Next Article in Journal / Special Issue
Delayed Discharge Bridging Two Sputtering Modes from Modulated Pulsed Power Magnetron Sputtering (MPPMS) to Deep Oscillation Magnetron Sputtering (DOMS)
Previous Article in Journal
Generation of a Metal Ion Beam Using a Vacuum Magnetron Discharge
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrostatic Dust-Acoustic Rogue Waves in an Electron Depleted Dusty Plasma

1
Faculty of Science and Information Technology, Daffodil International University, Dhaka 1207, Bangladesh
2
Department of Physics, Jahangirnagar University, Savar, Dhaka 1342, Bangladesh
3
Plasma Physics Division, Atomic Energy Centre, Dhaka 1000, Bangladesh
*
Author to whom correspondence should be addressed.
Submission received: 15 March 2021 / Revised: 11 April 2021 / Accepted: 12 April 2021 / Published: 16 April 2021
(This article belongs to the Special Issue Feature Papers in Plasma Sciences)

Abstract

:
The formation of gigantic dust-acoustic (DA) rouge waves (DARWs) in an electron depleted unmagnetized opposite polarity dusty plasma system is theoretically predicted. The nonlinear Schrödinger equation (NLSE) is derived by employing the reductive perturbation method. It is found that the NLSE leads to the modulational instability (MI) of DA waves (DAWs), and to the formation of DARWs, which are caused by to the effects of nonlinearity and dispersion in the propagation of DAWs. The conditions for the MI of DAWs and the basic properties of the generated DARWs are numerically identified. It is also seen that the striking features (viz., instability criteria, amplitude and width of DARWs, etc.) of the DAWs are significantly modified by the effects of super-thermality of ions, number density, mass and charge state of the plasma species, etc. The results obtained from the present investigation will be useful in understanding the MI criteria of DAWs and associated DARWs in electron depleted unmagnetized opposite polarity dusty plasma systems like Earth’s mesosphere (where the D-region plasma could suffer from electron density depletion), cometary tails, Jupiter’s magnetosphere, and F-ring of Saturn, etc.

1. Introduction

Opposite polarity (OP) dusty plasma (OPDP) is demonstrated as fully ionized gas consisting of massive OP dust grains, ions, and electrons in the presence of electrostatic and gravitational force fields, and has been identified in astrophysical environments, viz., Earth polar mesosphere [1], cometary tails [2,3,4], Jupiter’s magnetosphere [5,6], solar system [3], and laboratory devices [7,8,9,10]. First, Rao et al. [11] traced how the presence of massive dust grains alters the dynamics of dusty plasma medium (DPM), and also theoretically predicted a new kind of low-frequency dust-acoustic (DA) waves (DAWs). The distinction of low-frequency DAWs from ion-acoustic waves (IAWs) was experimentally confirmed by Barkan et al. [12] in the DPM. The experimental identification of DAWs in the DPM has mesmerized many plasma physicists to investigate numerous modern electrostatic eigen-modes, viz., dust-drift waves [13], DA solitary waves (DA-SWs) [14], DAWs [15,16,17], dust lattice waves [18], DA shock waves (DA-SHWs) [19], and dust-ion-acoustic waves (DIAWs) [20], etc.
The mechanism of electron depletion is considered as the maximum number of electrons that are inserted onto the massive negative dust grains from the background of DPM during the dust charging process, and the medium in which the electron depletion process occurs is known as electron depleted DPM (EDDPM) [21,22,23]. The laboratory observation of low phase velocity DAWs is associated with significant depletion of the electron number density, suggesting that the wave dynamics is governed by the inertia of the dust fluid and the pressure of inertialess ions only [12,24,25,26]. The electron depletion process can also be observed in astrophysical environments, viz., Earth’s mesosphere (where the D-region plasma could suffer from electron density depletion) [27], cometary tails [4], Jupiter’s magnetosphere [5,6], and F-ring of Saturn [28], etc. Shukla and Silin [20] examined DIAWs in an unmagnetized collisionless EDDPM. Ferdousi et al. [19] considered a two-component plasma system containing inertialess positively charged ions and inertial negatively charged dust grains to investigate DA-SHWs, and found that the model supports both positive and negative electrostatic potentials. Mamun et al. [21] considered a two-component EDDPM to investigate the propagation of a nonlinear solitary pulse, and found that the height of the negative potential pulse increases with the number density of ions. Sahu and Tribeche [29] studied electrostatic double-layers (DLs) in an unmagnetized EDDPM with inertial dust grains and inertialess ions, and reported that both compressive and rarefactive DA-DLs are allowed by the plasma model. Hossen et al. [16] examined DAWs in a multi-component EDDPM with inertialess non-thermal ions and inertial massive OP dust grains, and found that the configuration of DA-DLs and DA-SWs is rigourously modified by the existence of positively charged dust grains.
The super-thermal κ -distribution [30,31,32,33,34,35,36,37] can describe the deviation, according to the values of the super-thermal parameter κ which manifests the presence of external force fields or wave-particle interactions, of plasma species from the thermally equilibrium state or Maxwellian distribution. The super-thermal κ -distribution exchanges with the Maxwellian distribution when κ tends to infinity, i.e., κ , and κ -distribution is normalizable for any values of κ rather than κ > 3 / 2 [32,33,34,35,36,37]. Uddin et al. [32] numerically analyzed the propagation of nonlinear electrostatic positron-acoustic waves in a super-thermal plasma, and reported that the amplitude of the electrostatic positive potential decreases with an increasing value of κ . Shahmansouri and Alinejad [33] examined DA-SWs in a DPM with super-thermal plasma species, and found that the depth of the potential well decreases with the increase in the value of κ . Kourakis and Sultana [34] examined the speed of the DIA solitons in the presence of super-thermal particles in a DPM.
The modulational instability (MI) of the carrier waves, energy re-distribution and localization are governed by the standard nonlinear Schrödinger equation (NLSE) [35,36,37,38,39,40,41,42]. Sultana and Kourakis [35] examined the electron-acoustic (EA) envelope solitons in a plasma medium with super-thermal electrons, and found that the stable domain of EA waves decreases with increasing κ . Ahmed et al. [36] reported IAWs in a four-component plasma medium, and highlighted that the critical wave number ( k c ) increases with a decrease in the value of κ . Saini and Kourakis [38] demonstrated the MI of the DAWs as a DPM with super-thermal ions, and obtained that excess super-thermality of the ions leads to a narrower envelope of solitons.
Recently, Hossen et al. [15] considered a three-component plasma model with inertial OP dust grains and non-extensive electrons to investigate the propagation of the DA-SWs. In this paper, we aim to extend on previous published work [15] by considering a real and novel four-component EDDPM with inertial OP dust grains and inertialess iso-thermal negative ions and super-thermal positive ions to examine the MI of DAWs and the formation of DA rogue waves (DARWs).
The layout of the paper is as follows: The governing equations are given in Section 2. The derivation of the standard NLSE is demonstrated in Section 3. The MI and rogue waves are presented in Section 4. The results and discussion are provided in Section 5. A brief conclusion is presented in Section 6.

