The dynamical effects of a cylindrically symmetric moving shock wave in a self-gravitating and rotating ideal gas medium influenced by a magnetic field and radiation heat flux are investigated. The dynamics of the shock wave are governed by the one-dimensional motion of the gas, and the energy equation contains the effect of thermal radiation in the setting of an optically thick limit. A mathematical model is established using the Lie symmetry method, and all possible cases of similarity solutions are obtained by choosing different subalgebras from the optimal system. The numerical computations are performed using a fourth-order Runge–Kutta method. For a power law shock path, the contour plots of flow variables such as density and velocity are presented for convenient visualization. The variations of the shock strength and the flow variables in the physical flow field region behind the shock front with changes in the values of the gravitational parameter, rotational parameter, adiabatic index, ambient magnetic field strength, similarity exponent, and radiative heat transfer parameter are investigated. It is found that increases in the ambient magnetic field strength and the radiative heat transfer parameter lead to decay of the shock wave. For a power law shock path, it is found that increases in the gravitational parameter and adiabatic index cause decay of the shock wave, whereas for an exponential law shock path, the shock strength is increased by increases in the values of these parameters.

Experimental investigations and observations in astrophysics reveal that the outer atmospheres of planets, stars, and galaxies undergo rotational motion due to the spin of these objects. This rotational motion induces macroscopic motion at supersonic velocities within the atmospheres, leading to the generation of shock waves. Consequently, the rotation of planets and stars exerts a significant influence on the processes occurring in their outer layers, and thus explosions within rotating gaseous atmospheres are of considerable astrophysical interest.1–4 Alzahrani and Khan5 discussed entropy generation, Joule heating, and homogenous and heterogeneous chemical reactions in Ree–Eyring fluid flow between double rotating disks. Chaturani6 examined the propagation of cylindrical shock waves through a gas experiencing solid body rotation using the similarity method of Sedov.7 Levin and Skopina8 obtained a condition for the existence of Chapman–Jouguet waves and studied shock and detonation waves in rotating gas flows. Nath9 discussed the similarity solution for a strong shock moving with an exponential-law shock path in rotating dusty gas flow.

During the last few decades, there has been considerable research interest in electrically conducting fluid flows, such as those involved in the motion of electrically conducting fluids, supersonic motion of interstellar gas, supernova explosions, the Earth’s magnetic field, and flow through magnetized nozzles. Problems related to the existence and properties of magnetic fields are extremely important in the study of geological and astrophysical events.10–18 Raza et al.19 studied the impact of magnetic fields on entropy generation in magnetohydrodynamic (MHD) nanofluids using a neural network approach. Mishra et al.20 studied the effects of magnetic field in dusty nanofluids containing nanoparticles such as copper and silver as the conducting medium. In studies of high-temperature and low-pressure phenomena, the gaseous medium is often assumed to follow the ideal gas law, which reduces the complexity of the physical model. We shall adopt this assumption in the present study. There have been many investigations of implosion shock propagation in a perfect gas with or without a magnetic field.21–25 The acceleration of plane shock waves described by solid body motion in a perfect gas in the presence of a magnetic field were discussed by Golubyatnikov and Kovalevskaya.26 Nath27 studied the behavior of a shock wave in rotating perfect gas flow in the presence of an azimuthal magnetic field and obtained a similarity solution.

