Next Article in Journal
Computation of Atomic Astrophysical Opacities
Previous Article in Journal
The ExoMol Atlas of Molecular Opacities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hybrid Theory of Scattering and Its Applications

Heliophysics Science Division, NASA/Goddard Space Flight Center, Greenbelt, MD 20772, USA
Atoms 2018, 6(2), 27; https://doi.org/10.3390/atoms6020027
Submission received: 12 April 2018 / Revised: 7 May 2018 / Accepted: 7 May 2018 / Published: 14 May 2018

Abstract

:
A number of formulations have been used to investigate scattering of low-energy electrons and positrons from various targets. The hybrid theory of scattering, which takes into account the short-range as well as the long-range correlations, and is variationally correct, is described in this article. This approach has been applied to calculate phase shifts for scattering of electrons and positrons, resonances in two-electron systems, photodetachment, and photoionization of two-electron systems. This approach has also been applied to calculate excitation of 2s state of atomic hydrogen by electron impact. In photoabsorption the target can be left in 2p state instead of 1s state, resulting in the emission of Lyman-alpha radiation. Cross sections for this process are also calculated.

1. Introduction and Calculations

Scattering by single-electron systems is always of interest because the wave function of the target is known exactly. Also, scattering from atoms helps us to understand their static and dynamic properties. It also has applications in astrophysics and in the investigation of controlled thermonuclear devices. Different approaches have been applied to study scattering of electrons and positrons from various targets, to infer the feasibility of various calculations and the accuracy of their results. In other words, this is an ideal problem for testing various theories. A simple method is to assume that the wave function of the incident electron on hydrogen atom is given by Ψ ( r 1 , r 2 ) = u ( r 1 ) Φ 0 ( r 2 ) , where Φ 0 ( r 2 ) is the ground state of the hydrogen atom and u ( r 1 ) is the function whose asymptotic limit determines phase shift of the scattering. Morse and Allis [1] introduced exchange between electrons, writing the wave function in the form:
Ψ ( r 1 , r 2 ) = u ( r 1 ) Φ 0 ( r 2 ) ± u ( r 2 ) Φ 0 ( r 1 ) .
The ground state function is given by:
Φ 0 ( r 2 ) = Z 3 π e Z r 2 .
The upper sign in Equation (1) corresponds to the singlet state and the lower sign corresponds to triplet state of the system. The equation for the scattering function u(r) (letting r1 = r) is obtained by the Kohn variation of the functional:
Y L 0 ( Δ 1 ) Φ 0 ( r 2 ) | H E | Ψ ( r 1 , r 2 ) = 0 ,
where E = k2Z2 and k2 is the kinetic energy of the incident electron, and Z is the charge of the nucleus. Considering the proton stationary during the interaction, the Hamiltonian H of the system in Rydberg units is given by:
H = 1 2 2 2 2 Z r 1 2 Z r 2 + 2 r 12
In the above equation, r 12 = | r 1 r 2 | . Using Equation (3), we get the integro-differential equation:
[ d 2 d r 2 + 2 ( Z 1 ) r L ( L + 1 ) r + e 2 Z r 2 ( Z + 1 r ) + k 2 ) ] u ( r ) ± 4 Z 3 e Z r [ ( k 2 + Z 2 ) r δ L 0 I 1 2 2 L + 1 ( I 2 r L + r L + 1 I 3 ) ] = 0 .
The exchange terms with I 1 , I 2 , and I 3 are the nonlocal potentials and are functions of u(r):
I 1 = 0 e Z x x u ( x ) d x ,
I 2 = 0 r e Z x x L + 1 u ( x ) d x ,
and
I 3 = r e Z x u ( x ) x L d x .
The singlet and triplet S-wave phase shifts in the exchange approximation are given in Table 1. Among the different methods used is the variational principle of Kohn [2]. Schwartz [3] used this principle to calculate phase shifts of electrons and positrons scattering from hydrogen atoms. There are no bounds on phase shifts in this principle except at incident energy k = 0. There are also singularities in this calculation. In spite of singularities, the results obtained in [3] are fairly accurate and have stood the test of time. Another extensively used method is the close-coupling approximation [4], in which the total wave function is expanded in eigenstates of the hydrogen atom. Accurate results have been obtained using the R-matrix formulation [5] in which most of the complications of interactions are taken into account within the inner region of a certain radius and outside this radius plane waves or Coulomb waves are used for continuum functions.
It has been emphasized by Wigner [8] that long range forces determine the nature of the scattering parameters at threshold. A method in which the longest-range potential proportional to −1/r4 is taken into account is the method of polarized orbital [6,9]. This method includes the essential physics of the problem and has been widely used not only to calculate scattering parameters for electrons and positrons but also to calculate photoabsorption cross sections. Using the first-order perturbation theory and the dipole part of the perturbed wave function, Temkin [9] showed that the effective target wave function in the presence of the incident electron can be written as
Φ p o l ( r 1 , r 2 ) = Φ 0 ( r 2 ) ε ( r 1 , r 2 ) r 1 2 u 1 s p ( r 2 ) r 2 cos ( θ 12 ) Z π
where
u 1 s p ( r 2 ) = e Z r 2 ( 0.5 Z r 2 3 + r 2 2 ) .
ε ( r 1 , r 2 ) = 1 , r 1 > r 2          = 0 , r 1 < r 2 .
Now, the function in Equation (1) is written as:
Ψ ( r 1 , r 2 ) = u ( r 1 ) Φ p o l ( r 1 , r 2 ) ± ( 1 2 ) .
The equation for the scattering function u L ( r 1 ) can be derived from:
Y L 0 ( Δ 1 ) Φ 0 ( r 2 ) | H E | Ψ ( r 1 , r 2 ) = 0
The equation for the scattering function u L ( r ) for all momenta has been given by Sloan [10]:
[ d 2 d r 2 + 2 ( Z 1 ) r L ( L + 1 ) r + e 2 Z r 2 ( Z + 1 r ) + k 2 ) ] u L ( r ) ± 4 Z 3 e Z r [ ( k 2 + Z 2 ) r δ L 0 I 1      2 2 L + 1 { r L I 2 + r L + 1 I 3 } ] = α ( Z r ) ( Z r ) 4 u L ( r )      4 Z e Z r [ δ L 1 3 { Z ( Z 2 + k 2 ) 2 r 3 + ( Z 2 k 2 ) r 2 } r e Z x x L u L ( x ) d x      + 2 ( Z r 3 + r 2 ) { L ( 2 L 1 ) ( 2 L + 1 ) r L 1 r e Z x x L 1 u L ( x ) d x      + L + 1 ( 2 L + 1 ) ( 2 L + 3 ) r L + 1 r e Z x x L 3 u L ( x ) d x } ]      ± 4 3 δ L 1 e 2 Z r [ ( 3 2 Z 2 r 2 + 1 2 Z r 3 ) u L ( r ) ( 1 2 Z r 2 + r ) d u L ( r ) d r ] ,
where
α ( x ) = 9 2 2 3 e 2 x ( x 5 + 9 2 x 4 + 9 x 3 + 27 2 x 2 + 27 2 x + 27 4 ) .
The scattering equation has the well-known attractive potential which for large r is equal to α / ( Z r ) 4 , α = 4.5 a 0 3 being the asymptotic polarizability for a hydrogen atom. Temkin and Lamkin [6], using this method of polarized orbitals, calculated S-wave phase shifts, as well as for higher partial waves, for scattering of an electron from a hydrogen atom. Phase shifts are determined by the asymptotic limit of the function u L ( r ) :
lim r u L ( r ) = sin ( k r π L / 2 + σ C + η ) r .
The Coulomb phase in the above equation is given by:
σ C = Z 1 k ln ( 2 k r ) + arg Γ ( L + 1 i ( Z 1 ) k ) .
S-wave phase shift obtained by this method are given in Table 1 and they differ from those obtained in the exchange approximation, showing the importance of polarization of the target resulting in the long-range potential −1/r4. However, this method does not provide variational bounds to the exact phase shifts.
The short-range correlations can be important in addition to the long-range correlations. These short-range correlations can be included using the projection-operator formalism of Feshbach [11]:
P = P1 + P2P1P2.  and    Q = 1 − P.
The wave function in Equation (1) is augmented by a correlation function Φ L ( r 1 , r 2 ) :
Ψ ( r 1 , r 2 ) = u ( r 1 ) Φ 0 ( r 2 ) ± ( 1 2 ) + Φ L ( r 1 , r 2 )
where
Φ L ( r 1 , r 2 ) = { f L κ , 1 ( r 1 , r 2 , r 12 ) D L κ , 1 ( θ , ϕ , ψ ) + f L κ , 1 ( r 1 , r 2 , r 12 ) D L κ , 1 ( θ , ϕ , ψ ) } .
The radial functions are of Hylleraas type and D’s are rotational harmonics [12].
The projections are defined by:
P 1 Φ ( r 1 , r 2 ) = Φ 0 ( r 1 ) Φ 0 ( r 1 ) Φ L ( r 1 , r 2 ) d r ¯ 1 .
P 2 Φ ( r 1 , r 2 ) = Φ 0 ( r 2 ) Φ 0 ( r 2 ) Φ L ( r 1 , r 2 ) d r 2 .