2. Basic Equations

We consider a four-component EDDPM consisting of inertial negatively charged dust grains (mass m ; charge q = e Z ), positively charged dust grains (mass m + ; charge q + = e Z + ), inertialess negatively charged ions (mass m 1 ; charge q 1 = e Z 1 ), and positively charged ions (mass m 2 ; charge q 2 = e Z 2 ); where Z , Z + , Z 1 , and Z 2 are the charge state of the negative dust, positive dust, negative ion, and positive ion, respectively, and e is the magnitude of a single electron charge. At equilibrium, the quasi-neutrality condition can be expressed as Z n 0 + Z 1 n 10 = Z + n + 0 + Z 2 n 20 ; where n 0 , n 10 , n + 0 , and n 20 are, respectively, the equilibrium number densities of negative dust grains, negative ions, positive dust grains, and positive ions. Now, after normalization, the set of basic equations can be written as
n t + x ( n u ) = 0 ,
u t + u u x = λ 1 ϕ x ,
n + t + x ( n + u + ) = 0 ,
u + t + u + u + x = ϕ x ,
2 ϕ x 2 = λ 2 n 1 λ 3 n 2 + ( 1 λ 2 + λ 3 ) n n + ,
where n and n + are the negative and positive dust grains number density normalized by their equilibrium value n 0 and n + 0 , respectively; u and u + are the negative and positive dust fluid speed normalized by DA wave speed C + = ( Z + k B σ 1 / m + ) 1 / 2 (with σ 1 being the temperature negative ion and k B being the Boltzmann constant); ϕ is the electrostatic wave potential normalized by k B σ 1 / e ; the time and space variables are normalized by ω p + 1 = ( m + / 4 π e 2 Z + 2 n + 0 ) 1 / 2 and λ D + = ( k B σ 1 / 4 π e 2 Z + n + 0 ) 1 / 2 , respectively; λ 1 = Z m + / Z + m , λ 2 = Z 1 n 10 / Z + n + 0 , and λ 3 = Z 2 n 20 / Z + n + 0 . We have considered m > m + , Z > Z + , and n 0 > n + 0 for our numerical analysis. The expression for heavy negative ion number density obeying iso-thermal Maxwelliann distribution is given by
n 1 = exp ( ϕ ) .
The expression for light positive ion number density obeying super-thermal κ -distribution is given by [36]
n 2 = 1 + λ 4 ϕ κ 3 / 2 κ + 1 / 2 ,
where λ 4 = σ 1 / σ 2 (with σ 2 being the temperature positive ion) and σ 2 > σ 1 . The super-thermality of the light positive ion can be represented by the parameter κ . Now, by substituting Equations (6) and (7) into Equation (5), and expanding up to the third order of ϕ , we get
2 ϕ x 2 + n + + λ 3 = λ 2 + ( 1 λ 2 + λ 3 ) n + T 1 ϕ + T 2 ϕ 2 + T 3 ϕ 3 + · · · ,
where
T 1 = λ 2 ( 2 κ 3 ) + λ 3 λ 4 ( 2 κ 1 ) ( 2 κ 3 ) , T 2 = λ 2 ( 2 κ 3 ) 2 λ 3 λ 4 2 ( 2 κ 1 ) ( 2 κ + 1 ) 2 ( 2 κ 3 ) 2 , T 3 = λ 2 ( 2 κ 3 ) 3 + λ 3 λ 4 3 ( 2 κ 1 ) ( 2 κ + 1 ) ( 2 κ + 3 ) 6 ( 2 κ 3 ) 3 .
We note that the terms containing T 1 , T 2 , and T 3 are due to the contribution of positive and negative ions.

3. Derivation of the NLSE

The reductive perturbation method is applicable to derive the standard NLSE as well as to study the MI of the DAWs in a four-component EDDPM. The stretched co-ordinates, to develop a standard NLSE, can be written as [39,40,41,42]
ξ = ϵ ( x v g t ) ,
τ = ϵ 2 t ,
where v g is the group velocity and ϵ is a small parameter which measures the nonlinearity of the plasma medium. Then, the dependent variables can be written as [39,40,41,42]
n = 1 + m = 1 ϵ m l = n l ( m ) ( ξ , τ ) exp [ i l ( k x ω t ) ] ,
u = m = 1 ϵ m l = u l ( m ) ( ξ , τ ) exp [ i l ( k x ω t ) ] ,
n + = 1 + m = 1 ϵ m l = n + l ( m ) ( ξ , τ ) exp [ i l ( k x ω t ) ] ,
u + = m = 1 ϵ m l = u + l ( m ) ( ξ , τ ) exp [ i l ( k x ω t ) ] ,
ϕ = m = 1 ϵ m l = ϕ l ( m ) ( ξ , τ ) exp [ i l ( k x ω t ) ] .
For the above consideration, the derivative operators can be represented as [39,40,41,42]
t t ϵ v g ξ + ϵ 2 τ ,
x x + ϵ ξ .
Now, by substituting Equations (9)–(17) into Equations (1)–(4), and (8), and collecting the terms containing ϵ , the first order ( m = 1 with l = 1 ) reduced equations can be written as
n 1 ( 1 ) = λ 1 k 2 ω 2 ϕ 1 ( 1 ) , u 1 ( 1 ) = λ 1 k ω ϕ 1 ( 1 ) , n + 1 ( 1 ) = k 2 ω 2 ϕ 1 ( 1 ) , u + 1 ( 1 ) = k ω ϕ 1 ( 1 ) ,
these relations provide the dispersion relation for DAWs
ω 2 = k 2 + λ 1 k 2 λ 1 λ 2 k 2 + λ 1 λ 3 k 2 k 2 + T 1 .
The second-order ( m = 2 with l = 1 ) equations are given by
n 1 ( 2 ) = λ 1 k 2 ω 2 ϕ 1 ( 2 ) 2 i λ 1 k ( v g k ω ) ω 3 ϕ 1 ( 1 ) ξ ,
u 1 ( 2 ) = λ 1 k ω ϕ 1 ( 2 ) i λ 1 ( v g k ω ) ω 2 ϕ 1 ( 1 ) ξ ,
n + 1 ( 2 ) = k 2 ω 2 ϕ 1 ( 2 ) + 2 i k ( v g k ω ) ω 3 ϕ 1 ( 1 ) ξ ,
u + 1 ( 2 ) = k ω ϕ 1 ( 2 ) + i ( v g k ω ) ω 2 ϕ 1 ( 1 ) ξ ,
with the compatibility condition
v g = ω + ω λ 1 ω λ 1 λ 2 + ω λ 1 λ 3 ω 3 k + k λ 1 k λ 1 λ 2 + k λ 1 λ 3 .
The coefficients of ϵ for m = 2 with l = 2 provide the second-order harmonic amplitudes which are found to be proportional to | ϕ 1 ( 1 ) | 2
n 2 ( 2 ) = T 4 | ϕ 1 ( 1 ) | 2 , u 2 ( 2 ) = T 5 | ϕ 1 ( 1 ) | 2 , n + 2 ( 2 ) = T 6 | ϕ 1 ( 1 ) | 2 , u + 2 ( 2 ) = T 7 | ϕ 1 ( 1 ) | 2 , ϕ 2 ( 2 ) = T 8 | ϕ 1 ( 1 ) | 2 ,
where
T 4 = λ 1 k 2 ( 3 λ 1 k 2 2 ω 2 T 8 ) 2 ω 4 , T 5 = λ 1 k ( λ 1 k 2 2 ω 2 T 8 ) 2 ω 3 , T 6 = k 2 ( 2 ω 2 T 8 + 3 k 2 ) 2 ω 4 , T 7 = k ( 2 ω 2 T 8 + k 2 ) 2 ω 3 , T 8 = 2 T 2 ω 4 3 k 4 ( 1 + λ 2 λ 1 2 λ 1 2 λ 3 λ 1 2 ) 2 ω 2 k 2 ( 1 + λ 1 λ 1 λ 2 + λ 1 λ 3 ) 2 ω 4 ( 4 k 2 + T 1 ) .
Now, we consider the expression for ( m = 3 with l = 0 ) and ( m = 2 with l = 0 ), which leads the zeroth harmonic modes. Thus, we obtain
n 0 ( 2 ) = T 9 | ϕ 1 ( 1 ) | 2 , u 0 ( 2 ) = T 10 | ϕ 1 ( 1 ) | 2 , n + 0 ( 2 ) = T 11 | ϕ 1 ( 1 ) | 2 , u + 0 ( 2 ) = T 12 | ϕ 1 ( 1 ) | 2 , ϕ 0 ( 2 ) = T 13 | ϕ 1 ( 1 ) | 2 ,
where
T 9 = λ 1 2 k 2 ( ω + 2 k v g ) T 13 λ 1 ω 2 v g 2 ω 3 , T 10 = λ 1 ( λ 1 k 2 ω 2 T 13 ) v g ω 2 , T 11 = ω ( k 2 + ω 2 T 13 ) + 2 v g k 3 v g 2 ω 3 , T 12 = k 2 + ω 2 T 13 v g ω 2 , T 13 = 2 v g ( v g ω 3 T 2 k 3 ) ω k 2 + ( 1 λ 2 + λ 3 ) ( ω k 2 λ 1 2 + 2 v g k 3 λ 1 2 ) ω 3 + λ 1 ω 3 λ 1 λ 2 ω 3 + λ 1 λ 3 ω 3 T 1 ω 3 v g 2 .
Finally, the third harmonic modes ( m = 3 ) and ( l = 1 ) of Equations (1)–(4), and (8), and with the help of Equations (18)–(26) give a set of equations which can be reduced to the following NLSE:
i Φ τ + P 2 Φ ξ 2 + Q Φ 2 Φ = 0 ,
where Φ = ϕ 1 ( 1 ) for simplicity. In Equation (27), P is the dispersion coefficient which can be written as
P = 3 v g ( v g k ω ) 2 ω k ,
and also Q is the nonlinear coefficient which can be written as
Q = 3 ω 3 T 3 + 2 ω 3 T 2 ( T 8 + T 13 ) ω k 2 ( T 6 + T 11 ) + 2 k 3 ( T 7 + T 12 ) F 1 2 k 2 + 2 λ 1 k 2 2 λ 1 λ 2 k 2 + 2 λ 1 λ 3 k 2 ,
where F 1 = ( ω λ 1 k 2 ω λ 1 λ 2 k 2 + ω λ 1 λ 3 k 2 ) ( T 4 + T 9 ) + ( 2 λ 1 k 3 2 λ 1 λ 2 k 3 + 2 λ 1 λ 3 k 3 ) ( T 5 + T 10 ) . It may be noted here that both P and Q are functions of various plasma parameters such as k, λ 1 , λ 2 , λ 3 , λ 4 , and κ . So, all the plasma parameters are used to maintain the nonlinearity and the dispersion properties of the EDDPM.

4. Modulational Instability and Rogue Waves

The nonlinear property of DPM as well as the stability conditions of the DAWs in an EDDPM can be determined according to the sign of dispersive (P) and nonlinear (Q) coefficients of the standard NLSE (27) [39,40,41,42]. The sign of the dispersive coefficient is always negative for any kind of wave number k in the “P vs. k curve”, while the sign of the nonlinear coefficient is positive for small values of k and negative for large values of k in “Q vs. k curve” (figures are not included). Both P and Q have the same sign (i.e., positive or negative), then they determine a modulationally unstable parametric regime (i.e., P / Q > 0 ) of DAWs in the “ P / Q vs. k curve”, and when both P and Q have the opposite sign (i.e., positive and negative), then they determine a modulationally stable parametric regime (i.e., P / Q < 0 ) of DAWs in the presence of external perturbation [39,40,41,42]. The stable and unstable parametric regimes are differentiated by a vertical line at which P / Q (i.e., Q = 0 and because P is always negative), and the wave number for which P / Q is known as the critical waves number ( k k c ) [39,40,41,42]. We have numerically analyzed the modulationally stable and unstable parametric regimes of DAWs in Figure 1.
The modulationally unstable parametric regime of the DAWs allows to generate highly energetic and mysterious DARWs associated with DAWs, and the governing equation of the puzzling DARWs can be written as [43,44]
Φ ( ξ , τ ) = 2 P Q 4 ( 1 + 4 i P τ ) 1 + 16 P 2 τ 2 + 4 ξ 2 1 exp ( 2 i P τ ) .
We have numerically analyzed Equation (28) in Figure 2 to understand how the nonlinear properties of a four-component EDDPM and the configuration of the DARWs associated DAWs have been changed by the variation of different plasma parameters.

5. Results and Discussion

The variation of P / Q with k for different values of κ as well as the stable and unstable parametric regimes of the DAWs can be observed in the left panel of Figure 1, and it is obvious from this figure that (a) DAWs become modulationally stable for small values of k, while unstable for large values of k; (b) the k c decreases with the increase in the value of κ . The right panel of Figure 1 indicates the effects of the number density and charge state of negative ions and positive dust grains (via λ 2 = Z 1 n 10 / Z + n + 0 ) to recognize the stable and unstable parametric regimes of the DAWs, and it can be seen from this figure that (a) the modulationally stable parametric regime increases with λ 2 ; (b) the modulationally stable parametric regime increases with an increase in the value of negative ion number density ( n 10 ) while it decreases with an increase in the value of the positive dust grains number density ( n + 0 ) for a constant value of Z 1 and Z + ; (c) the modulationally stable parametric regime of the DAWs increases (decreases) with Z 1 ( Z + ) for a fixed negative ion and positive dust number density.
We have numerically analyzed first-order rogue waves [using Equation (28)] in Figure 2. The left panel of Figure 2 highlights the effects of the mass and charge state of the positive and negative dust grains (via λ 1 = Z m + / Z + m ) in recognizing the shape of the DARWs in an EDDPM, and it can be manifested from this figure that (a) the increase in the value of λ 1 does not only cause to change the amplitude of the DARWs but also causes to change the width of the DARWs; (b) the amplitude and width of the DARWs decrease with the increase in the value of λ 1 ; (c) actually, the nonlinearity of the plasma medium as well as the amplitude of the DARWs increases with increasing m , while the nonlinearity as well as the amplitude of the DARWs decreases with increasing m + for a fixed value of Z + and Z . The right panel of Figure 2 reflects how the number density and charge state of the positive ion and dust grains (via λ 3 = Z 2 n 20 / Z + n + 0 ) can contribute to generate highly energetic rogue waves in a four-component EDDPM. The amplitude and width of the electrostatic DARWs associated with DAWs decreases with an increase in the value of positive ion charge state and number density, while increases with an increase in the value of positive dust grain charge state and number density. The physics of this result is that the nonlinearity of a four-component EDDPM increases (decreases) with positive dust grain (negative ion) number density as well as with positive dust grain (negative ion) charge state.

6. Conclusions

In this present article, we have considered a four-component EDDPM with inertial opposite polarity dust grains and inertialess positive and negative ions. A nonlinear Schrödinger equation (NLSE) is derived by employing the reductive perturbation method. We have theoretically predicted the modulationally stable and unstable parametric regimes of DAWs, and associated DARWs in the modulationally unstable parametric regime of DAWs in an EDDPM. The results that have been found from our investigation can be summarized as follows:
  • Both modulationally stable and unstable parametric regimes of DAWs are allowed by the plasma model.
  • The critical wave number ( k c ) decreases with the increase in the value of κ .
  • The nonlinearity of the plasma medium as well as the amplitude of the DARWs increases (decreases) with increasing negative (positive) dust mass.
It may be noted here that the gravitational effect is very important to consider in the governing equations but is beyond the scope of our present work. However, electrons’ depletion process can be observed in astrophysical OPDP mediums, viz., Earth’s mesosphere (where the D-region plasma could suffer from electron density depletion) [27], cometary tails [4], Jupiter’s magnetosphere [5,6], F-ring of Saturn [28], and also laboratory EDDPM [12,24,25,26]. To conclude, the results obtained from the present investigation will be useful in the future to understand the MI of DAWs and associated DARWs in Earth’s mesosphere (where the D-region plasma could suffer from electron density depletion) [27], cometary tails [4], Jupiter’s magnetosphere [5,6], F-ring of Saturn [28], and laboratory EDDPM [12,24,25,26].

Author Contributions

All authors contributed equally to complete this work. All authors have read and agreed to the published version of the manuscript.

Funding

The research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable—no new data generated.

Acknowledgments

The authors are grateful to anonymous reviewers for their constructive suggestions which have significantly improved the quality of our manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Havnes, O.; Trim, J.; Blix, T.; Mortensen, W.; Nsheim, L.I.; Thrane, E.; Tnnesen, V. First detection of charged dust particles in the Earth’s mesosphere. J. Geophys. Res. 1996, 101, 10839. [Google Scholar] [CrossRef]
  2. Ellis, T.A.; Neff, J.S. Numerical simulation of the emission and motion of neutral and charged dust from P/Halley. Icarus 1991, 91, 280. [Google Scholar] [CrossRef]
  3. Horányi, M. Charged dust dynamics in the solar system. Annu. Rev. Astron. Astrophys. 1996, 34, 383. [Google Scholar] [CrossRef]
  4. Mendis, D.A.; Rosenberg, M. Cosmic dusty plasma. Annu. Rev. Astron. Astrophys. 1994, 32, 419. [Google Scholar] [CrossRef]
  5. Shukla, P.K.; Mamun, A.A. Introduction to Dusty Plasma Physics; IOP: London, UK, 2002. [Google Scholar]
  6. Horányi, M.; Morfill, G.E.; Grün, E. Mechanism for the acceleration and ejection of dust grains from Jupiter’s magnetosphere. Nature 1993, 363, 144. [Google Scholar] [CrossRef]
  7. Zhao, H.; Castle, G.S.P.; Lnculet, I.I.; Bailey, A.G. Bipolar charging of poly-disperse polymer powders in fluidized beds. IEEE Trans. Ind. Appl. 2003, 39, 612. [Google Scholar] [CrossRef]
  8. Trigwell, S.; Grable, N.; Yurreri, C.U.; Sharma, R.; Mazumder, M.K. Effects of surface properties on the tribocharging characteristics of polymer powder as applied to industrial processes. IEEE Trans. Ind. Appl. 2003, 39, 79. [Google Scholar] [CrossRef]
  9. Zhao, H.; Castle, G.S.P.; Lncullet, I.I. The measurement of bipolar charge in polydisperse powders using a vertical array of Faraday pail sensors. J. Electr. 2002, 55, 261. [Google Scholar] [CrossRef]
  10. Ali, F.S.; Ali, M.A.; Ali, R.A.; Lncullet, I.I. Minority charge separation in falling particles with bipolar charge. J. Electr. 1998, 45, 139. [Google Scholar]
  11. Rao, N.N.; Shukla, P.K.; Yu, M.Y. Dust-acoustic waves in dusty plasmas. Planet. Space Sci. 1990, 38, 543. [Google Scholar] [CrossRef]
  12. Barkan, A.; Merlino, R.L.; D’Angelo, N. Laboratory observation of the dust-acoustic wave mode. Phys. Plasmas 1995, 2, 3563. [Google Scholar] [CrossRef] [Green Version]
  13. Shukla, P.K.; Yu, M.; Bharuthram, Y.R. Linear and nonlinear dust drift waves. J. Geophys. Res. 1991, 96, 21343. [Google Scholar] [CrossRef]
  14. Shahmansouri, M.; Alinejad, H. Dust acoustic solitary waves in a magnetized electron depleted superthermal dusty plasma. Phys. Plasmas 2013, 20, 033704. [Google Scholar] [CrossRef]
  15. Hossen, M.M.; Alam, M.S.; Sultana, S.; Mamun, A.A. Low frequency nonlinear waves in electron depleted magnetized nonthermal plasmas. Eur. Phys. J. D 2016, 70, 252. [Google Scholar] [CrossRef]
  16. Hossen, M.M.; Alam, M.S.; Sultana, S.; Mamun, A.A. Oblique propagation of low frequency nonlinear waves in an electron depleted magnetized plasma with positive and negative dust. Phys. Plasmas 2016, 23, 023703. [Google Scholar] [CrossRef]
  17. Hossen, M.M.; Nahar, L.; Alam, M.S.; Sultana, S.; Mamun, A.A. Electrostatic shock waves in a nonthermal dusty plasma with oppositely charged dust. High Energy Density Phys. 2017, 24, 9. [Google Scholar] [CrossRef]
  18. Melandso, F. Lattice waves in dust plasma crystals. Phys. Plasmas 1996, 3, 3890. [Google Scholar] [CrossRef]
  19. Ferdousi, M.; Miah, M.R.; Sultana, S.; Mamun, A.A. Dust-acoustic shock waves in an electron depleted nonextensive dusty plasma. Astrophys. Space Sci. 2015, 360, 43. [Google Scholar] [CrossRef]
  20. Shukla, P.K.; Silin, V.P. Dust ion-acoustic wave. Phys. Scr. 1992, 45, 508. [Google Scholar] [CrossRef]
  21. Mamun, A.A.; Cairns, R.A.; Shukla, P.K. Solitary potentials in dusty plasmas. Phys. Plasma 1996, 3, 702. [Google Scholar] [CrossRef]
  22. Dialynas, K.; Krimigis, S.M.; Mitchemm, D.G.; Hamilton, D.C.; Krupp, N.; Brandt, P.C. Energetic ion spectral characteristics in the Saturnian magnetosphere using Cassini/MIMI measurements. J. Geophys. Res. 2009, 114, A01212. [Google Scholar] [CrossRef] [Green Version]
  23. Mayout, S.; Tribeche, M. Arbitrary amplitude dust-acoustic double-layers in an electron-depleted dusty plasma with two high-energy tail ion distributions. J. Plasma Phys. 2012, 78, 657. [Google Scholar] [CrossRef]
  24. Tadsen, B.; Greiner, F.; Groth, S.; Piel, A. Self-excited dust-acoustic waves in an electron-depleted nanodusty plasma. Phys. Plasmas 2015, 22, 113701. [Google Scholar] [CrossRef]
  25. Goertz, I.; Greiner, F.; Piel, A. Effects of charge depletion in dusty plasmas. Phys. Plasmas 2011, 18, 013703. [Google Scholar] [CrossRef] [Green Version]
  26. Kim, S.H.; Merlinoa, R.L. Charging of dust grains in a plasma with negative ions. Phys. Plasmas 2006, 13, 052118. [Google Scholar] [CrossRef] [Green Version]
  27. Friedrich, M.; Rapp, M.; Plane, J.M.C.; Torkar, K.M. Bite-outs and other depletions of mesospheric electrons. J. Atmos. Sol. Terr. Phys. 2011, 73, 2201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Goertz, C.K. Dusty plasma in the solar system. Rev. Geophys. 1989, 27, 271. [Google Scholar] [CrossRef]
  29. Sahu, B.; Tribeche, M. Small amplitude double-layers in an electron depleted dusty plasma with ions featuring the Tsallis distribution. Astrophys. Space Sci. 2012, 341, 573. [Google Scholar] [CrossRef]
  30. Ferdousi, M.; Sultana, S.; Hossen, M.M.; Miah, M.R.; Mamun, A.A. Dust-acoustic shock excitations in κ-nonthermal electron depleted dusty plasmas. Eur. Phys. J. D 2017, 71, 102. [Google Scholar] [CrossRef]
  31. Vasyliunas, V.M. A survey of low-energy electrons in the evening sector of the magnetosphere with OGO 1 and OGO 3. J. Geophys. Res. 1968, 73, 2839. [Google Scholar] [CrossRef]
  32. Uddin, M.J.; Alam, M.S.; Mamun, A.A. Positron-acoustic shock waves associated with cold viscous positron fluid in superthermal electron-positron-ion plasmas. Phys. Plasmas 2015, 22, 062111. [Google Scholar] [CrossRef]
  33. Shahmansouri, M.; Alinejad, H. Arbitrary amplitude dust ion acoustic solitary waves in a magnetized suprathermal dusty plasma. Phys. Plasmas 2012, 19, 123701. [Google Scholar] [CrossRef]
  34. Kourakis, I.; Sultana, S. Electrostatic waves in superthermal dusty plasmas: Review of recent advancement. AIP Conf. Proc. 2011, 1397, 86. [Google Scholar]
  35. Sultana, S.; Kourakis, I. Electrostatic solitary waves in the presence of excess superthermal electrons: Modulational instability and envelope soliton modes. Plasma Phys. Control. Fusion 2011, 53, 045003. [Google Scholar] [CrossRef] [Green Version]
  36. Ahmed, N.; Mannan, A.; Chowdhury, N.A.; Mamun, A.A. Electrostatic rogue waves in double pair plasmas. Phys. Plasmas 2010, 17, 013701. [Google Scholar] [CrossRef] [Green Version]
  37. Gill, T.S.; Bains, A.S.; Bedi, C. Modulational instability of dust acoustic solitons in multicomponent plasma with kappa-distributed electrons and ions. Chaos 2018, 28, 123107. [Google Scholar] [CrossRef]
  38. Saini, N.S.; Kourakis, I. Dust-acoustic wave modulation in the presence of superthermal ions. Phys. Plasmas 2018, 15, 123701. [Google Scholar] [CrossRef]
  39. Shikha, R.K.; Chowdhury, N.A.; Mannan, A.; Mamun, A.A. Electrostatic dust-acoustic envelope solitons in an electron-depleted plasma. Contrib. Plasma Phys. 2021, 61, e202000117. [Google Scholar] [CrossRef]
  40. Noman, A.A.; Chowdhury, N.A.; Mannan, A.; Mamun, A.A. Dust-acoustic envelope solitons in super-thermal plasmas. Contrib. Plasma Phys. 2019, 59, e201900023. [Google Scholar] [CrossRef] [Green Version]
  41. Zaman, D.M.S.; Mannan, A.; Chowdhury, N.A.; Mamun, A.A. Dust-acoustic rogue waves in opposite polarity dusty plasma featuring nonextensive statistics. High Temp. 2020, 58, 789. [Google Scholar] [CrossRef]
  42. Banik, S.; Shikha, R.K.; Noman, A.A.; Chowdhury, N.A.; Mannan, A.; Roy, T.S.; Mamun, A.A. First and second-order dust-ion-acoustic rogue waves in non-thermal plasma. Eur. Phys. J. D 2021, 75, 43. [Google Scholar] [CrossRef]
  43. Akhmediev, N.; Ankiewicz, A.; Soto-Crespo, J.M. Rogue waves and rational solutions of the nonlinear Schrödinger equation. Phys. Rev. E 2009, 80, 026601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ankiewicz, A.; Devine, N.; Akhmediev, N. Are rogue waves robust against perturbations? Phys. Lett. A 2009, 373, 3997. [Google Scholar] [CrossRef]
Figure 1. The variation of P / Q with k for different values of κ when λ 2 = 2.0 (left panel), and the variation of P / Q with k for different values of λ 2 when κ = 1.8 (right panel). Other plasma parameters are λ 1 = 0.8 , λ 3 = 1.5 , and λ 4 = 0.4 .
Figure 1. The variation of P / Q with k for different values of κ when λ 2 = 2.0 (left panel), and the variation of P / Q with k for different values of λ 2 when κ = 1.8 (right panel). Other plasma parameters are λ 1 = 0.8 , λ 3 = 1.5 , and λ 4 = 0.4 .
Plasma 04 00015 g001
Figure 2. The variation of | Φ | with ξ for different values of λ 1 when λ 3 = 1.5 (left panel), and the variation of | Φ | with ξ for different values of λ 3 when λ 1 = 0.8 (right panel). Other plasma parameters are λ 2 = 2.0 , λ 4 = 0.4 , κ = 1.8 , and k = 1.1 .
Figure 2. The variation of | Φ | with ξ for different values of λ 1 when λ 3 = 1.5 (left panel), and the variation of | Φ | with ξ for different values of λ 3 when λ 1 = 0.8 (right panel). Other plasma parameters are λ 2 = 2.0 , λ 4 = 0.4 , κ = 1.8 , and k = 1.1 .
Plasma 04 00015 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sikta, J.N.; Chowdhury, N.A.; Mannan, A.; Sultana, S.; Mamun, A.A. Electrostatic Dust-Acoustic Rogue Waves in an Electron Depleted Dusty Plasma. Plasma 2021, 4, 230-238. https://0-doi-org.brum.beds.ac.uk/10.3390/plasma4020015

AMA Style

Sikta JN, Chowdhury NA, Mannan A, Sultana S, Mamun AA. Electrostatic Dust-Acoustic Rogue Waves in an Electron Depleted Dusty Plasma. Plasma. 2021; 4(2):230-238. https://0-doi-org.brum.beds.ac.uk/10.3390/plasma4020015

Chicago/Turabian Style

Sikta, Jebun Naher, Nure Alam Chowdhury, Abdul Mannan, Sharmin Sultana, and A. A. Mamun. 2021. "Electrostatic Dust-Acoustic Rogue Waves in an Electron Depleted Dusty Plasma" Plasma 4, no. 2: 230-238. https://0-doi-org.brum.beds.ac.uk/10.3390/plasma4020015

Article Metrics

Back to TopTop