SciELO - Scientific Electronic Library Online

 
vol.64 issue5Vegas: Software package for the atomistic simulation of magnetic materialsIn-situ monitoring by thermal lens microscopy of a photocatalytic reduction process of hexavalent chromium author indexsubject indexsearch form
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • Have no similar articlesSimilars in SciELO

Share


Revista mexicana de física

Print version ISSN 0035-001X

Rev. mex. fis. vol.64 n.5 México Sep./Oct. 2018  Epub Nov 30, 2019

 

Research

Differential and integral cross sections for electron elastic scattering by ammonia for incident energies ranging from 10 eV to 20 KeV

N. Sali1 

H. Aouchiche1 

1Laboratoire de mécanique, structures et énergétique, université Mouloud Mammeri de Tizi-Ouzou, B.P. 17, Tizi-Ouzou 15000, Algeria. e-mail: h_aouchiche@yahoo.fr


Abstract

Differential and integral cross sections for elastic scattering of electron by NH3 molecule are investigated for the energy ranging from 10 eV to 20 keV. The calculations are carried out in the framework of partial wave formalism, describing the target molecule by means of one center molecular Hartree-Fock functions. The potential used includes a static part -obtained here numerically from quantum calculation- and fine effects like correlation, polarization and exchange potentials. The results obtained in this model point out clearly the role played by the exchange and the correlation-polarization contributions in particular at lower scattering angles and lower incident energies. Both differential and integral cross sections obtained are compared with a large set of experimental data available in the literature and good agreement is found throughout the scattering angles and whole energy range investigated here.

Keywords: Electron-ammonia interaction; elastic scattering; correlation-polarization and exchange potentials; differential and integral cross sections

PACS: 34.80.-i

1. Introduction

Electron collisions with NH3 molecule are of prime interest in atomic and molecular processes in chemical reactions, collisions physics, interstellar space, atmosphere, radioactivity, plasmas (plasma etching, plasma deposition), switching device 1,2, biological matter and medicine. Ammonia is a colorless gas with a characteristic pungent smell; it is found in trace quantities in nature, being produced from the nitrogenous animal and vegetable matter, throughout the solar system, and in small quantities in rainwater. Ammonia is also known as caustic and hazardous in its concentrated form; it is currently used as fertilizer, cleaner, fuel, antimicrobial agent for food production, in fermentation, and textile. In addition, NH3 is one of the important molecules considered as a source of nitrogen atoms for the fabrication of nitride films and other nitrogen compounds 3. Considering these different domains of use, the e-NH3 interaction has been the subject of many investigations in various energy ranges theoretically as well as experimentally.

On the experimental side, the absolute cross section of e-NH3 interaction was measured first by Brüche in 1929 below 50 eV, using a Ramsauer-type apparatus 4, and then followed by the differential cross section (DCS) measurements reported by Harshbarger et al. (1971) 5 versus the momentum transfer between 2o-10o for 300, 400, and 500 eV, mentioning also unpublished data of Bromberg 6. The measurements of the total cross sections (TCS) on NH3 are renewed in 1987 by Sueoka et al. 7 using a linear transmission type time-of-flight apparatus from low to intermediate impact energies (0.7-400 eV). We also found the data reported by Szmytkowski et al. (1989) for impact energy from 1 to 80 eV, using a non-magnetic linear transmission technique 8. Particularly, the results show a very broad hump centered around 10 eV. At the same time, Pritchard et al. 9 mentioned that Shyn has measured the relative DCSs for e-NH3 interactions at 8.5 eV and 15 eV at intermediate and large angles 10. The TCSs have been also measured in a transmission technique using a Ramsauer-type configuration in the 75-4000 eV energy range by Zeeca et al. (1992) 11. At the same time, Alle et al. reported absolute TCS and DCSs measurements for vibrationally elastic electron scattering from NH3 et al at incident energies 2-30 eV, using a crossed electron-molecular beam apparatus 12. While in a wider energy range 300-5000 eV, Garcia and Manero (1996) used a method based on a transmission-beam technique including a detailed error-source analysis 13. In another range of energy 400-4000 eV, Ariyasinghe et al. (2004) have measured the data from a linear transmission technique based on the electron-beam intensity attenuation through a gas 14. Recently, Jones et al. (2008) reported the data in the low energy range 20 meV-10 eV including both integral scattering and scattering into the backward hemisphere with high energy resolution 15. Furthermore, the elastic and inelastic scattering DCSs of high energy (35 keV) electrons scattered by NH3 molecules were measured separately for the first time by Lahmam-Bennani et al. 16-17 and Dugu et al. 18, using two independent experimental methods. Finally, let us notice that many discrepancies exist between the cross sections produced in different laboratories. As mentioned by Karwazs et al. the most serious deviations arise in the TCS maximum resonant region (up to 25%), around 10 eV, and at high energies (to 35%) 19.

