization search to be unconstrained.
We did not start with the best conjugate pair values. The
fits that these &ee fitted S functions give to the data are
shown in the upper portions of the two components of
Fig. 6. Shown below them, the data are compared with
the results found &om calculations made with diverse
potentials. The solid curves are the results obtained by
using the (complex) potentials determined with the fully
quantal inversion method. The other two results were
obtained by using the model potentials [9j, the double
Yukawa form, and the folded density one being used to
real
imaginary
2.80217431
1.03890829
3.12388397
0.18162345
-0.263431881
-0.38222030
-0.47514827
-0.98578740
give the results displayed by the long dashed and short
dashed curves, respectively. The differences are marked,
as is evident &om the comparisons given in Fig. 7. The
real part of the (complex) inversion potentials is stronger
than the model forms, especially at 100 eV. That is so as
the refractive processes must compensate for absorptive
processes to yield the same elastic scattering. However,
the absorption is still quite weak, far weaker than one
finds for other scattering problems such as nuclear scattering, for example. The potentials have been truncated
at small radii as well. The data are just not sensitive
to the specifics of the potentials below about 0. 1 a.u.
As a consequence, the WKB scheme has utility since its
use leads to the (complex) potentials shown in Fig. 8.
They concur with the fully quantal scheme results and
are shown to the small radial limit allowed.
Finally, the complex interactions lead to absorption
cross sections and in the cases studied, those values are
shown in Table IV. The total elastic, absorption, total reaction, and momentum transfer cross-section values are
10
100 MeV
10
0
10
10
CV
0+
b( c".
~i~
10
100 eV
10
500 MeV
0
10
10
—5 —
I
I
60
120
0
180
I
(deg)
FIG. 6. The cross sections for 500 eV (top) and 100 eV
(bottom) electrons scattered off water compared with the rational S function fits, with the results of using the fully quantal (complex) inversion potentials (solid curve) and with the
two model (real) interactions of Katase et al. [9].
2
Radius Ia. u. )
FIG. 7. The complex potentials found by quantal inversion
of the data shown in Fig. 6 with the solid and long dashed
curves, giving the real and imaginary parts, respectively.
real model interactions are shown also.
The
49
INVERSION OF ELECTRON-WATER ELASTIC-SCATTERING DATA
3795
Those same methods have also been used to ascertain
the potential for the elastic scattering of 200 eV electrons
&om CH4.
The starting point for each of the inverse scattering
calculations was a rational function representation of the
S function. The pole-zero (complex) pairs of angular momenta were determined by fitting the differential crosssection data. Under the constraint that the pole-zero
pairs all be complex conjugates, and with a two-pair set,
excellent fits to the data were found. The inversion potentials that result then were purely real functions. At
the highest energies, those potentials and the associated
total elastic and momentum transfer cross sections resemble the results found with model forms of the interaction used in direct solution of the Schrodinger equation
[9l
—10—
Radius (a.u. j
FIG. 8. The complex potentials for 100 and 500 eV e-Hqo
elastic scattering as de6ned by WEB inversion. The real and
imaginary parts are shown by the solid and dashed curves
respectively.
listed therein and are compared with the elastic and momentum transfer values specified by Katase et aL [9]. The
100 eV results are the most aH'ected although the 500 eV
momentum transfer cross section reveals the largest percentage variation.
We do not claim that the complex interactions we have
found are the appropriate ones for electron scattering
&om water. We have shown that the data can be well
fitted by such interactions and so stress that other physical information must be used to settle upon what is the
actual representative interaction for this scattering.
IV. CONCLUSIONS
Fixed energy inverse scattering methods have been applied to extract electron-water molecule potentials &om
measured diBerential cross sections. Both semiclassical
(WKB) and fully quantal inversion schemes have been
used to ascertain those real local interactions from data
taken with incident energies in the range 200—1000 eV.
TABLE IV. Integrated cross sections (units 10
to the complex interactions.
corresponding
Reference
The derived potentials are smooth, well-behaved functions but they are energy dependent, especially for energies below 500 eV. That energy dependence can be interpreted as the eKect of nonlocality in the interaction
and, using a Frahn-Lemmer form, we find that the nonlocal function based upon an energy of 350 MeV is also
smooth and relatively weak. It closely resembles a sum
of two Yukawa functions.
We also analyzed the 100 and 500 eV electron-water
scattering data allowing the search for the rational poleGood fits to the
zero pair values to be unconstrained.
data in those cases were found but the inversion potentials then became complex. A complex nature to the
scattering potential is consistent with Aux loss in the experiment and leads to absorption cross sections as well.
The total elastic and momentum transfer cross sections
then vary markedly &om the predictions of Katase et al.