Using the projection operators P and Q, we get the equation for the scattering function u L ( ± ) ( r ) , having an optical potential [7]. The equation for the scattering function u L ( ± ) ( r ) is given by the following equation:
[ d 2 d r 2 + L ( L + 1 ) r 2 + V d ± V e x + V ( ± ) o p ( P , Q ) k 2 ] u L ( ± ) ( r ) = 0
where V d and V e x are the direct and nonlocal exchange potentials of the ‘exchange approximation’ [1]. V o p ( ± ) ( P , Q ) is the optical potential which is negative definite and therefore attractive. The optical potential acting on u L ( r ) is given by:
V o p ( ± ) ( P , Q ) u L = r Y L 0 P H Q 1 E Q H Q Q H P Ψ L .
We find S-wave phase shifts, given in Table 1, increase compared to the exchange approximation when short-range correlations are included but not the long-range correlations at the same time. However, the phase shifts obtained by this method have lower bounds to the exact phase shifts.
Is it possible to consider the long-range and short-range correlations at the same time? The projection-operator formalism, when both type of correlations are included, does not work because it is not possible to construct projection operators when Φ p o l ( r 1 , r 2 ) instead of Φ 0 ( r 2 ) is used in the definitions given in Equation (21). However, we can write a wave function which has a long-range and a short-range parts at the same time and we call this formulation hybrid theory because it includes both type of correlations: short-range as well as long-range at the same time. We describe this formalism by confining ourselves to the e-H partial wave (denoted by L) problem. The total spatial wave function is written as:
Ψ L ( r 1 , r 2 ) = u L ( r 1 ) r 1 Y L 0 ( Δ 1 ) Φ p o l ( r 1 , r 2 ) ± ( 1 2 ) + λ = 1 N C λ Φ L λ ( r 1 , r 2 ) .
The functions Φ p o l ( r 1 , r 2 ) and Φ L ( r 1 , r 2 ) are given in Equations (9) and (20), respectively. This approach is variationally correct and phase shifts obtained have lower bounds to the exact phase shifts. This has been achieved by introducing separate correlation functions and then amalgamating them into the scattering problem, via an optical potential, in order to replace the many-particle Schrödinger equation with a single-particle Schrödinger equation.
We can derive the equation for the scattering function u(r) from (letting r1 = r) from the functional:
I = Ψ L ( r 1 , r 2 ) | H E | Ψ L ( r 1 , r 2 ) = 0 .
For S-waves, we can write:
Φ L = 0 ( r 1 , r 2 ) = l m n N ω C l m n [ e γ r 1 δ r 2 r 1 l r 2 m r 12 n ± ( 1 2 ) ]
Clmn in Equation (26) are coefficients which are determined when eigenvalues are calculated using the Ritz variational principle. For N ω = 1, the functional can be written as:
I = A + C 1 B + C 1 2 D .
where
A = [ Φ p o l u ( r 1 ) ± ( 1 2 ) ] | H E | [ Φ p o l u ( r 1 ) ± ( 1 2 ) ]
and
B = 2 Φ L 1 ( r 1 , r 2 ) | H E | [ Φ p o l u ( r 1 ) ± ( 1 2 ) ]       = 4 V 1 ( r 1 ) u ( r 1 ) .
We can determine C1 by:
d I d C 1 = 0 .
which implies that:
B + 2 C 1 D = 0
This gives
C 1 = B / 2 D = 2 V 1 ( r 1 ) u ( r 1 ) E ε 1 .
where
V 1 ( r 1 ) = Φ L ( 1 ) ( r 1 , r 2 ) | H E | Φ p o l ) ,
D = Φ L ( 1 ) | H E | Φ L ( 1 ) = ε 1 E .
We can generalize the above treatment to any number of unknown coefficients in Equation (24). Now all the quantities in Equation (24) are known except u L ( r ) . The equation for the scattering function u L ( r ) can be derived from:
Y L 0 ( Δ 1 ) Φ p o l | H E | Ψ ( r 1 , r 2 ) = 0 .
The resulting equation is fairly complicated and the various quantities are given in [13]. We give a few of them below. We can generalize the above treatment to any number of unknown coefficients in Equation (24). We write the wave function in the form:
Ψ L ( r 1 , r 2 ) = u L ( r 1 ) Φ p o l ( r 1 , r 2 ) ± ( 1 2 ) + λ C λ Φ L λ ( r 1 , r 2 ) .
We define now:
Φ p o l ( r 1 , r 2 ) = Φ 0 ( r 2 ) χ β ( r 1 ) r 1 2 u 1 s p ( r 2 ) r 2 cos ( θ 12 ) π Z .
We have replaced the step function ε ( r 1 , r 2 ) by:
χ β ( r 1 ) = ( 1 e β r 1 ) n .
The exponent n is found to be 3 or 4. Now, polarization takes place for any value of r1, the coordinate of the incident electron. In the method of polarized orbitals, the orbital of the target was perturbed only when the incident electron was outside the target. Now perturbation takes place whether the incident electron is outside or inside of the target.
In addition to the nonlinear parameters in the correlation function, β is another nonlinear parameter and it is a function of the incident momentum k. This further helps in getting improved results. There is another form of the smooth cutoff function given by Shertzer and Temkin [14]:
χ S T ( r 1 ) = 1 e 2 Z r 1 ( 1 3 ( Z r 1 ) 4 + 4 3 ( Z r 1 ) 3 + 2 ( Z r 1 ) 2 + 2 Z r 1 + 1 ) .
All the quantities are now known in the wave function given in Equation (24). We can derive the equation for the scattering function using Equation (35). The equation for the scattering function is again is of the same form as Equation (22), but with more terms:
[ D ( r ) d 2 d r 2 L ( L + 1 ) r 2 + V d + V d p o l ± ( V e x + V e x p o l ) V o p p o l + k 2 ] u ( r ) = 0 .
As stated earlier, most of the quantities are given in [13]. Some of these quantities occurring in the direct potential are given below:
D ( r ) = 1 + 43 8 Z 6 ( χ ( r ) r 2 ) 2 .
The direct potentials are given by:
V d = 2 ( Z 1 ) r + 2 e Z r ( Z + 1 r ) .
We repeat here the exchange terms already indicated in Equation (5):
V e x u L ( r ) = 4 Z 3 e Z r { ( k 2 + Z 2 ) r δ L 0 I 1 2 2 L + 1 ( I 2 r L + 1 + r L + 1 I 3 ) }
V d p o l = ( x 1 + x 2 ) + x 3 d d r .
x 1 = 43 8 Z 6 ( 2 Z r 2 r 2 + k 2 ) ( χ ( r ) r 2 ) 2 .
x 2 = B 1 ( r ) 43 8 Z 6 ( χ ( r ) r 2 ) .
x 3 = B 2 ( r ) 43 8 Z 6 ( χ ( r ) r 2 ) + 2 α ( r ) ( Z r ) 4 χ ( r ) 9 2 χ ( r ) 2 ( Z r ) 4 8 3 Z d ( r ) ( χ ( r ) r 2 ) 2 ,
B 1 ( r ) = 1 r 2 [ 2 Z χ ( r ) + 2 Z 2 r + e 2 Z r ( 2 3 Z 4 r 3 4 3 Z 3 r 2 + 2 Z + 2 r ) ] ,
B 2 ( r ) = 2 Z B 1 ( r ) 4 Z r 3 + 6 r 4 + e 2 Z r ( 4 3 Z 5 r + 2 Z 4 4 Z 2 r 2 8 Z r 3 6 r 4 ) .
α ( r ) = 9 2 2 3 e 2 Z r ( 3 2 ( Z r ) 4 + 15 2 ( Z r ) 3 + 27 2 ( Z r ) 2 + 27 2 ( Z r ) + 27 4 ) .
The last term in x 3 has:
d ( r ) = 129 32 Z 5 1 r + 18 Z 7 1 r 3 e 2 Z r ( 3 16 r 4 + 27 16 Z r 3 + 54 8 Z 2 r 2 + 135 8 Z 3 r + 975 32 Z 4 + 1281 32 Z 5 1 r + 36 Z 6 1 r 2 + 18 Z 7 1 r 3 )
For r , x 3 has a term
9 2 Z 4
which is the dipole polarizability of the target with nuclear charge Z. V o p p o l = optical potential which is attractive, but does not depend on projection operators, and is given by:
V o p p o l u L ( r 1 ) = r 1 s N ω Y L 0 Δ 1 ) Φ p o l ( r 1 , r 2 ) | H E | Φ L s Φ L s | H E | Ψ L ' E ε s .
The function Ψ L is the function given in Equation (24) without the correlation terms. Phase shifts, given in Table 2, are calculated using the asymptotic limit of the function u L ( r ) given in Equation (16).
We see that the method of polarized orbitals over estimates phase shifts. In the variational formulation, phase shifts have lower bounds to the exact phase shifts, that is, they can be higher than those given in Table 2 when the number of correlations terms in Equation (24) is increased. The phase shifts obtained using hybrid theory agree well with those obtained by Schwartz [3] and also with those obtained by Scholz et al. [15] using the close-coupling approximation. Similar calculations have been carried out for S-wave electron scattering from He+ and Li2+ [16]. These results are shown in Table 3 and they agree with those obtained by Gien [17,18] using the Harris-Nesbet variational method.
Calculations have been carried out for partial wave L = 1 for electron scattering from H, He+, and Li2+ [19,20]. The P-wave phase shifts for electron-H scattering are given in Table 4. They are compared with those obtained using the R-matrix method [15] and by the finite-element method [21]. We see that the singlet P phase shifts have a peak at k = 0.3, while the triplet P phase shifts increase continuously.
Calculations have also been carried for partial wave L = 2 for electron scattering from H, He+, and Li2+ and photoabsorption when the final state is a 2p state [22].

2. Scattering Length

The scattering length a (in units of Bohr radius a 0 ) is defined by:
lim k 0 k cot ( η ) = 1 / a .
Temkin [23] has shown that the scattering length is significantly affected by the long-range polarization potential. It is given by:
a = a ( R ) α ( 1 R a R 2 + a 2 3 R 3 )
where α is the fine structure constant.
For 1S, for N ω = 70 we find a(R) = 6.00239 at R = 117.3038, using the above correction, we get a = 5.96595 compared to the value 5.965 ± 0.0003 obtained by Schwartz [3] who used the Kohn variational principle. The 3S scattering length for N ω = 84 is a(R) = 1.781542 at R = 349.0831. The corrected value is 1.76815 is little lower compared to the Schwartz a = 1.7686 ± 0.0002. The agreement with Schwartz’s values for both singlet and triplet S scattering lengths is very good.
The hybrid theory is akin to the close-coupling approximation in which the wave function is expanded in the orthonormal eigenstates of the target 1s, 2s, 2p…. However, in the close-coupling approximation, the asymptotic polarizability of the target depends on the number of p states included in the wave function. If 1s, 2s, and 2p eigenstates are included, then 66% of the polarizability is included, while in the hybrid theory the total asymptotic polarizability of the target is included and these calculations are also variationally correct. Also, the convergence in the close-coupling approximation is very slow as the number of eigenstates is increased.

3. Resonances

Calculations for phase shifts have been carried out for He+ and Li+2 targets in the resonance region [16]. These phase shifts are fitted to the Breit-Wigner form to obtain the resonance parameters:
η c a l c . ( E ) = η 0 + A E + tan 1 ( 0.5 Γ ( E R E ) ) .
E = k2 is the incident energy, ER is the resonance position and Γ is its width. The present calculation gives [16] ER = 57.8481 and Γ = 0.1233 eV for 1S state of He. The resonance position is with respect to the ground state of the He atom. These results agree well with those obtained using the Feshbach formalism [24]:
ER = 57.8435 and Γ = 0.125 eV.
In the Feshbach formalism, E R = ε Q + Δ , where ε Q = Q H Q and Δ is the correction due to the fact that the resonance is embedded in the continuum and there are other resonance states nearby. These corrections have to be calculated separately in order to get the position given above, as indicated in [24]. There are no such corrections when the resonance position is obtained by fitting resonance-region phase shifts to the Breit-Wigner expression given in Equation (55). There is no bound on the resonance position because of fitting of the phase shifts to E R and Γ in Equation (55) even when there is a lower bound on the phase shifts.

4. Low-Energy Scattering

It is known [25] that at low energies L = 1 scattering, the long-range correlations contribute most to the scattering phase shifts:
tan ( η ) / k 2 = π α / 15 A k .
So that
tan η ( k 1 ) / k 1 2 tan η ( k 2 ) / k 2 2 = A ( k 1 k 2 ) .
The first term in Equation (57) is due to the long-range potential (α is the polarizability) and the second term has contributions from the short-range correlations. Using the phase shifts given in Table 5, we find:
A T = 1.002 a n d A S = 2.942 .
These value compare reasonably well with the results A T = −1.3 and A S = 1.6 obtained in [25] by using the phase shifts obtained by the method of polarized orbitals [6]. The results in [6] are not accurate enough because of lack of any variational principle which is the reason for discrepancy.
For L equal to 1 and greater than 1, the first term in the expansion of tan η ( k ) is always proportional k2 as given below:
tan η ( k ) = π α k 2 ( 2 L 1 ) ( 2 L + 1 ) ( 2 L + 3 ) A k 2 L + 1 .
For L = 0, the expansion is complicated and involves terms like k, k2, ln(k), and the effective range. It is given in Equation (49) of reference [25].

5. Photoabsorption

It was pointed out by Wildt [26] that the opacity in the sun is due to the absorption of photons:
h v + H = H + e , h v + H e = H e + + e .
The photoabsorption or photodetachment cross section (in length form and in units of ( a 0 2 ) for a transition from an initial state ‘i’ to the final state ‘f’ is given by:
σ = 4π αkω|〈 Ψf | z1 + z2 | Φi 〉|2,
where α is the fine structure constant, k is the momentum of the outgoing electron, and ω is the energy of the incident photon:
ω = I + k 2 ,
where I is the ionization of the system absorbing the photon and k2 is the energy of the ejected electron. The Hylleraas-type normalized function Φ i is the ground-state wave function of the two-electron systems and it is given by:
Φ i = 1 8 π 2 C l m n [ e a r 1 b r 2 r 1 l r 2 m ± ( 1 2 ) ] r 12 n .
The upper sign refers to the singlet states and the lower sign refers to the triplet states. In the Table 6, the photodetachment cross sections of H [27] are given for a few values of k, momentum of the ejected electron. These results have been obtained without the short-range correlations. They are compared in Table 6 with those obtained by Bell and Kingston [28] who used the method of polarized orbitals. Their results are higher than those obtained using the variational approach [27]. Addition of short-range correlations in the final state wave function gave improved results for cross sections, given in Table 7. However, we find that at low energies the cross sections do not change considerably, showing that at low-energies the long-range correlations are more important than the short-range correlations.
Ohmura and Ohmura [29], using the effective range theory and the loosely bound structure of hydrogen ion, obtained:
σ = 6.8475 x 10 18 γ k 3 ( 1 γ ρ ) ( γ 2 + k 2 ) 3 cm 2 .
where γ = 0.2355883 and ρ = 2.646 ± 0.0004, γ is the square root of the binding energy of the electron and ρ is the effective range. The cross sections obtained using the effective-range theory are higher than those obtained in [27]. However, at low energies they are close to each other. Photodetachment cross sections for H are shown in Figure 1 from Ref. [27].
A similar calculation, including short-range and long-range correlations, has been carried out for the photoionization of He [27]. The results, given in Table 8, agree well with those obtained using the R-matrix approach [30] and the experimental results of Samson et al. [31]. Photoionization cross sections for Li ion have also been calculated and they are also given in [27].
We notice that in the limit k 0 , the photodetachment cross section for H goes to zero because the plane wave is normalized as sin(kr)/kr which is equal to 1 for r 0 , while the photoionization cross section of He goes to a finite value because of the final state Coulomb wave, which is proportional to 1 k for r 0 .

6. Recombination Rates

Knowing the photoabsorption cross sections, we can calculate radiative attachment rate coefficients which are important in the solar and astrophysical problems.
e + H H + h v e + H e + H e + h v
The attachment cross section σa is given by:
o a = ( h v c ) 2 1 2 m E g ( f ) g ( i ) σ .
This relation follows from the principle of the detailed balance. The weight factors in the initial and final states are g(i) and g(f), E = k2 is the energy of the electron, and k B is the Boltzmann constant in Equation (66), which gives the rate coefficient:
α R ( T ) = ( 2 S + 1 ) 10.2509 x 10 10 ( T ) 1.5 0 d E ( E + I ) 2 σ e E / k B T .
In Table 9, the rate coefficients for the singlet states of H, He, and Li+ are given at various temperatures [27]. We see that the rate coefficients for H and Li+ ions increase with temperature, attain a maximum value and then decrease, while in He atoms they decrease monotonically with the increase of the temperature.
Photoabsorption cross sections for excited states of the above systems and the recombination rate coefficients associated with these states have also been calculated [27]. In Table 10, rate coefficients for the excited states of He atoms and Li+ ions are given. As is well known, there are no excited states of H, therefore no results are given for this ion. We see that the singlet and triplet states the rate coefficients differ quite a bit.

7. Excitation of the 2S State

Cross section for excitation of the 2S state of atomic hydrogen at low incident electron energies has been carried out using different methods. We mention a few: Burke et al. [32] used close-coupling approximation retaining 1s, 2s, and 2p eigenstates. Callaway et al. [33] carried out calculations in the range 12 to 54 eV using 11-state expansion including seven pseudostates. A similar calculation in the above energy range has been carried by Callaway et al. [34] using six-state close-coupling expansion including an optical potential. The results of these calculations are essentially based on the close-coupling approximation. Scott el al. [35] used the standard R-matrix close-coupling method using nine-state basis, consisting of three eigenstates and six pseudostates. Poet [36] carried out calculations by assuming spherically symmetric wave functions. Lloyd and McDowell [37] used the method of polarized orbitals for the first three partial waves and Bessel functions for the higher partial waves. All these results are different from each other. Even though the hydrogen atom is the simplest system, it seems it is not easy to get exact cross sections for excitation. Excitation cross sections have been calculated using the variational method of polarized orbitals [13]. The excitation cross section [38] from state ‘i’ to state ‘f’ is given by:
σ = k f k i | T f i | 2 d Δ .
where ki and kf are the initial and final momenta and Tfi is a matrix element given by:
T f i = ( 1 / 4 π ) ψ f | V | Ψ i .
The interaction potential is:
V = 2 Z r 1 + 2 r 12
The initial state wave function Ψ i , in principle, is an exact solution of the Schrödinger equation [38]. Using this method, excitation cross sections have been calculated in the energy range 10.30 to 54.5 eV by using nine partial waves to get convergence, which is good up to the fifth and sixth decimal places for low energies, while at high energies it is good up to the third decimal place [38]. Some of the results of the various calculations are given in Table 11. The maximum of the cross section 0.137 π a 0 2 at 11.4 eV which is close to the experimental result 0.163 ± 0.2 π a 0 2 at 11.6 ± 0.2 eV of Kauppila et al. [39]. There is a minimum of the cross section at 16.46 eV and another maximum at 34.82 eV, as shown in Figure 2 from Ref. [38]. Differential and spin cross sections have also been calculated [38].

8. Lyman-Alpha Radiation

When the transition (2p- > 1s) takes place, we get electric dipole Lyman-alpha radiation which has a wavelength of 1216 A0 in H atoms and 304 A0 in He atoms. This radiation has been seen from the sun and from various astrophysical sources, and also from the Milky Way galaxy [41]. In most calculations, the 2p state is excited by electron impact. However, photoexciation is also possible: instead of leaving the target in the 1s state, it is left in the 2p state. The outgoing electron can then be in the final state lf = 0 or 2. In the dipole approximation, the cross section is:
σ = 4 π α k ω 3 ( 2 l i + 1 ) ( | M 0 | 2 + | M 2 | 2 ) .
The matrix element M is defined as:
M = ( 2 l f + 1 ) 0.5 | Ψ f | z 1 + z 2 | Φ i | .
In Table 12, we give photoabsorption cross sections [40] for two-electron systems for various values of the outgoing electron leaving the photoionized target in 2p 2P state. The calculation in this problem has been carried out using only the exchange approximation for the final continuum states. No short-range and long-range correlations have been included in this calculation. The purpose was to show that the cross sections are significant and this process should be considered along with excitation by electron impact. Photoabsorption cross sections when the final state is 2s 2S instead of 2p 2P are also given in [40], along with the recombination rate coefficients.

9. Positron-Hydrogen Scattering

It should be noted that for positron-hydrogen scattering Φ p o l has a positive sign instead of the negative sign as in the case of electron-hydrogen scattering:
Φ p o l ( r 1 , r 2 ) = ϕ 0 ( r 2 ) + χ S T ( r 1 ) r 1 2 u 1 s p ( r 2 ) r 2 cos ( θ 12 ) π Z .
In Table 13, we compare phase shift for S-wave scattering obtained in [42] using 35-term correlation functions with those obtained using 84-term correlation functions where projection operators were used [43]. With shorter expansions, improved results have been obtained, as also indicated in Figure 3 from Ref. [42].
The scattering length is −2.10258. Using the correction given in Equation (54), we find the scattering length is −2.104004 which agrees with −21036 ± 0.0004 obtained by Houston and Drachman [44], and with −2.103 ± 0.001 obtained by Humberston and Wallace [45].

10. Zeff

There is a possibility of positronium annihilation for which the cross section is given by:
σ a ( π a 0 2 ) = Z e f f α 3 / k .
Annihilation takes place with the emission of two gamma rays [46]. For hydrogen:
Z e f f = d r 1 d r 2 | Ψ L ( r 1 , r 2 ) | 2 δ ( r 1 , r 2 ) .
The normalization of the scattering function for r 1 is given by:
u L ( r 1 ) = [ 4 π ( 2 L + 1 ) ] 0.5 sin ( k r 1 L π 2 + η ) / k .

11. Positronium Formation

Positronium formation takes place when the incident positron captures an electron of the hydrogen target:
e + H ( 1 s ) P s + P .
Zeff and positronium formation have been calculated [42] using hybrid theory. The results for Zeff are given in Table 13 and positronium formation cross sections are given in [42]. A similar calculations have been carried out for P-wave scattering. Zeff and positronium formation cross sections have also been calculated [47].

12. Conclusions

In summary, we have used an approximation for scattering calculations in which the long-range and short-range correlations have been included at the same time variationally. The phase shifts for electron and positron scattering have a lower bound to the exact phase shifts. We have used these continuum functions to calculate resonance parameters, photoabsorption, and excitation cross sections, Zeff and positronium formation cross sections. The results are in agreement with those obtained in other calculations.

Acknowledgments

Thanks are extended to R.J. Drachman for discussions on low-energy scattering.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Morse, P.M.; Allis, W.P. The effect of Exchange on the Scattering of Slow Electrons from Atoms. Phys. Rev. 1933, 44, 269. [Google Scholar] [CrossRef]
  2. Kohn, W. Variational Methods in Nuclear Collision Problems. Phys. Rev. 1948, 74, 1763. [Google Scholar] [CrossRef]
  3. Schwartz, C. Electron Scattering from Hydrogen. Phys. Rev. 1961, 124, 1468. [Google Scholar] [CrossRef]
  4. Burke, P.G.; Smith, K. The Low-Energy Scattering of Electrons and Positrons by Hydrogen Atoms. Rev. Mod. Phys. 1962, 34, 458. [Google Scholar] [CrossRef]
  5. Burke, P.G.; Robb, W.D. The R-matrix Theory of Atomic Processes. Adv. At. Mol. Phys. 1976, 11, 143. [Google Scholar]
  6. Temkin, A.; Lamkin, J.C. Application of the Method of Polarized Orbitals to the scattering of Electrons from Hydrogen. Phys. Rev. 1961, 121, 788. [Google Scholar] [CrossRef]
  7. Bhatia, A.K.; Temkin, A. Complex-correlation Kohn T-matrix method of calculating total and elastic cross sections: Electron-hydrogen elastic scattering. Phys. Rev. A 2001, 64, 032709. [Google Scholar] [CrossRef]
  8. Wigner, E.P. On the Behavior of Cross Sections Near Threshold. Phys. Rev. 1948, 73, 1002. [Google Scholar] [CrossRef]
  9. Temkin, A. A note on the Scattering of Electrons from Atomic Hydrogen. Phys. Rev. 1959, 116, 358. [Google Scholar] [CrossRef]
  10. Sloan, I.H. The method of polarized orbitals for the elastic scattering of slow electrons by ionized helium and atomic hydrogen. Proc. R. Soc. Lond. A 1964, 281, 151–163. [Google Scholar] [CrossRef]
  11. Feshbach, H. A unified theory of nuclear reactions. II. Ann. Phys. 1962, 19, 287–313. [Google Scholar] [CrossRef]
  12. Bhatia, A.K.; Temkin, A. Symmetric Euler-Angle Decomposition of the Two-Electron Fixe-Nucleus Problem. Rev. Mod. Phys. 1964, 36, 1050. [Google Scholar] [CrossRef]
  13. Bhatia, A.K. Hybrid theory of electron-hydrogen elastic scattering. Phys. Rev. A 2007, 75, 032713. [Google Scholar] [CrossRef]
  14. Shertzer, J.; Temkin, A. Direct calculation of the scattering amplitude without partial-wave analysis III. Inclusion of correlation effects. Phys. Rev. A 2006, 74, 052701. [Google Scholar] [CrossRef]
  15. Scholz, T.; Scott, P.; Burke, P.G. Electron-hydrogen-atom scattering at intermediate energies. J. Phys. B 1988, 21, L139. [Google Scholar] [CrossRef]
  16. Bhatia, A.K. Applications of the hybrid theory in the scattering of electrons from He+ and Li2+ and resonances in these systems. Phys. Rev. A 2008, 77, 052707. [Google Scholar] [CrossRef]
  17. Gien, T.T. Accurate calculation of phase shifts for electron-He+ collisions. J. Phys. B 2002, 35, 4475. [Google Scholar] [CrossRef]
  18. Gien, T.T. Accurate calculation of phase shifts for electron collisions with positive ions. J. Phys. B 2003, 36, 2291. [Google Scholar] [CrossRef]
  19. Bhatia, A.K. Hybrid theory of P-wave electron-hydrogen elastic scattering. Phys. Rev. A 2012, 85, 052708. [Google Scholar] [CrossRef]
  20. Bhatia, A.K. Application of P-wave hybrid theory in the scattering of electrons from He+ and resonances in He and H. Phys. Rev. A 2012, 86, 032709. [Google Scholar] [CrossRef]
  21. Botero, J.; Shertzer, J. Direct numerical solution of the Schrödinger equation for the quantum scattering problems. Phys. Rev. A 1992, 46, R1155. [Google Scholar] [CrossRef] [PubMed]
  22. Bhatia, A.K. D-wave electron-H, He+, and Li2+ elastic scattering and photoabsorption in P states of two-electron system. Phys. Rev. A 2014, 89, 062720. [Google Scholar] [CrossRef]
  23. Temkin, A. Polarization and Triplet Electron-Hydrogen Scattering Length. Phys. Rev. Lett. 1961, 6, 354. [Google Scholar] [CrossRef]
  24. Bhatia, A.K.; Temkin, A. Calculation of autoionization of He and H using the projection-operator formalism. Phys. Rev. A 1975, 11, 2018. [Google Scholar] [CrossRef]
  25. O’Malley, T.F.; Rosenberg, L.; Spruch, L. Low-Energy Scattering of a charged Particle by a Neutral Polarizable System. Phys. Rev. 1962, 125, 1300. [Google Scholar] [CrossRef]
  26. Wildt, R. Electron Affinity in Astrophysics. Astrophys. J. 1939, 89, 295–301. [Google Scholar] [CrossRef]
  27. Bhatia, A.K. Hybrid theory of P-wave electron-Li2+ elastic scattering and photoabsorption in two-electron systems. Phys. Rev. A 2013, 87, 042705. [Google Scholar] [CrossRef]
  28. Bell, K.L.; Kingston, A.E. The bound-free absorption coefficient of the negative hydrogen ion. Proc. Phys. Soc. Lond. 1967, 90, 895. [Google Scholar] [CrossRef]
  29. Ohmura, T.; Ohmura, H. Electron-Hydrogen Scattering at low energies. Phys. Rev. 1960, 118, 154. [Google Scholar] [CrossRef]
  30. Nahar, S.N. New Quests in Stellar Astrophysics. II. The Ultraviolet Properties of Evolved Stellar Populations; Chavez, M., Bertone, E., Rosa-Gonzalez, D., Rodriguez-Merino, L.H., Eds.; Springer: New York, NY, USA, 2009; p. 245. [Google Scholar]
  31. Samson, J.A.R.; He, Z.X.; Yin, L.; Haddad, G.H. Precision measurements of the absolute photoionization cross sections of He. J. Phys. B 1994, 27, 887. [Google Scholar] [CrossRef]
  32. Burke, P.G.; Schey, H.M.; Smith, K. Collisions of Slow Electrons and Positrons with Atomic Hydrogen. Phys. Rev. 1963, 129, 1258. [Google Scholar] [CrossRef]
  33. Callaway, J. Scattering of electrons by atomic hydrogen at intermediate energies: Elastic Scattering and n = 2 excitation from 12 to 54 eV. Phys. Rev. A 1985, 32, 775. [Google Scholar] [CrossRef]
  34. Callaway, J.; Unnikrishnan, K.; Oza, D.H. Optical potential of electron-hydrogen scattering at intermediate energies. Phys. Rev. A 1987, 36, 2576. [Google Scholar] [CrossRef]
  35. Scott, M.P.; Scholz, T.T.; Walters, H.R.; Burke, P.G. Electron scattering by atomic hydrogen at intermediate energies: Integrated elastic, is-2s and 1s-2p cross sections. J. Phys. B 1989, 22, 3055. [Google Scholar] [CrossRef]
  36. Poet, R. The exact solution for a simplified model of electrons scattering by hydrogen atoms. J. Phys. B 1978, 11, 3081. [Google Scholar] [CrossRef]
  37. Lloyd, M.D.; McDowell, M.R.C. An application of the polarized approximation to electron-impact excitation of atomic hydrogen. J. Phys. B 1969, 2, 1313. [Google Scholar] [CrossRef]
  38. Bhatia, A.K. Excitation of the 2S State of Atomic Hydrogen by Electron Impact. Atoms 2018, 6, 7. [Google Scholar] [CrossRef]
  39. Kauppila, W.E.; Ott, W.R.; Fite, W.L. Excitation of Atomic Hydrogen to the Metastable 22S1/2 State by Electron-Impact. Phys. Rev. A 1970, 1, 1099. [Google Scholar] [CrossRef]
  40. Bhatia, A.K.; Drachman, R.J. Photoejection with excitation in H and other systems. Phys. Rev. A 2015, 91, 012702. [Google Scholar] [CrossRef]
  41. Lallement, R.; Que’merais, E.; Bertaux, J.-L.; Sandel, B.R.; Izmodenov, V. Voyager Measurements of Hydrogen Lyman-α Diffuse Emission from the Milky Way. Science 2011, 334, 1665–1669. [Google Scholar] [CrossRef] [PubMed]
  42. Bhatia, A.K. Positron-Hydrogen Scattering, Annihilation, Positronium Formation. Atoms 2016, 4, 27. [Google Scholar] [CrossRef]
  43. Bhatia, A.K.; Temkin, A.; Drachman, R.J.; Eiserike, H. Generalized Hylleraas Calculation of Positron-Hydrogen Scattering. Phys. Rev. A 1971, 3, 1328. [Google Scholar] [CrossRef]
  44. Houston, S.K.; Drachman, R.J. Positron-Atom by Kohn and Harris Methods. Phys. Rev. A 1971, 3, 1335. [Google Scholar] [CrossRef]
  45. Humberston, J.W.; Wallace, J.B.G. The elastic scattering of positrons by atomic hydrogen. J. Phys. B 1972, 5, 1138. [Google Scholar] [CrossRef]
  46. Ferrall, R.A. Theory of Positron Annihilation in Solids. Rev. Mod. Phys. 1956, 28, 308. [Google Scholar] [CrossRef]
  47. Bhatia, A.K. P-wave Positron-Hydrogen Scattering, Annihilation, and Positronium Formation. Atoms 2017, 5, 17. [Google Scholar] [CrossRef]
Figure 1. Photodetachment of detachment of a hydrogen ion. The lowest curve is obtained only when the long-range correlations are included; the middle curve is obtained when the short-range and long-range correlations are included. The top [27] curve is obtained when the effective range theory and the loosely bound structure of hydrogen ion used [29].
Figure 1. Photodetachment of detachment of a hydrogen ion. The lowest curve is obtained only when the long-range correlations are included; the middle curve is obtained when the short-range and long-range correlations are included. The top [27] curve is obtained when the effective range theory and the loosely bound structure of hydrogen ion used [29].
Atoms 06 00027 g001
Figure 2. (Color online) Total 1S–2S excitation cross section ( π a 0 2 ) vs. the incident momentum k(Ryd).
Figure 2. (Color online) Total 1S–2S excitation cross section ( π a 0 2 ) vs. the incident momentum k(Ryd).
Atoms 06 00027 g002
Figure 3. (Color online) The upper curve represents the present results, while the lower curve represents obtained by Bhatia et al. [43].
Figure 3. (Color online) The upper curve represents the present results, while the lower curve represents obtained by Bhatia et al. [43].
Atoms 06 00027 g003
Table 1. Phase shifts (radians) for the 1S and 3S states of e-H.
Table 1. Phase shifts (radians) for the 1S and 3S states of e-H.
KSingletTriplet
EAaPO [6]QHQ [7]EAaPO [6]QHQ [7]
0.08.1005.9 2.3491.9
0.12.3962.5832.553582.9082.9452.93853
0.21.8702.1442.066782.6792.7322.71741
0.31.5081.7501.698162.4612.5192.49975
0.41.2391.4691.415402.2572.3202.29408
0.51.0311.2511.200942.0702.1332.10454
0.60.869 1.040831.901 1.93272
0.70.7440.9470.931111.7491.8151.77950
0.80.6510.8540.887181.6141.6821.64379
EAa are exchange approximation phase shifts, and k = 0 results are scattering lengths.
Table 2. Comparison of phase shifts (radians) obtained in hybrid theory [13] with those obtained by Schwartz [3] and Scholz et al. [15], the results for k = 0 are scattering lengths.
Table 2. Comparison of phase shifts (radians) obtained in hybrid theory [13] with those obtained by Schwartz [3] and Scholz et al. [15], the results for k = 0 are scattering lengths.
kSingletTriplet
η H B [13] η s c h w a r t z [3] η S S B [15] η H B [13] η S c h w a r t z [3] η S S B [15]
0.05.995675.965 1.781541.7686
0.12.553702.5532.5502.938562.93882.939
0.22.067172.06732.0622.717512.71712.717
0.31.696841.69641.6912.499872.49962.500
0.41.415541.41461.4102.294572.29382.294
0.51.201951.2021.1962.105742.10462.105
0.61.041911.0411.0351.933361.93291.933
0.70.930840.9300.9251.779991.77971.780
0.80.888020.886 1.644441.643
Table 3. Phase shifts (radians) for scattering of electrons from He+ and Li2+ [16].
Table 3. Phase shifts (radians) for scattering of electrons from He+ and Li2+ [16].
kHe+Li2+
1S3S1S3S
0.10.438080.930650.231880.56084
0.20.435500.927040.231760.56020
0.30.431420.921140.231480.55869
0.40.426080.913020.231090.55678
0.50.419740.902820.230640.55435
0.60.412650.890570.230120.55142
0.70.405680.876450.229600.54799
0.80.398650.860660.229060.54413
0.90.391230.843660/228550.53925
1.00.386440.825360.228070.53514
1.10.382000.806360.227690.53000
1.20.379140.786770.227400.52456
1.30.378460.766960.227240.51880
1.40.381580.747080.227240.51276
1.50.398020.727460.227420.50646
1.60.344800.708150.22782049997
Table 4. P-wave phase shifts (radians) for electron-hydrogen scattering. η H B are the results obtained using hybrid theory [19], η R M are the R-matrix results [15], and η F E M are finite-element method results [21].
Table 4. P-wave phase shifts (radians) for electron-hydrogen scattering. η H B are the results obtained using hybrid theory [19], η R M are the R-matrix results [15], and η F E M are finite-element method results [21].
K1P3P
η H B η R M η F E M η H B η R M η F E M
0.10.006350760.0060.0060.010382340.0100.0100
0.20.015065560.0150.01480.045367350.0450.0452
0.30.016706340.0160.01600.10693120.1070.1067
0.40.010153470.0090.00900.18888730.1870.1873
0.5−0.00061223−0.002−0.00200.27097620.2700.2708
0.6−0.01009367−0.012−0.01170.34167490.3410.3417
0.7−0.01321557−0.016−0.01490.39321000.3920.3933
0.8−0.0046818 −0.00680.4277296 0.4283
Table 5. Low-energy P-wave phase shifts for electron-hydrogen scattering.
Table 5. Low-energy P-wave phase shifts for electron-hydrogen scattering.
Statek1 = 0.04k2 = 0.05
1P0.0013036920.001963464
3P0.0015643460.002469346
Table 6. Photodetachment cross sections (Mb) of the ground state of H, without short-range correlations [27] and comparison with those obtained by Bell and Kingston [28].
Table 6. Photodetachment cross sections (Mb) of the ground state of H, without short-range correlations [27] and comparison with those obtained by Bell and Kingston [28].
KNω = 220286364Ref. [28]
0.010.02430.2440.0245
0.052.70152.71482.7480
0.115.214715.232415.246512.34
0.238.376338.367538.368840.48
0.334.978334.980634.968436.40
0.424.249824.247224.253725.296
0.515.866315.869915.869216.43
0.610.497210.494210.492411.29
0.77.12347.12437.1258
0.746.15086.15216.1530
0.84.97624.97694.97685.31
0.85444.14234.14254.1421
0.86314.02304.02284.0224
0.86603.98523.98513.9846
Table 7. Photodetachment cross sections (Mb) of the ground state of H, with short-range correlations included in the final state [27].
Table 7. Photodetachment cross sections (Mb) of the ground state of H, with short-range correlations included in the final state [27].
KNω = 220286364
0.041.44641.45451.4750
0.052.70502.71852.7517
0.115.252615.270415.3024
0.238.551638.542938.5443
0.2339.576439.592739.6366
0.2539.069939.092539.1350
0.335.242035.244335.2318
0.424.473424.470924.4774
0.516.083016.086616.0858
0.610.745910.742810.7410
0.77.48377.48477.4862
0.746.60506.60636.6072
0.8005.65065.65145.6512
0.85444.14264.14254.1421
0.86316.89806.89846.8976
0.86607.62297.62307.6223
Table 8. Photoionization cross sections (Mb) of He [27] obtained with correlations.
Table 8. Photoionization cross sections (Mb) of He [27] obtained with correlations.
k N ω = 120165220Ref. [30]Ref. [31]
0.17.33197.33057.33007.2957.44
0.27.15637.15497.15447.1157.13
0.36.87336.87206.87166.8386.83
0.46.49656.49536.49516.4746.46
0.56.04716.04616.04616.0066.02
0.65.59295.59245.59255.5355.55
0.75.01215.01185.01204.9855.04
0.84.47384.47384.47404.4824.51
0.93.96473.96483.9649
1.03.46523.46543.46543.4763.48
1.13.02053.02063.02063.0233.00
1.32.25602.25612.25612.2712.19
1.41.98201.98211.98211.9431.89
1.51.68161.68171.6817
1.61.63241.63291.6329
Table 9. Recombination coefficients (cm3/s) for singlet states of H, He, and Li+ [27].
Table 9. Recombination coefficients (cm3/s) for singlet states of H, He, and Li+ [27].
TαR(T) × 1015, HαR(T) × 1013, HeαR(T) × 1013, Li+
10000.992.500.12
20001.282.391.04
50002.401.872.62
70002.821.662.92
10,0003.201.45 3.03
12,0003.371.353.02
15,0003.561.232.95
17,0003.651.172.89
20,0003.751.102.79
22,0003.791.052.73
250,0003.830.992.63
30,0003.830.922.49
35,0003.770.872.36
40,0003.670.822.25
Table 10. Rate coefficients (cm3/s) for (1s2s) 3S and (1s2s) 1S singlet excited states of He atom and Li+ ion [27].
Table 10. Rate coefficients (cm3/s) for (1s2s) 3S and (1s2s) 1S singlet excited states of He atom and Li+ ion [27].
THeLi+
αR(T) × 1014, 3SαR(T) × 1015, 1SαR(T) × 1014, 3SαR(T) × 1014, 1S
10002.138.274.682.99
20002.087.974.472.87
50001.737.303.482.27
70001.565.713.092.03
10,0001.405.052.681.78
12,0001.324.732.491.66
15,0001.234.352.261.52
17,0001.184.152.141.45
20,0001.123.901.981.36
22,0001.093.751.901.31
25,0001.043.571.791.24
30,0000.983.311.641.15
35,0000.933.101.521.08
40,0000.892.931.431.02
Table 11. Excitation cross sections ( π a 0 2 ) obtained in using variational polarized orbital method [40] and comparison with the results of other calculations.
Table 11. Excitation cross sections ( π a 0 2 ) obtained in using variational polarized orbital method [40] and comparison with the results of other calculations.
KHybrid Theory [38]Lloyd and McDowell [37]Burke et al. [32]Callaway [33]
0.876.8208(-2)
0.901.3589(-1)0.17890.09918
0.9041.3667(-1)
0.9051.3673(-1)
0.9061.3662(-1)
0.9071.3660(-1)
0.989.9106(-2) 0.161
1.008.9474(-2)0.14430.17830.154
1.028.1852(-2) 0.150
1.058.4057(-2) 0.142
1.106.9444(-2) 0.16540.135
1.207.3317(-2) 0.13490.111
1.508.5797(-2) 0.08910.087
1.608.4611(-2) 0.081
1.807.1804(-2) 0.074
2.006.9460(-2)0.11980.063550.066
Table 12. Photoabsorption cross sections (Mb) of the H anion, He, Li+, Be2+, and C4+. The final state is the 2P state [40].
Table 12. Photoabsorption cross sections (Mb) of the H anion, He, Li+, Be2+, and C4+. The final state is the 2P state [40].
KHHeLi+Be2+C4+
0.11.25113.4706(-2)
0.21.69832.2972(-2)3.8740(-3)6.9296(-3)1.8170(-4)
0.31.65783.0459(-2)3.8183(-3)6.7374(-3)1.8144(-4)
0.41.17092.7627(-2)3.7482(-3)6.4820(-3)1.8104(-4)
0.56.4369(-1)2.4948(-2)3.6694(-3)6.1657(-3)1.8048(-4)
0.64.4867(-1)2.2717(-2)3.5894(-3)5.8193(-3)1.7987(-4)
0.74.1448(-1)2.1120(-2)3.5102(-3)5.4450(-3)1.7916(-4)
0.83.5327(-1)2.0058(-2)3.4337(-3)5.0471(-3)1.7836(-4)
0.92.6847(-1)1.9375(-2)3.3712(-3)4.6498(-3)1.7505(-4)
1.01.508(-1)1.8811(-2)3.3083(-3)4.2507(-3)1.7641(-4)
1.1 1.8219(-2)3.2481(-3)3.8543(-3)1.7552(-4)
1.2 1.7409(-2)3.1894(-3)3.4825(-3)1.7437(-4)
1.3 1.6392(-2)3.1220(-3)3.1270(-3)1.7322(-4)
1.4 1.5181(-2)3.0533(-3)2.8077(-3)1.7198(-4)
1.5 1.3866(-2)2.9777(-3)2.4963(-3)1.7072(-4)
1.6 1.2498(-2)2.8919(-3)2.2171(-3)1.6963(-4)
1.7 1.1158(-2)2.7898(-3)1.9516(-3)1.6819(-4)
1.8 9.8391(-3)2.6849(-3)1.7123(-3)1.6678(-4)
1.9 8.6095(-3)2.5735(-3)1.4960(-3)1.6521(-4)
2.0 7.4791(-3)2.4546(-3)1.3015(-3)1.6371(-4)
Table 13. Comparison of phase shifts (radians) for e+-H scattering for various k obtained in [42] with those obtained in [43].
Table 13. Comparison of phase shifts (radians) for e+-H scattering for various k obtained in [42] with those obtained in [43].
kHybrid Theory [42]P, Q Projections [43] Z e f f
0.10.149180.14838.092
0.20.188030.18775.357
0.30.168310.16774.264
0.40.120830.12013.370
0.50.062780.06242.424
0.60.009030.00392.249
0.7−0.04253−0.05122.069

Share and Cite

MDPI and ACS Style

Bhatia, A.K. Hybrid Theory of Scattering and Its Applications. Atoms 2018, 6, 27. https://doi.org/10.3390/atoms6020027

AMA Style

Bhatia AK. Hybrid Theory of Scattering and Its Applications. Atoms. 2018; 6(2):27. https://doi.org/10.3390/atoms6020027

Chicago/Turabian Style

Bhatia, Anand K. 2018. "Hybrid Theory of Scattering and Its Applications" Atoms 6, no. 2: 27. https://doi.org/10.3390/atoms6020027

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop