Skip to content
Publicly Available Published by De Gruyter August 10, 2020

Resonant interactions between the fundamental and higher harmonic of positron acoustic waves in quantum plasma

  • Jit Sarkar ORCID logo , Swarniv Chandra EMAIL logo and Basudev Ghosh

Abstract

In this paper, we study the second harmonic generation and its interaction with the fundamental mode in a magnetised dense positron-ion plasma interacting with laser pulses. It has been shown that different harmonics propagate with different phase velocities. The gradual evolution of the fundamental wave into higher harmonics is studied, and the conversion efficiency is calculated. Dependence of conversion efficiency on wavenumber shifts and the applied magnetic field has also been examined.

1 Introduction

In recent years, particle acceleration by the interaction of high-intensity laser pulse with plasma has become of great interest because of its tremendous applicability in fundamental research and industrial use. Such laser plasmas have vast applications in medical and biological fields. The development for trapping and storing positrons [1], [2] allows a large number of low-temperature positrons to form a plasma. The physical properties of such plasma may vary from the conventional ones, but the collective behaviours shown by them are complimentary to electrons in plasmas.

Positrons are fundamentally important in the study of physics because of their recombination with electron to form neutral plasmas with dynamical symmetry between charged particles [3], [4]. Such positron-hydrogen interactions have been often observed in astrophysical environments. Many wave modes and associated instabilities are found to occur in hydrogen plasmas and solar system plasmas with an excess of positrons [5], [6], [7].

Many a time high energy electron beam propagates through a dense positron plasma. A part of them recombines releasing a tremendous amount of energy that causes newer wave modes which may or may not survive nonlinear Landau damping. Under such conditions, oblique propagation of waves is conventional [7]. If there is a magnetic field present in such a homogeneous plasma environment, both electrons and positrons may often transform from a neutral plasma to two oppositely charged species. Such a flip of state can be obtained by meticulously varying the particle densities. The easy availability of low energy positron is helpful in experimental studies of magnetic confinement, positron-atom and positron-molecule interaction, as well as often fundamental particles [8]. Because of the plasma ray astronomical studies, such interactions are of immense value [9], [10]. Slow positrons can be easily obtained from 22Na isotope (with half-life 2.6 years) used along with a tungsten moderator. This 22Na isotope emits positron with an energy of 545 KeV which can be slowed down to 2 eV (nonrelativistic) by a tungsten crystal [11], [12]. Now, that such a positron plasma has been contained by any physical mean, there exists a plasma vacuum boundary interface. When a wave is generated in such a plasma due to nonlinear interactions, higher-order harmonics are often found to originate. The resonant interactions between the original wave and the newly generated harmonics give rise to damping-like effects, as well as energy conversion between them. In order to study plasma harmonic generation by free particles, without the competing effect of atomic harmonic generation by bound particles, a preformed plasma is required [13].

In the present paper, we consider a laser pulse with a phase velocity comparable to the velocity of light in a dense magnetised positron plasma with low energy. We further study the conversion efficiency of the harmonic generated and the energy distribution arising out of such resonant interactions.

The paper is organized in the following way: in section 2, starting from the fundamental equations, we derive the linear dispersion relation of second harmonics. In the next section, we study the generation and interaction of the second harmonics with the fundamental mode. Here, we study the conversion efficiency and resonant interaction with associated energy distributions. Finally, we conclude with some remark on the scope for further study and its possible applications.

2 Basic equations and model

We consider a magnetised quantum plasma consisting of positrons and negatively charged ions occupying half-space (x > 0) and bounded by vacuum (x < 0), as shown in Figure 1. Wave is travelling along the x-axis, and the plasma is considered collision less (almost) due to the Pauli blocking mechanism. Electrostatic waves are considered to be travelling in such a completely degenerate dense plasma. The following basic equations govern the dynamics of such a plasma:

(1)npt+(npγup)=0
(2)(t+up)(upγ)=eEm+empup×(B+b)1mpnpPp+22mp(2npnp)
(3)E=4πe(npZieni)

where ni = density of the negative ions, Zie = charge of the negative ions. The Laplace equation in vacuum is given as follows:

(4)ϕν=0
Figure 1: Schematic of polarized electromagnetic wave propagating in homogeneous plasma.
Figure 1:

Schematic of polarized electromagnetic wave propagating in homogeneous plasma.

Here, up, Pp, , E, B, b, ni and ϕν are the fluid velocity, degeneracy pressure, the Planck constant, electric field, radiation magnetic field, applied magnetic field, ion density and electric potential in vacuum, respectively. The last term of Eq. (2) is Bohm potential term. According to Chandrasekhar [14], one can write the degeneracy pressure in the following form:

(5)Pdegeneracy=πme4c53h3[R(2R23)1+R2+3sinh1R]

where ‘c’ is the speed of light in vacuum and R = PF/mc (PF Fermi relativistic momentum).

Now, the basic set of Eqs. (1)–(4) is normalised by employing the following scheme (for simplicity, we will confine ourselves in one dimension in the linear analysis) xxωpcs, ttωp, ωjωjωp, ϕeϕ2kBTF, npnpnp0, ninini0 and ueuecs. Here, ωp=4πne0e2me is the plasma oscillation frequency and cs=2kBTFme is the positron acoustic speed. We obtained the following equations:

(6)npt+npupx=0
(7)[t+upx]up=ϕxωcupGpnpx+H22x[1np2npnp2]
(8)2ϕx2=(npZieni)
(9)2ϕνx2=0

where ωc is the cyclotron frequency normalised with respect to plasma oscillation frequency, H is the quantum diffraction parameter defined as H=ωp/2kBTFe, (TFe Fermi temperature) and Gp=mpc26kBTFe(R02/1+R02)[R0(n/n0)(1/3)]

3 Linear analysis

We assume that every field quantity (f) has the following form:

(10)f=f0+f(x)exp[i(kzωt)]+c.c

Using Eq. (10) in Eqs. (6)–(9) and solving the resulting equation under the boundary condition that electric potential is continuous across the interface, we obtain the following linear dispersion relation in one dimension:

(11)ω=12(1+k2ΛR03+H2k24)

where Λ=mec22kBTFeZie. The normalised frequency (ω) depends on the wavenumber (k), electron Fermi temperature (TFe), Quantum diffraction parameter (H) and relativistic degeneracy factor (R0).

4 Generation of second harmonic

Now, consider a linearly polarized (along the x-axis) laser pulse propagating along the z-direction in a uniform quantum plasma containing positive and negative ions. The plasma is embedded in a transverse magnetic field in y-direction. The electric field component of the laser beam is given as follows:

(12)El=12e^xE0ei(k0zω0t)+c.c

where ω0=ωp1μ2, µ is the refractive index of plasma, ωp is plasma frequency and e^x is a unit vector along x-direction. As the wave propagates in the plasma, transverse current densities at double the fundamental frequency are generated due to nonlinear laser-plasma interactions. Now, we can write the corresponding electric fields for the fundamental and double fundamental frequencies (ω0 and 2ω0) as follows:

(13)E1=12e^xE1ei(k1zω0t)+c.c
(14)E2=12e^xE2ei(k2z2ω0t)+c.c

Here, E1 and E2 are the amplitudes; k1 and k2 are the propagation vectors which are given as follows:

(15)k1=ω0cμ1
(16)k2=2ω0cμ2

where μ1 and μ2 are the corresponding wave refractive indices for the plasma medium.

The wave equation governing the propagation of laser beam in the plasma is given by as follows:

(17)(21c22t2)E=4πc2Jt

The total electric field vector due to interaction is given as follows:

(18)E=E1+E2

The plasma current density due to positrons is given as follows:

(19)J=enup

Using reductive perturbation expansion and following the standard procedure, we obtain the first-order transverse and longitudinal velocities as follows:

(20)upx(1)=i2[cλ1ω02ω02ωc2ei(k1zω0t)+2cλ2ω044ω02ωc2ei(k2z2ω0t)λ2(2ω0+iωc)n(1,2)(4ω02ωc2)λ1(ω0+iωc)Θq1n(1,1)(ω02ωc2)]+constant
(21)upz(1)=12[cλ1ω02ω02ωc2ei(k1zω0t)+4cλ2ω044ω02ωc2ei(k2z2ω0t)+λ1(ω0iωc)Θq2n(1,1)(ω02ωc2)+λ2(2ω0iωc)n(1,2)(4ω02ωc2)]+constant
(22)λ1=eE1mcω0λ2=eE2mcω0ωc=ebmcΘq1={VF2n0+24m21n0k12}k1Θq2={VF2n0+24m21n0k22}k2}

While obtaining above velocity components (Eqs. 20 and 21), we assumed the first-order perturbed plasma density n(1) containing only the contribution from the fundamentals and first harmonic terms n(1)=n(1,1)+n(1,2), where n(1,1) and n(1,2) vary as n1ei(k1zω0t) and n2ei(k2z2ω0t). Now, substituting perturbed positron density into continuity equation, we derive the first-order plasma positron density as follows:

(23)n(1)=12[(cωcω02ω0Θq1n1+2iωcΘq1n1)k1n0λ1(ω03ωc2ω0)ei(k1zω0t)+(cωcω04ω0Θq2n1+2iωcΘq2n2)k2n0λ24(ω03ωc2ω0)ei(k2z2ω0t)]+constant

in which

n1=ck1n0ωcω022[ω02(ω02ωc2)+n0ω0k1(ω0iωc)Θq1]n2=2ck2n0ωcω022[4ω02(4ω02ωc2)+n0ω0k2(ω0iωc)Θq2]

Similarly, we can calculate the second ordered transverse and longitudinal velocities as follows:

(24)upx(2)=[i4c2k1ω02ωc(ω024ωc2)λ12(ω02ωc2)(4ω02ωc2)+λ12k1{p2Θq2n12p1Θq1n1}(ω02ωc2)(4ω02ωc2)]e2i(k1zω0t)
(25)upz(2)=λ12k1c2{3ω03ωc2+(ω03ωc2ω0)(4ω022ωc2)}4(ω02ωc2)(4ω02ωc2)e2i(k1zω0t)[2λ12k1(ω0ωciωc2)(cω02+2ω022ωc2)Θq1n1(ω02ωc2)(4ω02ωc2)+2(ω02iωcω0)(cωc2ωc2+2ω02)λ12k1Θq12n12(ω02ωc2)(4ω02ωc2)]e2i(k1zω0t)

The second-order x-component of velocity is generated due to uniform magnetic field and reduces to zero in its absence. However, the z-component of velocity is due to the magnetic vector of the radiation field.

Substituting first-order densities and second-order velocities in the continuity equation, we get the following equation:

(26)n(2)=[3λ12k12n0c2ω02ωc24(4ω02ωc2)(ω02ωc2)+λ12k12n0c2(ω02ωc2)(4ω022ωc2)4(4ω02ωc2)(ω02ωc2)n0k12ω0(4ω02ωc2)(ω02ωc2){ωc(ω0iωc)(ω02ωc2+cω022)Θq1n1+iω0(ω0iωc)(ω02ωc2+cωc22)Θq12n12}]e2i(k1zω0t)

The perturbed velocities and densities are used to obtain the transverse current density from Eq. (19)

(27)Jx={ien0cλ1ω022(ω02ωc2)en0λ1(ω0+iωc)n1Θq1ω02ωc2}ei(k1zω0t){iecn0ω02λ2(4ω02ωc2)en0λ2(2ω0+iωc)n2Θq24ω02ωc2}ei(k1z2ω0t)+βe2i(k1zω0t)+constant

where

β=ek1n0(4ω02ωc2)(ω02ωc2)[i4λ12c2ω02ωc(ω02+ωc2)λ12n2Θq1{p1+(c(4ω02ωc2)2)×(i(ω0iωc)ω0(ω0+iωc)ωc)}+{(ω02+ωc2)(4ω02ωc2)ω0p2}](Θq12n12ω0)
p1=ω0(ω0iωc){2i(ω02ωc2)+ic(ω02+ωc2)+(ω0+iωc)(ω02ωc2+cω0ωc)}p2=2ω0(ω02+ωc2)+iωc(ω0iωc)2

Using Jx, we obtain the first and second harmonic dispersion relations as follows:

(28)c2k12=ω02ω02ωp2(ω02ωc2)
(29)c2k22=4ω024ω02ωp2(4ω02ωc2)

To obtain the amplitude of the second harmonic term, we substitute the current density (Eq. 19) in the wave Eq. (17), equate the second harmonic terms and then after proper manipulation, we obtain the following equation:

(30)λ2(z)z=3iλ12k12cωp2ω02ωc(ω02+ωc2)k2(ω02ωc2)(4ω02ωc2)eiΔkz
(31)λ2=[3λ12ωcωP28cω02{{1ωp2ω02(1ωc2ω02)}1/2(1+ωc2ω02){1ωp24ω02(1ωc24ω0)1}1/2(1ωc2ω02)2(1ωc24ω02)}4ωp2λ1k1n1Θq1c3k2(4ω02ωc2)(ω02ωc2)2Γ]eiΔkz/2sin(Δkz/2)Δk

where, wavenumber shift Δk is given as follows:

(32)Δk=k22k1

The second harmonic conversion efficiency is given as follows:

(33)η=μ1|λ22|μ2|λ12|
(34)η=9λ12ωc2ωp416c2ω04[{1ωp2ω02(1ωc2ω02)}1/2(1+ωc2ω02){1ωp24ω02(1ωc24ω0)1}1/2(1ωc2ω02)2(1ωc24ω02)+4ωp4k1n1Θq1Γ2c6k2(ω02ωc2)4(4ω02ωc2)2]sin2(Δkz/2)Δk
(35)ηmax=9λ12ωc2ωp416c2ω04[{1ωp2ω02(1ωc2ω02)}1/2(1+ωc2ω02){1ωp24ω02(1ωc24ω0)1}1/2(1ωc2ω02)2(1ωc24ω02)+4ωp4k1n1Θq1Γ2c6k2(ω02ωc2)4(4ω02ωc2)2]

in which

Γ=[n1Θq1(ω02+ωc2)(4ω02ωc2)ω0p2+p1+c2(4ω02ωc2){i(ω0iωc)ω0(ω0+iωc)ωc}]

5 Results

We have calculated the conversion efficiency of the fundamental into the second harmonic. We have studied the relationship between the frequency and wavenumber shift (Δk) and its dependence on the applied magnetic field. The corresponding results are shown graphically in Figures 2 and 3. Figure 4 shows the dependence of conversion efficiency on the wavenumber shift for various values of the applied magnetic field. It shows that conversion efficiency is large for small wavenumber shift. Figure 5 shows the conversion efficiency length of wavenumber shift (ηzΔk) plots. The figure is symmetric and there is gradual decay along the length and flip in directionality due to the phase factor of the propagating wave. As we know harmonics carry energy with it, the energy exchange during such a harmonic conversion becomes important to study losses in the dispersive media. Figure 6 shows the variation of maximum conversion efficiency (ηmax) with the magnetic field. It is evident that for low magnetic field strength, ηmax increases very slowly with magnetic field, but for strong field, it increases rapidly.

Figure 2: : ω0${\omega }_{0}$ versus Δk${\Delta}k$ with a varying magnetic field.
Figure 2:

: ω0 versus Δk with a varying magnetic field.

Figure 3: A contour plot of ω0${\omega }_{0}$ in (Δk−ωc)$\left({\Delta}k-{\omega }_{c}\right)$ plane.
Figure 3:

A contour plot of ω0 in (Δkωc) plane.

Figure 4: η versus Δk${\Delta}k$ with varying ωcω0$\frac{{\omega }_{c}}{{\omega }_{0}}$.
Figure 4:

η versus Δk with varying ωcω0.

Figure 5: Variation of conversion efficiency (η) with (z−Δk)$\left(z-{\Delta}k\right)$.
Figure 5:

Variation of conversion efficiency (η) with (zΔk).

Figure 6: Variation of maximum conversion efficiency ηmax${\eta }_{max}$ with ωcω0$\frac{{\omega }_{c}}{{\omega }_{0}}$.
Figure 6:

Variation of maximum conversion efficiency ηmax with ωcω0.

6 Conclusions

In this work, analytical study of second harmonic generation by a polarized laser pulse in homogeneous dense positron-ion plasma has been presented. The Lorentz force acting on plasma positrons introduces changes in relativistic mass and causes positron density perturbations, leading to change in the propagation characteristics of the laser beam. The wave equation governing the evolution of second harmonic is set up by using nonlinear current density arising due to the fundamental radiation. The slowly varying second harmonic amplitude and the conversion efficiency are obtained by solving the wave equation. This present study reveals that the maximum conversion efficiency increases with magnetic field strength and decreases with wavenumber shift. Our finding will help in the study of laser-plasma interaction, as well as laser beam interacting, with any nonlinear optical medium.


Corresponding author: Swarniv Chandra, Department of Physics, Government General Degree College at Kushmandi, Kolkata 733121, West Bengal, India, E-mail:

Acknowledgement

The authors are grateful to the anonymous referees for their constructive criticism which led to the improvement of this article. Authors would like to thank the physics departments of Jadavpur University and Government General Degree College at Kushmandi for providing facilities to carry out this work. Jit Sarkar would also like to thank Jyotirmoy Goswami for support and inspiration.

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: None declared.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

[1] R. G. Greaves, M. D. Tinkle, and C. M. Surko, “Creation and uses of positron plasmas,” Phys. Plasmas, vol. 1, no. 5, pp. 1439–1446, 1994, https://doi.org/10.1063/1.870693.Search in Google Scholar

[2] T. J. Murphy and C. M. Surko, “Positron trapping in an electrostatic well by inelastic collisions with nitrogen molecules,” Phys. Rev., vol. 46, no. 9, p. 5696, 1992, https://doi.org/10.1103/physreva.46.5696.Search in Google Scholar

[3] R. J. Gould, “Energy loss of fast electrons and positrons in a plasma,” Physica, vol. 60, no. 1, pp. 145–154, 1972, https://doi.org/10.1016/0031-8914(72)90227-3.Search in Google Scholar

[4] R. J. Gould, “Direct positron annihilation and positronium formation in thermal plasmas,” Astrophys. J., vol. 344, pp. 232–238, 1989, https://doi.org/10.1086/167792.Search in Google Scholar

[5] H. Alfvén, Cosmic plasma, Dordrecht, Holland, Reidel Publ. Co., 1981.10.1007/978-94-009-8374-8Search in Google Scholar

[6] V. Tsytovich and C. B. Wharton, “Laboratory electron-positron plasma-a new research object,” Comments Plasma Phys. Contr. Fusion, vol. 4, no. 4, pp. 91–100, 1978.Search in Google Scholar

[7] N. Iwamoto, “Collective modes in nonrelativistic electron-positron plasmas,” Phys. Rev. E, vol. 47, no. 1, p. 604, 1993, https://doi.org/10.1103/physreve.47.604.Search in Google Scholar

[8] Y. J. Gu, O. Klimo, S. Weber, and G. Korn, “High density ultrashort relativistic positron beam generation by laser-plasma interaction,” New J. Phys., vol. 18, no. 11, 2016, Art no. 113023, https://doi.org/10.1088/1367-2630/18/11/113023.Search in Google Scholar

[9] M. Amoretti, C. Amsler, G. Bonomi, et al., “Positron plasma diagnostics and temperature control for antihydrogen production,” Phys. Rev. Lett., vol. 91, no. 5, 2003, Art no. 055001, https://doi.org/10.1103/physrevlett.91.055001.Search in Google Scholar

[10] R. Funakoshi, M. Amoretti, G. Bonomi, et al., “Positron plasma control techniques for the production of cold antihydrogen,” Phys. Rev. A, vol. 76, no. 1, 2007, Art no. 012713, https://doi.org/10.1103/physreva.76.012713.Search in Google Scholar

[11] E. Gramsch, J. Throwe, and K. G. Lynn, “Development of transmission positron moderators,” Appl. Phys. Lett., vol. 51, no. 22, pp. 1862–1864, 1987, https://doi.org/10.1063/1.98495.Search in Google Scholar

[12] K. G. Lynn, B. Nielsen, and J. H. Quateman, “Development and use of a thin film transmission positron moderator,” Appl. Phys. Lett., vol. 47, no. 3, pp. 239–240, 1985, https://doi.org/10.1063/1.96231.Search in Google Scholar

[13] X. Liu, D. Umstadter, E. Esarey, and A. Ting, “Harmonic generation by an intense laser pulse in neutral and ionized gases,” IEEE Trans. Plasma Sci., vol. 21, no. 1, pp. 90–94, 1993, https://doi.org/10.1109/27.221106.Search in Google Scholar

[14] S. Chandrasekhar, “An Introduction to the Study of Stellar Structure, Dover Publications, INC.,” 1957, https://books.google.co.in/books?id=VwPLAgAAQBAJ&lpg=PP1&pg=PP3#v=onepage&q&f=false.Search in Google Scholar

Received: 2020-01-11
Accepted: 2020-07-20
Published Online: 2020-08-10
Published in Print: 2020-10-25

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/zna-2020-0012/html
Scroll to top button