Thermal radiation is a mode of transfer of heat energy as electromagnetic waves. The effect of radiation becomes significant for high-temperature phenomena in medium at low density, such those encountered in blast waves due to powerful explosions, large electrical discharges in rotating medium, the solar corona and solar wind, and supernova explosions.28 The interaction of radiation heat flux and gasdynamics in the absence or presence of magnetic fields was analyzed by Pai,29 and the behavior of shock waves in the presence of radiation was studied by Marshak.30 Raza et al.31 discussed the thermal properties of graphene oxide (GO) and molybdenum disulphide (MoS2) nanoparticles with the help of fractional derivatives. There have been a number of studies of thermal radiative phenomenon with the aim of improving the heat transfer behavior. Sulochana et al.32 investigated the impact of multiwalled carbon nanotubes (MWCNTs) and hybrid biodiesel on a compression ignition engine with regard to emission, combustion, and performance. Khan et al.33 analyzed the thermal mechanisms involved in mixed convective flow of a Reiner–Philippoff nanofluid, with an assessment of entropy generation. Thermal performance is affected by thermal radiation, Ohmic dissipation, and nonuniform heat sources and sinks. Nazir et al.34 examined the convective flow of a nanofluid in a square container under the influence of thermal boundary conditions. Nicastro35 studied the gasdynamic equations in the presence of radiation under the assumption that the gas was optically thick and obtained a similarity solution. When the mean free path of radiation in a gas is small, the gas is said to be optically thick, and conversely, when the mean free path is long, the gas is optically thin. For an optically thick gas, the expression for the radiation heat flux takes a very simple form. The mean free path of radiation represents the mean distance over which a photon travels before it is absorbed by the molecules of the gas. Thus, it is similar to the mean free path considered in the kinetic theory of gases, which represents the mean distance between collisions. The mean free path of radiation is defined as NR = 1/κρ, where κ is the Rosseland mean absorption coefficient and ρ is the density of the medium. Therefore, if a gas is optically thick, then the value of the absorption coefficient is high, and the effects of radiation in the medium are significant.29 Summers36 and Hutchens37 discussed blast waves generated owing to the release of huge amounts of energy in conducting and nonconducting media, respectively. Chu et al.38 discussed the effects of magnetic fields and radiation on steady MHD flow of Maxwell fluid generated by the stretching of two infinite disks. Vishwakarma et al.39 and Vishwakarma and Patel40 discussed shock wave propagation in a nonrotating and a rotating nonideal gas in the presence of a magnetic field and radiation heat flux. Vishwakarma and Nath41 and Nath42 studied the behavior of shock waves in the presence of radiation and conduction heat fluxes in rotating dusty gas. The propagation of MHD waves in the presence of a magnetic field was discussed by Rosenau and Frankenthal,43,44 with and without consideration of the thermal conductivity of the medium. Cioffi et al.45 studied supernova remnants in the interstellar medium produced by blast waves due to supernova explosion. Nath et al.46 investigated the effects of magnetic field, radiation, and gravitation on shock wave propagation, using the similarity method of Sedov.7 Nath and Vishwakarma47 discussed shock wave propagation in a nonideal gas in the presence of an azimuthal magnetic field, radiation, and heat conduction, without taking into account the effects of a rotating medium. Nath48 extended the problem studied by Nath and Vishwakarma47 by considering these rotational effects.

The effects of a self-gravitating gaseous medium are important for the formation of stars and galaxies and for sudden expansion of instantaneous line explosions. Gravitational effects on shock wave propagation have been the subjects of a number of studies.49–52 Larson53 numerically solved the hydrodynamic equations governing the evolution of a protostar during its collapse under self-gravitation. Rogers54 obtained analytical solutions for blast wave propagation in an atmosphere with varying density ahead of the wavefront. Sakurai55 extended the work of Kopal50 by considering variable shock strength. Shock wave propagation in a self-gravitating gas with or without magnetic field effects has been studied using a variety of techniques.56–58 Similarity solutions for the propagation of cylindrical shock waves in gravitating or nongravitating perfect gas with radiation heat flux were obtained by Singh59 and Singh and Vishwakarma,60 and Vishwakarma and Singh61 extended this work by considerating radiation as well as heat conduction fluxes. An exact solution was obtained by Nath et al.62 for a shock wave in a self-gravitating perfect gas with radiation flux and magnetic field, but without considering the effects of a rotating medium.

In most of the above studies, the method of dimensional analysis developed by Sedov7 was used to discuss the propagation of shock waves. An alternative approach is represented by the use of Lie symmetries. The Lie symmetry method was originally developed by Sophus Lie in the 19th century.63 It is a powerful method for solving nonlinear partial differential equations (PDEs) and for studying nonlinear mathematical models. This method is based on the invariance properties of differential equations under an infinitesimal group of transformations. Using invariance conditions on differential equations, the infinitesimal generators of the transformation group containing arbitrary constants are obtained. Different choices of arbitrary constants in the infinitesimal generators provide all possible solutions of the system of basic equations of motion.63,64 A brief discussion of the Lie theoretic method and its applications in different areas can be found in Refs. 63 and 65–68. Similarity solutions for the propagation of shock wave in gases were obtained by Nath and Devi69 and Singh70 using a Lie group transformation method. Using the Lie symmetry method, an optimal system was discussed by Ovsiannikov,71 who constructed a global matrix for the adjoint transformation. The concept of an optimal system is very useful for minimizing the search required to obtain all different solutions of a system of PDEs that leave the system invariant. Recently, Nath and Maurya72 and Nath and Kadam73 used the Lie symmetry method to obtain the optimal system for the propagation of shock waves in a self-gravitating nonideal or ideal gas in the absence of presence of monochromatic radiation in a rotating medium. Hu et al.74 proposed a promising new approach for finding the optimal system of Lie subalgebras for a one-dimensional general system of ordinary differential equations (ODEs). An optimal system of one-, two-, and three-dimensional Lie subalgebras for two-phase mass flow model was discussed by GhoshHajra et al.,75 and a similar approach for three-dimensional gasdynamical equations was presented by Mandal et al.76 Using a Lie symmetry method, Nadjafikhah and Zaeim77 studied the classification of subalgebras in a hyperbolic space for wave equations.

In the problem considered in the present paper, the Lie symmetry method is used to derive all possible similarity solutions of a mathematical model that describes shock wave propagation in a self-gravitating perfect gas under the influence of a magnetic field, the rotation of the medium, and a radiation heat flux. We determine the inequivalent classes of subalgebras (optimal classes), of which there are 14 in the problem under consideration. Then, from these 14 inequivalent optimal classes of subalgebras, we obtain the optimal system. For each optimal class, group-invariant solutions of the problem exist, but our aim is to obtain the similarity solution, and it turns out that of the 14 cases, only three allow the existence of similarity solutions.

The present study generalizes the work of Singh59 by taking into consideration the rotational effect of the medium and the presence of a magnetic field as well as a gravitational force. It also represents a generalization of the work of Singh and Vishwakarma,60 again by taking into account the effect of a magnetic field and the rotational effect of the medium. In Refs. 59 and 60, the similarity solution for the power-law shock path was obtained by using the dimensional analysis method develop by Sedov,7 but we obtain similarity solutions in three cases, two with a power-law shock path and one with an exponential-law shock path, using a Lie symmetry method.63,71 Our mathematical model may provide an improved basis for understanding a variety of astrophysical and aerodynamic phenomena, such as supernova explosions, the central regions of starburst galaxies, nuclear explosions, and exploding wire experiments on the pinch effect in rotating media.1,2,29

To obtain similarity solutions of the problem under consideration, we assume that the azimuthal and axial fluid velocities and the density and magnetic field in the ambient medium all vary with the spatial coordinate. We also assume that the fluid is inviscid and perfectly conducting. In addition, we take the radial fluid velocity in the ambient medium to be zero.

Under these assumptions, numerical solution of the problem indicates that when the value of the gravitational parameter is increased, the shock wave strength decreases in the case of a power-law shock path and increases in the case of an exponential-law shock path. It is also found that the impact of the rotational parameter on the shock strength is the reverse of that of the gravitational parameter. However, the impacts of the magnetic field strength and radiation parameter on shock strength are similar for both power-law and exponential-law shock paths. In addition, the shock strength varies with a change in the magnetic field from axial to azimuthal.