On the theoretical side, Itikawa (1971) adopted a general form of electron-molecule interaction in the Born approximation and derived a cross section formula for the rotational transition in a symmetric-top molecule and applied it to the e-NH3 collisions to evaluate the DCSs for energies lesser than 0.1 eV 20. Many years later, Jain (1988) carried out calculation of the TCSs (elastic + inelastic) from 10 eV to 1000 eV, using a parameter-free spherical complex optical potential with different potential component models 21. As the author mentioned, the results were so distinct that their comparison with the experimental data can help to select the model describing well the scattering process. At the same time, Benarfa and Tronc (1988) investigated the vibrational excitation and then calculated the DCSs and integral cross sections (ICS) at low energies (3-10 eV); particularly they analyzed the angular distributions of the elastic peak of the nv1 series and of the v 4 vibrational mode 22. Furthermore, Pritchard et al. (1989) reported differential and momentum transfer cross sections of elastic scattering for e-NH3 collisions from 2.5 to 20 eV in the fixed-nuclei static-exchange approximation using the Schwinger variational principle 9. Sometime later, Gianturco (1991) developed a model - to calculate the DCSs and ICSs for electron-NH3 interactions - based on a free parameter correlation-polarization and exchange potentials, which play a fundamental role in describing the scattering of slow electrons from molecular targets23. In another calculation, Yuan and Zhang (1992) used the first-order Born approximation with the rotating molecule model, thus the total, differential, and momentum-transfer cross sections are reported for the vibrationally elastic scattering of electrons in the energy range 0.5-20 eV 24. Besides, Rescigno et al. (1992) developed an ab-initio optical potential study using Kohn variational model including the static-exchange and polarized-self consistent-field of low-energy 25. Their results demonstrate the sensitivity of the CSs to the short-range repulsion, long-range polarization, and the effects of the dipolar field particularly between 2 and 7.5 eV. Few years later, elastic and inelastic CSs for e-NH3 interactions are also calculated in the energy range 50-200 eV by Joshipura and Patel (1996) using a modified additivity rule by separating e-molecule interactions into short-range and long-range parts26. Furthermore, Liu et al. (1997) have used the semi-empirical formula, complex optical potential, and the additivity rule in the intermediate and high energy range to evaluate the TCSs from 10 eV to 1 keV 27. At the same time, Garcia and Manero (1997) used the empirical formula in the energy range 0.5-1 keV; their results reproduce sufficiently well (within 6%) their measurements for the series of investigated molecules (NH3, CH4, N2, CO and CO2) 28. The DCSs are also obtained from 8.5 to 30 eV by Lino (2005) using the Schwinger variational principle 29. As the Cartesian Gaussian functions basis - usually used in the method - is very effective only for short-range potentials, the authors introduced the plane waves as a trial basis set 29. Recently, Munjal and Baluja (2006) reported also the DCSs, ICSs, momentum transfer, and excitation cross sections for the low-energy (0.025-20 eV) electron-NH3 scattering, using the R-matrix method 30. Few years later, Limbachya et al. proposed two different methods to calculate the rotationally elastic CSs for electron scattering from NH3 attempting to demonstrate the possibility of producing robust cross sections over a wide energy range (0.01 eV to 2 keV) adaptable to any target 31. The UK molecular R-matrix code through the Quantemol-N software package is used to calculate elastic plus electronic excitation cross sections for incident energies below the ionization threshold of the target, while at higher energies the spherical complex optical potential formalism was used 31. Finally, note that the CSs for numerous polyatomic targets have been reviewed in these last decades by Shimamura 32, Hayashi 33, Morgan 34 and Karwazs et al. 19.

Despite the above-cited works, the differential and integral cross sections for electron elastic scattering by ammonia have not received all the attention merited. Regarding the works dedicated to the elastic scattering differential cross sections, they are very scarce in intermediate and higher energy range. Under these conditions, we propose in this paper to calculate elastic DCSs and ICSs for e-NH3 molecule interactions in the energy ranging from 10 eV up to 20 keV. Furthermore, the spherical complex optical potential including static potential and the fine effects like correlation-polarization and exchange potentials are considered and their effects discussed. The static potential is obtained here numerically from quantum calculation using molecular wave functions determined by Moccia 35 while the correlation-polarization and exchange potentials are rigorously selected from the literature. In the following sections, atomic units (a.u.) are used everywhere.

2.Theoretical model

The calculations of differential cross section dσ(Ω))/dΩ are developed in the coplanar geometry within partial wave formalism and non-relativistic approach, where Ω is the solid angle of the scattered electrons. Under these conditions dσ(Ω))/dΩ can be given by the square of the scattering amplitude f(θ), whith

f(θ)=1kl=0(2l+1)eiδlsinδlPl(cosθ), (1)

k is linked to the kinetic energy E by 2E=k2,l the quantum number of the kinetic momentum, Pl(cosθ) the Legendre polynomial and δl is the phase shift induced by the spherical potential V(r) in the outgoing wave relatively to the free wave (for more details see Ref. 36). NH3 being a pyramidal molecule with a heavy atom N, the charge distribution can be assumed as a spherical molecule centered at the nucleus N. Under these assumptions, the total potential V(r) may be approximated by a spherical one including the static potential Vst(r) and the two fine effects called polarization potential, Vp(r), and exchange potential, Vex(r)

V(r)=Vst(r)+Vp(r)+Vex(r). (2)

2.1.Static potential

To evaluate the static potential, we use the ammonia molecular orbital functions determined by Moccia 35. In this description, each molecular orbital was expressed in terms of Slater-type functions all centered at a common origin (the heaviest atom). Thus, the ammonia target can be described by means of Norb(Norb=5) molecular functions labeled 1A1, 2A1, 1A3, 1Ex and 1Ey, written as 37

Ψj(r)=k=1NjajkRnjk,ljkξik(r)Sljk,mjk(r^), (3)

where N j is the number of Slater functions used to construct the jth (with j varying from 1 to Norb) molecular orbital Ψj(r); r^ refers to the solid angle of the position vector r, and ajk the weight of each real atomic component. The function Rnjk,ljkξik(r) represents the radial part expressed by Slater-type functions (38) as

Rnjk,ljkξik(r)=(2ξjk)njk+1/2(2njk)!r(njk-1)e-ξjkr; (4)

and the angular part Sljk,mjk(r^) are the real spherical harmonics linked to the complex form Yljk,mjk(r^) by 39

*20lSljk,mjk(r^)=mjk2|mjk|1/2Yljk,-|mjk|(r^)+(-1)mjkmjk|mjk|Yljk,|mjk|(r^)ifmjk0Sljk,0(r^)=Yljk,0(r^)ifmjk=0 (5)

For more details, we refer the reader to Refs. 35 and 37 where all the coefficients (ajk,ξjk) and all quantum numbers (njk,ljk,mjk) used are reported. Besides note that the static potential includes electronic and ionic contributions

Vst(r)=[Vstj(r)]elec+[Vstj(r)]ion (6)

The electronic contribution is given for each molecular orbital j by

Vstj(r)elec=2k=1,k´=1Njajkajk,δ(mjk-mjk´)×δ(ljk-ljk´)0Rnjk,ljkξikr´r´2r>×Rnjk´,ljk´ξik´r´dr´ (7)

and the ionic one by

[Vstj(r)]ion=-Zr-1R>. (8)

The symbols r> and R> are respectively given by r>=max{r,r'}, and R>=max{r,RNH} where RNH=1.928 a.u. 35 is the internuclear distance between the two atoms N and H, and Z = 10 for ammonia molecule.

2.2.Correlation-polarization potential

It is well established that NH3 is polar covalent moleculedue to the electronegativity difference between the nitrogen (higher) and the hydrogen atoms. Hence, e-NH3 interactions induce an additive potential due to the polarizability of the target induced by the incident electrical charge particularly at low velocities. The polarization potential Vp(r) used here refers to the well-known Buckingham type given by

Vp(r)=-αd2(r2+rc2)2 (9)

where αd is the polarizability (αd=14.984 a.u. for NH340) and rc is a cut-off-parameter expressed by Mittleman and Watson 41 as

rc=12αdz-1/3bpol21/4, (10)

with bpol an adjustable parameter. In our case, we found that (see Salvat et al. for more details 42)

bpol=max{(E-0.5)/0.01;1} (11)

where E refers to the incident electron energy.

As proposed by Salvat et al. 42, we use here the polarization potential combined with the correlation one, called correlation-polarization potential. Different expressions of the correlation potential are found in the literature; the first expression chosen here is given by Padial and Norcross (1981) 43

*20lVc1(r)=0.0311log(rs)-0.0584+0.006log(rs)-0.015rsforrs<0.7Vc1(r)=-0.07356+0.02224log(rs)for0.7rs10 (12)

and the second one is given by Carr and Maradudin (1964) 44

*20lVc2(r)=0.0311log(rs)+0.0584-rs(0.00133log(rs)-0.0084)forrs1Vc2(r)=-β01+(7/6)β1rs1/2+(4/3)β2rs(1+β1rs(1/2)+β2rs)2forrs1 (13)

where rs=(3/4πρ(r))1/3) and β0=0.1423, β1=1.0529 and β2=0.3334. Following Salvat et al. 42 the correlation-polarization potential can be given by

Vcp(r)=*20lmax{Vc(r),Vp(r)}forr<rcrVp(r)forr>rcr; (14)

where rcr is defined as the outer radius where the two potetials Vp(r) and Vc(r) cross.

2.3.Exchange potential

In addition to the polarization phenomena of the target in the electron-molecule interactions, the incident electron can be captured by the target - particularly at low incident energies - and then the latter releases one of their bound electrons. This phenomenon known as electron exchange process needs another potential called exchange potential. Among the various formulas available in the literature, we have selected two expressions; the first one Vex1(r), depending clearly on the static potential and the correlation-polarization one, was given by Furness and McCarty 45

Vex1r=12E-Vstr+Vcpr-(E-Vstr+Vcpr)2+4πρ(r)1/2 (15)

where ρ(r) is the spherical density of the electronic charge given by 45

ρ(r)=1Norbj=1Norb|Rnjk,ljkξik(r)|2. (16)

The second exchange potential Vex2(r), used here, was initially given by Mittleman and Watson 46 and re-written by Riley and Truhlar 47:

Vex2r=-12πkLkLkF-12(kL2-kF2)×LnkL+kFkL-kF (17)

where kL is the local wave number of the electron projectile defined by kL2=k2+2I++kF2,I + being the first ionization potential of the molecule target, kF(r)=(3π2ρ(r))1/3 the Fermi wave number of the atomic electron cloud, and k2=2E. For illustration, we have reported in Fig. 1 the various potentials used in this work and described above. In Fig. 1(a), the static potential - calculated according to Eqs. (6), (7) and (8) - decreasing rapidly up to large distances present a discontinuity around =1.93 a.u due to the presence of the proton shell in the NH3 molecule where the internuclear distance RNH=1.928 a.u. In panel (b), the two correlation-polarization potentials Vcp1(r) and Vcp2(r) are compared, we clearly observe for r<rcr a quasi-linear behavior with the radius whereas they both mimic the asymptotic Vp(r) potential and then depend on the projectile energy at large distances. The two exchange potentials Vex1(r) and Vex2(r) are reported in Fig. 1(c), the first one presents no minimum contrary to Vex2(r) which exhibits clearly a minimum for r = 0.07 a.u. However, they have nearly the same values at large distances r.

FIGURE 1 (Color online). Different potentials used in this work. (a) Static potentials used to describe the electron elastic scattering by ammonia molecules, (b) Correlation- polarization potentials V cp1 ( black solid line) and V cp2 (red dotted line), (c) Exchange potentials V ex1 (black solid line) and V ex2 (red dashed line). 

Finally, note that here we have studied studied two expressions for the correlation-polarization potential (Vcp1(r) and Vcp2(r)) and two others for the exchange potential (Vex1(r) and Vex2(r)). So, according to Eq. (2), we get four combinations for the potential V(r), namely Vst(r)+Vcp1(r)+Vex1(r), Vst(r)+Vcp1(r)+Vex2(r), Vst(r)+Vcp2(r)+Vex1(r), and Vst(r)+Vcp2(r)+Vex2(r) which are used in our calculation.

3. Results and discussion

To evaluate the exchange potential (see Eq. (15)), we need, first, to calculate the electronic charge density, p(r), of the ammonia molecule. Figure 2 reports the obtained electronic charge density versus the radius r accordingly to Eq. (16), using the molecular wave functions determined by Moccia 35. The distribution exhibits two maxima located around 0.15 a.u. and 1.16 a.u, and one minimum around 0.5 a.u., as expected. The distribution shows no significant values over the radial distance r = 6 a.u. hence, the calculations can be achieved only in the range given by 0r6 a.u.

FIGURE 2 Electronic charge density of the ammonia molecule. The distance r refers to the center of the molecule, which is here merged with the heaviest atom center. 

3.1.Differential cross sections

Differential cross sections for elastic scattering of electron by NH3 molecule are calculated for scattering angles varying from 0 to 360o and for numerous incident energies. First, these differential cross sections are reported in Fig. 3 (left panel) for 15 eV, 20 eV and 30 eV, where experimental data are available and only from 0 to 180o since they are symmetrical with respect to the axis θ=180. The calculations are carried out for the four above-cited combinations of potentials investigated in this work. To the best of our knowledge, the only existing experimental DCSs related to the energies investigated here, were reported by Alle et al. 12 and Shyn10 and we are not aware of the existence of other measurements.

The DCSs obtained (left panel in Fig. 3) are important in the forward directions and lesser in the backward directions, while they are even smaller exhibiting a valley around the perpendicular directions, as expected. In general, the different exchange and correlation-polarization potentials, investigated here, provide quasi-identical DCSs and a good agreement is observed throughout the angle range for all the energies investigated. However, at lower scattering angles, the calculated DCSs exhibit little discrepancies particularly at lower energies. In fact, these discrepancies related to the form of the correlation-polarization potential considered are directly linked to the polar nature of the target since there is a mutual influence between the electric dipole of the molecule and the charge of the projectile. It is then evident that this influence, responsible of the discrepancies, became relatively important at lower incident energies, since the projectile spends a relatively important time near the target molecule. In addition, the DCS curves show the existence of a profound minimum, in the exhibited valley, whose depth depends on the form of the tested correlation-polarization and exchange potentials. The minima are located around θ=109 at 15 eV which are shifted to lower scattering angles when the incident energy increases, namely around θ=105 at 20 eV and around θ=99 at 30 eV. Note that, outside the observed minima, our calculated DCSs are in excellent agreement with the experimental data reported by Alle et al. 12 for vibrationally elastic scattering (black full squares in Fig. 3), using a crossed electron-molecular beam apparatus 12. Furthermore, our results are also in good agreement with the relative measures of Shyn 10 (red circles in Fig. 3)) mentioned and normalized at 90o by Pritchard et al. 9. We would note that the profound minima exhibited in these calculated DCSs for elastic scattering by NH3 are also found in our precedent works on H2S 48 and HCl 49 at lower energies using the same model. Otherwise, in Fig. 3 (right panel) the theoretical DCSs taken from 29 (solid navy line), 30 (dashed dark-yellow line), 25 (dotted magenta line), 23 (dashed-dotted blue line), 24 (dashed dotted-dotted red line), 9 (short dashed black line) are also reported to compare with. Note that these DCSs do not show a profound minimum except that obtained by the Schwinger variational principle with plane waves 29 (dashed dotted-dotted red line at 15 eV and 20 eV) based on fixed-nuclei approximation. As already pointed out by Brescansin et al50, these deep minima are probably due to the fact that only the spherical part of the interaction potential is considered in the calculations. Indeed, the spherical potential used here for describing the electron elastic scattering by the NH3 molecule is quite similar to that of the Neon atom for which the numerous DCSs reported in the literature 51 clearly point out an identical depth at lower energies. We would note here, that at intermediate and relatively high energies these very deep minima disappear completely, as we will see below. We can conclude that they are non-physical structures and probably underline the limitation of the current spherical approach for modeling the elastic scattering process at lower energies.

FIGURE 3 (Color online) In left panel: Comparison between calculated DCSs for electron elastic scattering by ammonia molecule at 15, 20, and 30 eV for various potentials: V st (r) + V cp1 (r) + V ex1 (r) (solid black line), V st (r) + V cp1 (r) + V ex2 (r) (dashed red line), V st (r) + V cp2 (r) + V ex1 (r) (dotted green line) and V st (r) + V cp2 (r) + V ex2 (r) (blue dash-dotted line). Experimental data are taken from 12 (black solid squares) and 10 (red open circles). In the right panel: theoretical results are taken from 9 (short-dashed black line), 23 (dash-dotted blue line), (24) (dash-dotted-dotted red line), 25 (dotted magenta line), 29 (solid navy line), and 30 (dashed dark-yellow line). 

In Fig. 4, we report the DCSs for electron elastic scattering at intermediate incident energies (300 eV, 400 eV, 500 eV, and 1 keV) for the four combinations of correlation-polarization and exchange potentials investigated in this work. These latter provide quasi-identical curves, and an overall agreement is observed in the whole angle range for all the energies investigated. The effects of the fine contributions (correlation-polarization and exchange potentials) no longer appear even in the range of lower scattering angles. Besides, the observed minima at low incident energies disappeared from the value of 200 eV (not reported here). In addition, we have reported in Fig. 4 the experimental data provided by Harsbarger et al. 5 (black solid squares) and by Bromberg 6 (red open circles) for the scattering angles from 2 to 10o. Our calculated DCSs appear slightly lesser than the experimental ones around 2o and seem to be in good agreement beyond this scattering angle value. Finally note that, to the best of our knowledge, there are no other results neither experimental nor theoretical in larger angle range to compare with.

FIGURE 4 (Color on line) As in Fig. 3 at 300 eV, 400 eV, 500 eV and 1 keV. Experimental data are taken from 5 (black solid squares) and 6 (red open circles). 

3.2.Integral cross sections

Integral cross sections for electron elastic scattering from NH3 molecules are also calculated using the four above-cited combinations including the static, correlation-polarization, and exchange potentials. The energy distribution for elastic scattering is obtained by integrating numerically the differential cross sections overall the scattering solid angle

σ(E)=2π0πdσ(E,θ)dθsinθdθ. (18)

The various obtained results are presented in Fig. 5a for the incident energies ranging from 10 eV to 20 keV, and for the four potential combinations investigated. As expected, the amplitude is important at lower incident energies and decreases monotonically with the increasing of the impact energy. In addition, the four calculated ICSs exhibit a good agreement at intermediate and high incident energies. In fact, the little discrepancies observed at low energies (30 eV) are undoubtedly due to the exchange and polarization effects since the ammonia molecule is a polar one with a dipole moment equal to 1.42 D. We can conclude here that the polarization-correlation and exchange phenomena have perceptible effects only at lower incident energies. We have also reported in Fig. 5a) the experimental data to compare with. Note that the available measurements in the literature concern essentially the total (elastic + inelastic) cross sections, see for example Refs. 7,11,13,14. Other experimental results are available in the literature but not reported since they are out of the energy range investigated here. First, note that some disagreements are observed in the experimental results as already mentioned by different authors 19,26. Besides, a simple comparison with our results shows that the predicted ICSs reproduce well the shape of the measures. Concerning the magnitude, the calculated ICSs with the potential combinations Vst(r)+Vcp1(r)+Vex2(r) (dashed red line), and Vst(r)+Vcp1(r)+Vex2(r) (dash-and-dotted blue line) are in better agreement with the experimental data of Alle et al. 12 (black full squares) and those of Sueoka et al. 7 (magenta open circles). At high energies, all the calculated ICSs are in satisfactory agreement with the available measurements except that reported by Garcia and Manero 13 (black full circles). Besides, the calculated ICSs seem slightly lesser than the experimental ones in magnitude over all the energy range since the latter include the elastic and inelastic contribution (rotational and/or vibrational).

FIGURE 5 a) (Color online) Comparison between calculated ICSs for electron elastic scattering by ammonia for various potential combinations: V st (r) + V cp1 (r) + V ex1 (r) (solid black line), V st (r)+V cp1 (r)+V ex2 (r) (dashed red line), V st (r)+V cp2 (r)+ V ex1 (r) (dotted green line) and V st (r)+V cp2 (r)+V ex2 (r) (dashand-dotted blue line). Experimental data are taken from 12 (black solid squares), 7 (magenta open circles), 11 (blue up triangles), 14 (red stars), 13 (solid black circles), and 8 (olive diamonds). b) (Color online) As in Fig. 5a. Thin lines are theoretical results taken from 21 (red solid line and dark-yellow dashed dotted line), 26 (dashed blue line), 27 (dotted magenta line and short dashed purple line) and 31 (dashed-dotted-dotted pink line). 

The obtained cross sections (thick lines) using different potentials are also reported in Fig. 5b where they are compared to the theoretical results existing in the literature (thin lines) taken from 21 (solid red line and dashed dotted dark-yellow line), 26 (dashed blue line), 27 (dotted magenta line and short dashed purple line), and 31 (dashed dotted-dotted pink line). Note that the shapes of all the different cross sections (thin lines) are identical with that calculated in this work (thick lines). However, the magnitudes of our CSs are, in general, lesser than the former except that of Jain 21 (solid red line). In fact all the different previous CSs include the inelastic contribution except that of Jain 21 where the elastic CSs are considered in the static-exchange-polarization (SPE) model. Finally, it is worth noting that the CSs reported by Jain 21 including the absorption part (dashed-dotted dark yellow line) are higher in magnitude than that obtained in this work confirming in this way these arguments.

4. Conclusion

In the present work the interactions of electron with ammonia molecule are investigated and the differential and integral cross sections are calculated for the elastic scattering in the partial wave formalism. The target molecular state has been described by means of a single-center molecular wave function, while the interactions by means of complex optical potentials including a static contribution - numerically obtained, here, from quantum calculations - and fine effects like correlation, polarization and exchange phenomena rigorously selected from the literature.

The obtained DCSs for different potentials investigated show clearly the role played by the correlation, polarization and exchange phenomena particularly at lower scattering angles and lower incident energies. The calculated DCSs have shown good agreement with the experimental data both in shape and magnitude, demonstrating in this way the power of the developed model, particularly at intermediate and higher energies.

In addition, the integral cross sections calculated from 10 eV to 20 keV for different combinations of potentials investigated here show clearly a good agreement except at lower energies (30 eV) where the effects of correlation, polarization and exchanges potentials are important, confirming in this way our concluding remarks in the case of the DCSs. These results can help us to select and recommend better expressions from that investigated. In this context, it is worth noting that new experimental works are highly encouraged in order to check the ability of our theoretical approach to model the electron elastic scattering process at the total and differential scales.

References

1. A. L. Broadfoot, Science 24 (1979) 979. [ Links ]

2. A. Cutgian, and S. Trajmar, Swar studies, and Inelastic electron-molecule collisions, Ed. L. C. Pichtford, and B. V. McKoy, (New-York 1987). [ Links ]

3. T. Sato, F. Shibata, and T. Goto, Chem. Phys. 108 (1986) 147. [ Links ]

4. E. Brüche Ann. Phys., Lpz. 1 (1929) 93. [ Links ]

5. W. R. Harshbarger, A. Skerbele, and E. N. Lassettre, J Chem. Phys. 54 (1971) 3784. [ Links ]

6. J. P. Bormberg, “unpublished work” as reported by Harshbarger et al. in [5]. [ Links ]

7. O. Sueoka, S. Mori, and Y. Katayama, J. Phys. B: At. Mol. Phys. 20 (1987) 3237. [ Links ]

8. C. Szmytkowskit, K. Maciag, G. Karwasz, and D. Filipović, J. Phys. B: At. Mol. Opt. Phys. 22 (1989) 525. [ Links ]

9. H. P. Pritchard, M. A. Lima, and V. McKoy, Phys. Rev. A 39 (1989) 2392. [ Links ]

10. T. W. Shyn, “unpublished work” as reported by Pritchard et al. in Ref. 9. [ Links ]

11. A. Zecca, G. P. Karwasz, and R. S. Brusa, Phys. Rev. A 45 (1992) 2777. [ Links ]

12. D. T. Alle, R. J. Gulley, S. J. Buckman, and M. J. Brunger, J. Phy. B: At. Mol. Opt. Phys 25 (1992) 1532. [ Links ]

13. G. García and F. Manero, J. Phys. B: At. Mol. Opt. Phys. 29 (1996) 4017. [ Links ]

14. W.M. Ariyasinghe, T. Wijeratne, P. Palihawadana, Nucl. Instrum. and Meth. in Phys. Res. B 217 (2004) 389. [ Links ]

15. N. C. Jones, D. Field, S. L. Lunt, and J.-P. Ziesel, Phys. Rev. A 78 (2008) 042714. [ Links ]

16. A. Lahmam-Bennani, A. Duguet, and H. F. Wellenstein, J. Phys. B 12 (1979) 461. [ Links ]

17. A. Lahrnarn-Bennani, A. Duguet , H. F. Wellenstein , and M. Rouault, J. Chem. Phys. 72 (1980) 6398. [ Links ]

18. A. Duguet , A. LahmamBennani, and M. Rouault , J. Chem. Phys. 78 (1983) 6595. [ Links ]

19. G. P. Karwasz , R. S. Brusa and A. Zecca , Riv. Nuovo Cimento 24 (2001) 1. [ Links ]

20. Y. Itikawa, J phys. Soc. Jpn 30 (1971) 835. [ Links ]

21. A Jain, J. Phys. B: At. Mol. Opt. Phys. 21 (1988) 905. [ Links ]

22. M. Ben Arfa and M. Tronc, J. Chim. Phys. 85 (1988) 889. [ Links ]

23. F. A. Gianturco, J. Phys. B: At. Mol. Opt. Phys. 24 (1991) 4627. [ Links ]

24. J. Yuan and Z. Zhang, Phys. Rev. A 45 (1992) 4565. [ Links ]

25. T. N. Rescigno, B. H. Lengsfield, C. W. McCurdy, and S. D. Parker, Phys. Rev. A 45 (1992) 7800. [ Links ]

26. K. N. Joshipura and P. M. Patel, J. Phys. B: At. Mol. Opt. Phys. 29 (1996) 3925. [ Links ]

27. Y. Liu, J. Sun, Z. Li, Y. Jiang, and L. Wan, Z. Phys. D 42 (1997) 45. [ Links ]

28. G. Garcia, F. Manero , Chem. Phys. Lett. 280 (1997) 419. [ Links ]

29. J. L. S. Lino, Rev. Mex. Fis. 51 (2005) 100. [ Links ]

30. H. Munjal and K. L. Baluja, Phys. Rev. A 74 (2006) 032712. [ Links ]

31. C. Limbachiya, M. Vinodkumar, and N. Mason, Phys. Rev. A 83 (2011) 042708. [ Links ]

32. I. Shimamura, Sci. Pap. Inst. Phys. Chem. Res. (Jpn.), 82 (1989) 1. [ Links ]

33. M. Hayashi, Electron Collision cross-sections, Handjournal on Plasma Material Science, 4 (1992) 9. [ Links ]

34. W. L. Morgan, International Symposium on Electron-Molecule Collisions and Swarms, Tokyo, edited by Y. Hatano et al., (Tokyo 1999). [ Links ]

35. R. Moccia, J. Chem. Phys. 40 (1964) 2176. [ Links ]

36. A. Messiah, Quantum Mechanics, Vol. 1 & 2; ch 9, 10, 19, J. Wiley & Sons Inc., New York (1968). [ Links ]

37. R. Moccia J. Chem. Phys. 40 (1964) 2164. [ Links ]

38. J. C. Slater, Quantum Theory of Atomic Structure. Vol. 2, McGraw-Hill, New-York, (1960). [ Links ]

39. H. Trygve, P. Jorgensen, and J. Olsen, In Molecular Electronic-Structure Theory, John Wiley Sons, Inc., (2000). [ Links ]

40. V. Spirko, H. G. Aa. Jensen, and P. Jorgensen , Chem. Phys. 144 (1990) 343. [ Links ]

41. M. H. Mittleman and K. M. Watson, Ann. Phys. 10 (1960) 286. [ Links ]

42. A. Salvat, A. Joblonski, and C. J. Powell, Comput. Phys. Commun. 165 (2005) 157. [ Links ]

43. N. T. Padial and D. W. Norcross, Phys. Rev. A 29 (1984) 1742. [ Links ]

44. W. J. Jr Carr and A. A. Maradudin, Phys. Rev. 133 (1964) 371. [ Links ]

45. B. Furness, I.E. McCarthy, J. Phys. B: At. Mol. Phys. 6 (1973) 2280. [ Links ]

46. M. H. Mittleman and K. M. Watson , Ann. Phys. 10 (1960) 286. [ Links ]

47. M. E. Riley and D. G. Truhlar, J. Chem. Phys. 63 (1975) 2182. [ Links ]

48. H. Aouchiche, F. Medegga, and C. Champion, Nucl. Instr. Meth. Phys. Res. B 333 (2014) 113. [ Links ]

49. F. Medegga , H. Aouchiche , High Energ. Chem. 51 (2017) 462. [ Links ]

50. L. M. Brescansin, L. E. Machado, M-T Lee, H. Cho and Y. S. Park, J. Phys B: At. Mol. Opt. Phys. 41 (2008) 185201. [ Links ]

51. H. Cho, R. P. McEachran, S. J. Buckman , and H. Tanaka, Phys. Rev. A 78 (2008) 034702 and references therein. [ Links ]

Received: January 13, 2018; Accepted: April 04, 2018

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License