[8
Much more data are required to resolve questions that
these analyses raise. Notably there is the question of just
what nonlocality is pertinent in electron-molecule scattering. Clearly such eKects arise through exchange amplitudes in microscopic model calculations. Then there is
the problem of just how strong absorption efkcts should
be. Neglect of channel coupling and the attendant scattering cross sections for inelastic events is a reason to
believe that the purely elastic channel models should use
complex interactions. Finally we observed that the current data set was not sensitive to the precise polarization
potential of the molecules. Altering the potentials from
quite small radii to have an „, behavior (instead of the
variation of the inversion potential) made very little
change to the fit to most of the data.
—
cm )
&mom
ACED O%V LEDC MENTS
100 eV
Katase et aL
This work
2.98
3.30
2.14
2.98
5.44
1.01
2.98
1.04
1.81
0.156
0.85
500 eV
Katase et aL
This work
1.04
1.12
0.69
We are grateful for the financial support given this
project by the Department of Industry Trade and Commerce under the Australia —People's Republic of China
scientific agreement and for research grants made by the
ARC and the University of Melbourne.
A. LUN, XUE JUN CHEN, L. J. ALLEN, AND K. AMOS
3796
APPENDIX: ASYMPTOTIC BEHAVIOR
OF INVERSION POTENTIALS
=
t')2
p2)
—n2n )
n=1 (A2
(p) are Jost functions which relate to
(2)
and
H& (p), Hankel functions of the first and
Hz (p)
second kind, respectively, by
fz
(1)
Both the semiclassical (WEB) and fully quantal (I I')
inversion schemes we have used to obtain the potentials
described herein take as input the rational form of the S
function,
Si, (A)
(+)
Therein
f,'"(p) =
(Al)
'(p)
f&
with the result that both schemes give potentials
asymptotically vary as „,.
—
=
2
— (&+ 2)
exp
exp
2
H~ '(p)
2
(A7)
(&+ 2) H~'(p).
2
(AS)
that
a. Behavior a8 r -+ 0
1. The semiclassical WKB scheme
As
r
~
limit
oo, the Sabatier variable tends as 0. m kr
~ oo
as well and the quasipotential
Q(o) =
1
2iE)
/02
p2
(A2)
It is useful to consider the series representation of the
Bessel function and its use in the Hankel function specifications [14], from which, for each pole or zero value,
(p)-A
H(i)
I'( —A+ 1)
) p-+o
sin(Am)
(p)2 —A
tends to zero. More specifically, to erst order,
): -:-~.'j
N
~(-).- —.,
const
r —+on
2
+ I'( —
A+ 2)
(2)"
(A3)
p3
The scale constant is a purely real number if the set
{n„,P„) are complex conjugates. Then, from the specification of the potential in terms of the quasipotential, it
is simple to show that
and similarly for H& (p). Note that the essential order
of the terms in the curly brackets depends on the value
of Re(A).
If Re(A)
( 1, then
one finds
(p)
const
T~oo
(A4)
p3
(2)'
Asymptotic properties of the quantal inversion scheme
have been published [7], but are given again herein in
slightly more detail. Further, the asymptotic forms for
1
conditions not previously considered, i.e. , for Re(A)
are given.
With the rational form of the 8 function, Eq. (Al),
in the quantal inversion scheme, the potential takes the
form (p = kr)
(
='. 2k
where the matrix
„-)
d
1
p dp
p
„,
v-'(p)
so that
d
H(i)( p )
H(i) P
p ( )
y„ (p)
=
i-
f
P2
"-H
f
(p )
(p)
P2
Pwi
P+21)
P)2P—
(2
p 2P
p„,
r(P)
2
+
a
~
~
and
H(2)( )
)(P )
(t
P~& p
(p )
—
"H( )(
(i P)
I'(1+P) e
P
p
p~o
)
r(i —~)e
(
(P)2
+ 1)
)envri(P)2n
1)
r(1+~)
n
n2
n2
Pmo
P~,'(P)
P 1 —2 r( P)
d
(+)
—P)
( I'(i
r(i+p)
~
p
cx
)
(AIO)
r(A+ 1)
(All}
(A5)
Y has elements
(
—A
I ( —A + 1)
p~& sin(Avr)
2. The quantal inversion scheme
U(. )
(A9)
p
1
—2
n),
c
I'( —
p
I'(n)
2
+
\
~
~
(A12)
(A6)
Consequently
INVERSION OF ELECTRON-%PATER ELASTIC-SCATTERING DATA
r(
2
...
+, r(I'(cr) )
which lead to
p.;
p—
)
+ (1—P)(1—a) (2)
—2 ( p)
+ (~~+0
z)(p x)
z
y p
(p)
2
p~o
+
p(p+
cr)
1
Z
(A13)
( Re(cr),
""P)
'-(-P")1 p-~
r(p)
p~o
(P —1)(n —1)
p(p+n)
p
2
(A21)
Then for the condition that Re(p)
i'
'
'i
Epy(p))
~
-
~
(P
+
~
~
~
(A14)
)
so that
p+n
—2k
U(p)
'
(P —l)(cr —1)
(A22)
a potential that is 6nite and smoothly varying &om the
origin.
so that
U()
2k
' —
~
P
I'(P)
P —n
In the limit
This potential oscillates because
2
2Re(P —2) i2Im(P) lnp
2P —
=p
'(P
b. Behavior aa
. (A15)
2
}.
P
~()(,)
I (cx)
A
However, if Re(A)
' —
) 1, then
2
r( —A + 1)
( )'
I'( —A + 2)
2
7I
whence the logarithmic
behave as
L-
"
"
——
H(2)
one finds expansions
(l)
sin(Am)
(A17)
1
1
——
——
e i(p 2 vm 4 m)
'
(A24)
p
derivatives
() „:1—,, ( '
of the 3ost functions
—-'. )+
and the matrix elements of Y as
n
+
(A23)
Kp
and
For the condition Re(p) )Re(n), the same process
yields an oscillating potential as well with the form
—2k2
Hankel functions have asymi(p ——vm —4 m)
~(1) ~
(A16)
U( ) p~&
~ oo, the
~ ao
totic form
)( cos[2Im(P)lnp]
+i sin[2Im(P)lnp]
r
7
m
(A18)
)
(A26)
Consequently,
U(r)
(A19)
and
&',
H-''(P)
a")( )
~
'
1
P
f~+',
I
:.
(-;)
7
—+OO
) (P
——
I(,-
—+
1
) P3
—n
(A27)
For complex conjugate pole-zero pairs, n„= P„', and so
U is purely real. Note that the corresponding S function
then varies simply for large A as [7]
&
1 — ' (P)'~
S(A) -+ 1
(A2O)
[1] K. Chadan and P. S. Sabatier, Inverse Problems in Quan
turn Scattering Theory, 2nd ed. (Springer, Berlin, 1989).
[2] U. Buck, Comput. Phys. Rep. 5, 1 (1986).
[3] H. Pauly, in Atom Molecule Collision-Theory, edited by
R. B. Bernstein (Plenum, New York, 1970).
[4) L. J. Allen, Phys. Rev. A 34, 2708 (1986); L. J. Allen
—+
.. .
(A28)
I. E. McCarthy, ibid. 36, 2570 (1987).
Allen, K. Amos, C. Steward, and H. Fiedeldey,
Phys. Rev. C 41, 2021 (1990); L. J. Allen, H. Fiedeldey,
S. A. Sofianos, K. Amos, and C. Steward, ibid. 44, 1606
and
[5]
)
— (P„—e„)
L.
J.
(1991); L. J. Allen, K. Amos, and H. Fiedeldey,
G 18, L179 (1992).
J. Phys.
3798
[71
[8]
[9]
[10]
A. LUN, XUE JUN CHEN, L. J. ALLEN, AND K. AMOS
H. Fiedeldey, R. I ipperheide, K. Naidoo, and S. A. Sofianos, Phys. Rev. C 30, 434 (1984).
R. Lipperheide and H. Fiedeldey, Z. Phys. A 286, 45
(19?8); 301, 81 (1981).
L. 3. Allen, L. Herge, C. Steward, K. Amos, H. Fiedeldey,
H. Leeb, R. Lipperheide, and P. Frobrich, Phys. Lett. B
298, 36 (1992).
A. Katase, K. Ishibashi, Y. Matsumoto, T. Sakae, S.
Maezono, E. Murakami, K. Watanabe, and H. Maki, J.
Phys. B 19, 2715 (1986).
K. Naidoo, H. Fiedeldey, S. A. Sofianos, and R. Lipper-
heide, Nucl. Phys. A 419, 13 (1984).
[11] H. Lech, W. A. Schnizer, H. Fiedeldey, S. A. Sopranos,
and R. Lipperheide, jnv. Prob. 5, 817 (1989).
[12] H. Fiedeldey, S. A. Sopranos, L. 3. Allen, and R. Lipperheide, Phys. Rev. C 33, 1581 (1986).
[13] B. Apagyi, K.-E, May, T. Hauser, and W. Scheid, 3.
Phys. G 16, 451 (1990).
[14] Handbook of Mathematica/ Functions, Nat. Bur. Stand. ,
Appl. Math. Ser. No. 55, edited by M. Abramowitz and
I. A. Stegun (U. S. GPO, Washington, DC, 1965).