The flow variables are denoted as follows: density by ρ, pressure by p, magnetic field by h, radial, azimuthal, and axial components of fluid velocity Z by u, v, and w, respectively, mass of fluid within a cylinder of unit length by m, and radiation heat flux by J. We assume that the gas is a perfect gas, and therefore the equation of state and internal energy per unit mass are expressed as6,10,56
(1)
where I, γ, R*, and T denote the internal energy per unit mass, adiabatic index, gas constant, and absolute temperature, respectively. We also assume that the gas is inviscid and with infinite electrical conductivity. We assume that the gaseous motion is one-dimensional unsteady adiabatic motion of a self-gravitating perfect gas and that a diverging cylindrical shock wave propagates symmetrically about the line of symmetry in the rotating medium under the influence of the radiation heat flux and the magnetic field. The mathematical model for such a type of problem is governed by the conservation laws of mass, momentum, energy, charge, etc. In an Eulerian formulation, the governing equations are expressed as30,40,42,47,60–62,70
(2)
(3)
(4)
(5)
(6)
(7)
(8)
where d/dt = /∂t + u∂/∂r, r and t are space and time coordinates, G and μ denote the constant of gravitation and magnetic permeability, respectively, and χ = 0 and 1 indicate an axial and azimuthal magnetic field, respectively.
Assuming local thermodynamic equilibrium and using the radiative diffusion model for an optically thick gray gas, the radiative heat flux J may be obtained from the differential approximation of the radiation transport equation in the diffusion limit as28,47,48,59,60
(9)
where ϰ is the Stefan–Boltzmann constant, and κ is the Rosseland mean absorption coefficient, the value of which is high for an optically thick gray gas and therefore the effect of radiation on the medium is significant.29 
The absorption coefficient κ is assumed to vary with temperature and density:48,60
(10)
where κo is a dimensional constant, and τ1 and τ2 are nondimensional constants. In Eq. (10), the exponents τ1, τ2 and the constant κo are to be determined from the gas-property data within the appropriate temperature range; if a self-similar solution is sought, they must also satisfy the similarity requirements.
The relation between the azimuthal fluid velocity v and the angular velocity of the medium Ω is expressed as
(11)
The vorticity vector is defined as8,42
(12)
and its components are given by
(13)
We assume that r = R(t) is the locus of the outwardly moving shock front and Us = dR/dt is its velocity. The pre-assumptions for the flow variables immediately ahead of the shock front are taken as
(14)
where a subscript “a” indicates the value of a flow variable ahead of the shock front.
We consider that the shock is isothermal (according to the approximation adopted here, a proportionality between the heat flux and the temperature gradient leads to the creation of an isothermal shock, i.e., the jump in temperature is negligible28). Thus, for a magnetogasdynamic isothermal shock, the Rankine–Hugoniot jump conditions are expressed as27,29,46,61
(15)
where a subscript “b” indicates the value of a flow variable immediately behind the shock wave front.
Using Eqs. (1) and (15), we obtain the shock jump conditions across the shock front as72 
(16)
where
AM=[ρaUs2/μ(ha)2]1/2 is the Alfvén Mach number, SM=(ρaUs2/γpa)1/2 is the shock Mach number referred to the frozen speed of sound (γpa/ρa)1/2, and ρa/ρb = β is the density ratio across the shock front, the value of which lies within the range from 0 to 1 for the existence of a shock wave.
The value of β is obtained from the quadratic equation
(17)
To discuss the similarity solutions of the present problem, the Lie symmetry method is used. To apply this method, the one-parameter group of infinitesimal transformations corresponding to the flow variables involved in the problem is taken into account. These are obtained through a Taylor’s series expansion of the one-parameter group of global transformations about the identity transformation in terms of a small quantity ϵ. We assume that the basic governing Eqs. (2)(8) and the shock jump conditions (16) are invariant under the infinitesimal transformations. Thus, the one-parameter group of infinitesimal transformations is taken as consisting of 65–67,72,73
(18)
where ξi and ηj, which are functions of (r, t, ρ, u, v, w, p, h, m, J), are known as the infinitesimal generators of the transformations (18), where i = 1, 2, 3, 4, 5, 6, 7, 8; j = 1, 2. In Eq. (18), ϵ is a small quantity, and so we can neglect second- and higher-order terms in ϵ.

We now introduce new symbols for the variables and their first-order partial derivatives as follows: t = x1, r = x2, ρ = X1, u = X2, v = X3, w = X4, p = X5, h = X6, m = X7, J = X8, and Yji=Xi/xj; i = 1, …, 8 and j = 1, 2.

The fundamental Eqs. (2)(8) can be written as
(19)
This system of equations is constantly and conformally invariant with respect to the transformations (18) if
(20)
where k, l = 1, …, 8, and the σkl are real numbers. L is called the Lie derivative and it defined as
(21)
with
(22)
where q = 1, 2 and e, s = 1, 2, 3, …, 8.
From Eqs. (19)(22), we obtain
(23)
Substituting the governing Eqs. (2)(9) one by one into Eq. (23) and setting the coefficients of Yqi, YqiYje, and the terms free from derivatives of dependent variables equal to zero gives a overdetermined system of equations in the infinitesimal generators called the system of determining equations. From these determining equations, the infinitesimal generators are obtained.
On solving the set of determining equations thereby obtained, the infinitesimal generators are obtained in the form
(24)
where c, σ22, σ44, k1, and a are real constants, and g(r, t) is taken as an unknown function of r and t as an integration constant for ξ8.
The infinitesimal generators (24) contain five constants, and therefore they generate a five-dimensional vector space (Lie algebra), which we denote by Π5. This is generated by five distinct Lie vectors, given by
(25)
To obtain the optimal system for the solution of the fundamental Eqs. (2)(8), the method described by Wambura et al.68 and Nadjafikhah and Zaeim77 is used. We express the commutation relations of the Lie vectors Vα, where α = 1, …, 5 using Lie brackets as
(26)
Further, the Lie series is defined by
(27)
where ϵα (α = 1, …, 5) are small quantities.

With the help of Eqs. (26) and (27), we obtain the commutator table and adjoint representation of the vector field (25) as shown in Tables I and II, respectively.

TABLE I.

Commutator table.

[Vα,Vξ]V1V2V3V4V5
V1 V1 
V2 
V3 V4 
V4 V4 V4 
V5 V1 V4 
[Vα,Vξ]V1V2V3V4V5
V1 V1 
V2 
V3 V4 
V4 V4 V4 
V5 V1 V4 
TABLE II.

Adjoint table.

RV1V2V3V4V5
V1 V1 V2 V3 V4 V5ϵ1V1 
V2 V1 V2 V3 V4 V5 
V3 V1 V2 V3 V4eϵ3 V5 
V4 V1 V2 V3ϵ4V4 V4 V5ϵ4V4 
V5 V1eϵ1 V2 V3 V4eϵ5 V5 
RV1V2V3V4V5
V1 V1 V2 V3 V4 V5ϵ1V1 
V2 V1 V2 V3 V4 V5 
V3 V1 V2 V3 V4eϵ3 V5 
V4 V1 V2 V3ϵ4V4 V4 V5ϵ4V4 
V5 V1eϵ1 V2 V3 V4eϵ5 V5 
The general adjoint matrix D is derived in the  Appendix. With the help of this general adjoint matrix, the optimal system is obtained as68,72
or
(28)
where ω ≠ 0, y=y1y2y3y4y5Trc1c2c3c4c5Tr, and superscript “Tr” indicates the transpose of a matrix. By solving Eq. (28) for different choices of z1, z2, z3, z4 and z5, we obtain all the possible subalgebras of Π5, as shown in Table III. It is important to note that some subalgebras are equal under translational transformation: for example, Φ3 = Φ13; Φ5 = Φ9 = Φ14 = Φ15 = Φ20 = Φ21 = Φ25 = Φ27; Φ7 = Φ19; Φ10 = Φ22; Φ12 = Φ18 = Φ24 = Φ28; Φ16 = Φ30; Φ23 = Φ26 = Φ29 = Φ31. This can easily be proved by applying a translational transformation to η1 or ξ4.
TABLE III.

Construction of subalgebras of Lie algebra Π5.

S. no.CaseBasis element (y)Sub-algebra Φd
z1 ≠ 0, z2 = z3 = z4 = z5 = 0. Let ϵ1 = 0 and z1/ω = c1 (c1, 0, 0, 0, 0) V1 
z1 = z3 = z4 = 0 = z5, z2 = ωc2 ≠ 0 (0, c2, 0, 0, 0) V2 
z1 = z2 = z4 = 0 = z5, z3 = ωc3 ≠ 0 (0, 0, c3, 0, 0) V3 
z1 = z2 = z3 = z5 = 0, z4 ≠ 0. