Nonlinear ion-acoustic solitary waves in an electron-positron-ion plasma with relativistic positron beam
Sarma Ridip1, Misra Amar P2, Adhikary Nirab C3, †
Department of Mathematics, University of Science and Technology, Techno City, Khanapara, Baridua, 9th Mile, Ri-Bhoi, Meghalaya, India
Department of Mathematics, Siksha Bhavana, Visva-Bharati University, Santiniketan-731235, West Bengal, India
Physical Sciences Division, Institute of Advanced Study in Science and Technology, Vigyan Path, Paschim Boragaon, Garchuk, Guwahati-781035, Assam, India

 

† Corresponding author. E-mail: iasst@yahoo.co.in

Abstract

The propagation characteristics of nonlinear ion–acoustic (IA) solitary waves (SWs) are studied in thermal electron–positron–ion plasma considering the effect of relativistic positron beam. Starting from a set of fluid equations and using the reductive perturbation technique, we derive a Korteweg–de Vries (KdV) equation which governs the evolution of weakly nonlinear IA SWs in relativistic beam driven plasmas. The properties of the IA soliton are studied, and it is shown that the presence of relativistic positron beam significantly modifies the characteristics of IA solitons.

1. Introduction

The nonlinear propagation of solitary waves and shocks (SWS) in multicomponent plasmas has been widely studied for understanding the electrostatic disturbances in various plasma environments.[15] The study of electron–positron–ion (EPI) plasmas has been growing extensively due to their presence in the early universe, in the active galactic nuclei, in the pulsar magnetosphere, ionosphere, in the solar atmospheres, etc.[611] It is well known that when positrons are introduced into an electron–ion plasma, the response of the electrostatic waves changes significantly in comparison to the usual two component plasmas.[5,12,13] Therefore, it is worthwhile to study the nonlinear wave propagation in electron–positron–ion plasmas for understanding the dynamical behaviour of astrophysical and laboratory plasmas.[1420] In the latter, the authors have shown that the nonlinear waves in the plasmas having positrons behave differently. Misra et al.[5] reported that the nonlinear mode of propagation and interaction in an electron–positron plasma exhibits a remarkable feature which is quite distinctive from the linear mode. Also, in some earlier works, Popel et al.[14] have found that the maximum amplitude of ion–acoustic (IA) solitons could be considerably reduced when a substantial fraction of hot positrons was present in the system. Interestingly, in a plasma system when a source of beam is inserted, it makes the system prone to a well-known streaming instability. Recently, Shan et al.[20] showed that due to the increased speed of positron beam and super-thermal electrons, the region of existence, amplitude, and width of solitons can be modified. Again, very recently, Shan et al.[21] found that with the variation in positron beam speed and positron population, and nonlinearities of the plasma system can be predominantly changed. Therefore, the options to create intense sources of positrons capable of fuelling extensive antimatter related research need to be developed, and this gap or requirement motivated us to start our investigation on ion–acoustic solitary waves (IASWs) in EPI plasmas with a relativistic positron beam impact which is a new kind of study not reported elsewhere as far as our knowledge is concerned.

Some of the earlier works considered electron-ion and EPI plasmas, however, most of them were non relativistic in nature.[3,22] But when the particle velocities are comparable to the speed of light, relativistic effects should be taken into account. Many researchers in due course of time exhibited the importance of relativistic effects in the modification of plasma wave dynamics.[2329] In the ionospheric plasma system,[30] there is always a possibility of its interaction with highly intense charged particle beam coming from outer space. Therefore, in such conditions, the solitary wave in nonrelativistic plasmas may interact with relativistic charged particle beam. Based on all such observations, our brief communication has heralded on a model of EPI plasma to know the various effects of relativistic positron beam on soliton dynamics.

In the present paper, we investigate the propagation characteristics of ion–acoustic solitons in a relativistic beam driven plasma, and employ a new potent mathematical formalism known as the homogeneous balance method probably first generated by Wang et al. to show the solution of a Korteweg–de Vries (KdV) equation.[32] Here, Wang et al. introduced a new direct method called the homogeneous balance method to look for travelling wave solutions of nonlinear evolution equations with variable coefficients. It is seen that the method presents a wider applicability for handling many kinds of nonlinear evolution equations, such as high-dimensional equations, variable-coefficient equations, and differential-difference equations. The method is based on the homogeneous balance principle and linear ordinary differential equation (LODE) theory. Being concise and straightforward, this method has been applied to handle nonlinear partial differential equations in a direct manner with no requirement for initial/boundary conditions or initial trial function at the outset.

This paper is structured as follows. Section 2 describes the basic equations governing the plasma dynamics. The nonlinear evolution equation in the form of a Korteweg–de Vries equation is derived by using the reductive perturbation technique. In this section, the homogeneous balance method has been employed to execute the salient features of solitons. Section 3 analyzes thoroughly the results and the discussion parts. Finally, we conclude the results in Section 4.

2. Basic equations governing the plasma dynamics

We consider the nonlinear propagation of IASWs in an unmagnetized plasma consisting of thermal electrons and ions with a flow of relativistic positron beam. The basic equations governing the plasma dynamics can be written as

and the Poisson’s equation

where nj, vj, pj, qj, and mj respectively denote the number density, the fluid velocity, the thermal pressure, the charge, and the mass of j-th species particles (j = i, e, and b respectively stand for ions, electrons, and positron beams). Furthermore, kB is the Boltzmann constant, φ is electrostatic potential, γ is the relativistic factor, and c1 is the ratio between the speed of light in vacuum c and the IA speed cs.

The collective effects in electron–positron plasmas can be studied if the electron–positron time scale is much longer than the electron–positron plasma effects, which is typically inverse of the plasma frequency. In a relativistic regime, the processes of creation and annihilation of electron–positron pairs become important. Before studying the nonlinear properties of the electron–positron plasma, it is necessary to take a look at the electron–positron pair annihilation time. Based on the theory of Svensson[32] concerning the electron-positron pair equilibria in relativistic plasma, to neglect the annihilation effect, the following inequality must be satisfied:

where (j = e and b respectively stand for electrons and positron beam) is the inverse of the plasma frequency equal to and τann is the annihilation time. The annihilation time τann in the relativistic regime can be expressed by the following relation:

where στ (6.65 × 10−25 cm2) is the electron Thomson cross section, and ETh(kBTj/(mjc2)) is the normalised thermal energy. Also, = eE ≡ 0.5615 with E( 0.5772) is the Euler’s constant. Combining Eqs. (11) and (12) and using ne0 = np0 = n0, the pair annihilation condition becomes

For some illustration purposes, we have chosen the number density value n0 = 104 cm−3, which is typically found in astrophysical environment (viz. ionospheric auroral region). By using the number density value into Eq. (13), one can obtain 0.229003 ≫ 2.6 × 10−15, which indicates that the electron–positron pair annihilation can be ignored in our plasma model Eqs. (1)–(10).

Now, we normalize the physical quantities as follows: T = pi, X = pi/cs, Vj = vj/cs, Nj = nj/nj0, Φ = /kBTe, Pj = pj/nj0kBTj, where is the ion–plasma oscillation frequency, is the ion–acoustic speed, nj0 is the unperturbed number density, and Tj the temperature of j-th species particles. Thus, we recast Eq. (1) to Eq. (10) in dimensionless forms as

where , β2 = mb/mi, α = Tb/Te, and c1 = c/cs,

where , or , and

In order to derive the evolution equation for the propagation of IASWs in a relativistic positron beam plasma, we use the reductive perturbation technique in which the independent variables are stretched as λζ = ε1/2(XT), τ = ε3/2T, where λ is the phase velocity and ε is a small nonzero constant measuring the weakness of the dispersion.

Now, using the expansions, Ne = 1 + εNe1 + ε2Ne2 + . . ., Ni = 1 + εNi1 + ε2Ni2 + . . ., Nb = 1 + εNb1 + ε2Nb2 + . . ., Ve = ε Ve1 + ε2Ve2 + . . ., Vi = ε Vi1 + ε2Vi2 + . . ., Vb = Vb0 + ε Vb1 + ε2Vb2 + . . ., Pb = 1 + ε Pb1 + ε2Pb2 + . . ., and Φ = εΦ1 + ε2Φ2 + . . ., and by following the standard procedure and substituting the values of Ne, Ni, Nb, Ve, Vi, Vb, Pb, and Φ in Eqs. (14)–(23), and then equating the coefficients of first order and second order in ε, we get together the dispersion relation for the phase velocity , where , , and eliminating the second-order quantities, we obtain the following KdV equation for the nonlinear propagation of IASWs in a thermal electron–ion plasma with a relativistic positron beam.

Now, we apply the transformation

Then, we have

Integrating it with respect to ξ once yields

where M is an integration constant.

In order to get the soliton solution of Eq. (25), we use the homogeneous balance method which needs to assume a solution (ξ) in the form

where Q(ξ) is the solution of the equation Q′(ξ) = Q2(ξ) − Q(ξ).

Now substituting the value of Eq. (29) into Eq. (28), we balance the leading order of the nonlinear term with that of the linear term. This manipulation evaluates the exact numbers for N of the series obtained as N = 2. Correspondingly, a series with N = 2 terms become

where a0, a1a2 are constants to be determined such that a2 ≠ 0.

Here Q(ξ) = 1/(1 ∓ eξ) satisfies the following equation

From Eqs. (29) and (30), we obtain

By substituting Eqs. (30), (32), and (33) into Eq. (28) and rearranging all coefficients of Q0, Q1, Q2, and so on, and then using Mathematica 10 we get the values of a’s, M, and Ω as . Also taking the stability condition into account we obtain

where Φm = 3V/A and , which is the well-known solitary wave solution of the KdV Eq. (25).

3. Results and discussion

In what follows, we study the effects of the relativistic positron beam and other plasma parameters on the profiles of IASWs. The effects of nonlinearity become significant when the amplitude of the wave is sufficiently large. Nonlinear structures such as solitons and shocks are formed due to relative competition between nonlinearity, dispersion, and dissipation. When the effect of dissipation is small compared to nonlinearity and dispersion, the system can be supportive for the formation of solitons. The existence of a soliton solution with finite amplitude (Eq. (25)) depends on the behavior of the product AB, where A is the nonlinear term and B represents that of dispersion in Eq. (25). The solitary wave becomes unstable when AB > 0 and gives rise to a compressive solitary structure. So, some interesting results can be expected when we are examining positron beam velocity, positron beam concentration, ion and positron temperature, and mass ratio along with its amplitude, width, and dispersion relation in our electron–positron–ion model. We have considered an external positron beam which is relativistic in nature under the effect of adiabatic responses of the positron beam to contribute additional sources of energy to the nonlinear behaviour of the usual electron positron ion plasma dynamical system.

We numerically examine in Fig. 1 the variation of the phase velocity Δ, plotted with respect to the positron concentration μb along with variable ion temperature σ. The other plasma parameters are α = 0.2, c1 = 103, and . Figure 1 shows that the dispersion properties of the present model have strong dependence on the plasma temperature. At elevated ion temperature σ, dispersion is triumphant and consequently the medium becomes more favourable for the propagation of compressive solitary waves, and the phase velocity increases as the ion temperature σ increases. The findings in Fig. 2, where variation of the phase velocity λ is plotted with respect to the positron concentration μb along with variable positron temperature α with other plasma parameters as c1 = 103, , and σ = 0.1, are in similar lines with the outcomes in Fig. 2. In Fig. 3, the variation of the phase velocity λ is plotted with respect to the positron concentration μb along with variable relativistic factor with other plasma parameters as c1 = 103, α = 0.2, and σ = 0.1. Here, from the investigation, it can be observed that the phase velocity λ is influenced by the relativistic effect even if is very small or considered fixed as shown in Figs. 1 and 2. When the ion temperature σ and positron temperature α are kept constant, phase velocity λ increases as increases. Thus from the present investigation it can be emphasized that the behaviour of ion acoustic solitary waves propagating in a relativistic electron–positron–ion plasma model is strongly dependent on the ion and positron temperature, the mass ratio, and the relativistic effect.

Fig. 1. (color online) The phase velocity λ is plotted with respect to the positron concentration μb along with variable ion temperature σ, where σ is 0.1, 0.3, and 0.5 for blue dotted, red dashed, and green lines, respectively, and other plasma parameters are c1 = 103, , and α = 0.2.
Fig. 2. (color online) The phase velocity λ is plotted with respect to the positron concentration μb along with variable positron temperature α, where α is 0.2, 0.4, and 0.6 for blue dotted, red dashed, and green bold lines, respectively, and other plasma parameters are c1 = 103, , and σ = 0.1.
Fig. 3. (color online) The phase velocity λ is plotted with respect to the positron concentration μb along with variable relativistic factor , where is 0.2, 0.4, and 0.6 for blue dotted, red dashed, and green bold lines, respectively, and other plasma parameters are c1 = 103, α = 0.2, and σ = 0.1.

Next, we examine the properties of the soliton solution, namely its amplitude and width for different values of the system parameters. The results are shown in Figs. 49. Figure 4 represents the variation of the width plotted with respect to the positron concentration μb along with the increase in positron temperature, and as the positron temperature increases, the width increases. At higher positron temperature, i.e., at α = 0.4 and α = 0.6, it can be observed that the level of the increase in width is greater than that at lower temperature. Figure 5 exhibits the variation in width plotted along with variable positron temperature α. It shows that the width decreases with the increase in the positron concentration. On the other hand, in Fig. 6, the variation of width is plotted along with variable ion temperature σ. The results obtained in Figs. 5 and 6 seem to differ with the results obtained in non-relativistic cases.[17] It may be due to the fact that in relativistic electron–positron–ion plasma, increasing the positron density enhances the depth of the potential well, resulting in narrower solitons. Also, when the relativistic factor is kept constant, the dispersion effect starts to dominate over the nonlinearity with increasing values of ion temperature ratio σ. Since the width of the soliton is inversely proportional to the dispersion effect, the decrease in the ion temperature ratio σ results in the increase of the width. Figures 7, 8, and 9 seem to agree quite well when compared with the results obtained from Figs. 4, 5, and 6. In Figs. 8 and 9, it can be observed that the amplitude increases with increasing μb. The physical reason behind this is may be due to the consideration of an external positron beam relativistic in nature under the effect of adiabatic responses, which contributes additional sources of energy to the nonlinear behaviour of the usual electron positron ion plasma dynamical system. This similar observation is also congruent with the result exhibited in Fig. 7. Also, an interesting fact can be observed in Fig. 9, that is, the graphs of the three regions (μb = 0.5, 0.7, and 0.9) tend to come closer to each other at higher ion temperature. This result is quite opposite to the character as observed for positron temperature in Fig. 8. This seems to be a quite interesting result, in view of its application in further study of astrophysical plasma regions.

Fig. 4. (color online) The soliton width Δ is plotted with respect to the positron concentration μb along with variable positron temperature α, where α is 0.2, 0.4, and 0.6 for blue dotted, red dashed, and green bold lines, respectively, and other plasma parameters are c1 = 103, , σ = 0.1, β1 = 0.005, and β2 = 0.0054.
Fig. 5. (color online) The soliton width Δ is plotted with respect to the positron temperature α along with variable positron concentration μb, where μb is 0.5, 0.7, and 0.9 for blue dotted, red dashed, and green bold lines, respectively, and other plasma parameters are c1 = 103, , σ = 0.1, β1 = 0.005, and β2 = 0.0054.
Fig. 6. (color online) The soliton width Δ is plotted with respect to the ion temperature σ along with variable positron concentration μb, where μb is 0.5, 0.7, and 0.9 for blue dotted, red dashed, and green bold lines, respectively, and other plasma parameters are c1 = 103, , α = 0.2, β1 = 0.005, and β2 = 0.0054.
Fig. 7. (color online) The soliton amplitude ϕm is plotted with respect to the positron concentration μb along with variable positron temperature α, where α is 0.2, 0.4, and 0.6 for blue dotted, red dashed, and green bold lines, respectively, and other plasma parameters are c1 = 103, , σ = 0.1, β1 = 0.005, and β2 = 0.0054.
Fig. 8. (color online) The soliton amplitude ϕm is plotted with respect to the positron temperature α along with variable positron concentration μb, where μb is 0.5, 0.7, and 0.9 for blue dotted, red dashed, and green bold lines, respectively, and other plasma parameters are c1 = 103, , σ = 0.1, β1 = 0.0050, and β2 = 0.0054.
Fig. 9. (color online) The soliton solution is plotted with respect to the ion temperature σ along with variable positron concentration μb, where μb is 0.5, 0.7, and 0.9 for blue dotted, red dashed, and green bold lines, respectively, and other plasma parameters are c1 = 103, , α = 0.2, β1 = 0.0050, and β2 = 0.0054.
4. Conclusion

We have investigated the propagation characteristics of ion–acoustic solitary waves in an unmagnetized thermal electron–ion plasma with a flow of relativistic positron beam. Using the reductive perturbation technique, the KdV equation is derived which governs the evolution of weakly nonlinear ion–acoustic solitons in relativistic beam driven plasmas. The effects of the plasma parameters, namely the ion to electron temperature ratio, the beam to electron temperature ratio, and the relativistic parameter due to the flow of positron beam on the profiles of the ion–acoustic solitons, are studied. It is found that the propagation characteristics of ion–acoustic solitons are significantly modified by these plasma parameters. To conclude, the results should be useful for understanding the properties of ion–acoustic solitons that may be relevant in laboratory and space plasmas.

Reference
[1] Zhao M Shan X Niu S Chen X 2017 Chin. Phys. 26 093103
[2] Gan C Y Xiang N Yu Z 2016 Chin. Phys. Lett. 33 085203
[3] Adhikary N C Deka M K Dev A N Sarmah J 2014 Phys. Plasmas 21 083703
[4] Ye C E Zhang W G 2010 Acta Phys. Sin. 59 5229 in Chinese
[5] Misra A P Roy Chowdhury A 2003 Chaos, Solitons and Fractals 15 801
[6] Abdelwahed H G El-Shewy E K Mahmoud A A 2017 Chin. Phys. Lett. 34 035202
[7] Alam M S Hafez M G Talukder M R Ali M H 2017 Chin. Phys. 26 095203
[8] Dev A N 2017 Chin. Phys. 26 025203
[9] Jian M J Rong Y J Ying L C 2012 Acta Phys. Sin. 61 020206 in Chinese
[10] Surko C M Leventhal M Passner A 1989 Phys. Rev. Lett. 62 901
[11] Surko C M Murphy T J 1990 Phys. Fluids A 2 1372
[12] Greaves R G Tinkle M D Surko C M 1994 Phys. Plasmas 1 1439
[13] Sarri G Dieckmann M E Kourakis I Di Piazza A Reville B Keitel C H Zepf M 2015 J. Plasma Phys. 81 455810401
[14] Popel S I Vladimirov S V Shukla P K 1995 Phys. Plasmas 2 716
[15] Nejoh Y N 1996 Phys. Plasmas 3 1447
[16] Saleem H 2006 Phys. Plasmas 13 34503
[17] Esfandyari-Kalejahi A Akbari-Moghanjoughi M Saberian E 2010 Plasma Fusion Res. 5 045
[18] Chatterjee P Ghosh U Roy K Muniandy S V Wong C S Sahu B 2010 Phys. Plasmas 17 122314
[19] Chatterjee P Saha T Muniandy S V Wong C S Roychoudhury R 2010 Phys. Plasmas 17 012106
[20] Shan S A El-Tantawy S A Moslem W M 2013 Phys. Plasmas 20 082104
[21] Shan S A El-Tantawy S A 2016 Phys. Plasma 23 072112
[22] Adhikary N C Mishra A P Deka M K Dev A N 2017 Phys. Plasmas 24 073703
[23] Dev A N Deka M K Sarma J Saikia D Adhikary N C 2016 Chin. Phys. 25 105202
[24] Gill T S Singh A Kaur H Saini N S Bala P 2007 Phys. Lett. 361 364
[25] Gill T S Bains A S Saini N S 2009 Can. J. Phys. 87 861
[26] Han J Du S Duan W 2008 Phys. Plasmas 15 112104
[27] Shah A Haque Q Mahmood S 2011 Astrophys. Space Sci. 335 529
[28] Hafez M G Talukder M R Sakthivel R 2016 Indian J. Phys. 90 603
[29] Hafez M G Talukder M R 2015 Astrophys. Space Sci. 359 27
[30] Gsponer A 2004 arXiv: physics/0409157v3[physics.plasm-ph]
[31] Wang M L 1995 Phys. Lett. 199 169
[32] Svensson R 1982 Astrophys. J. 258 335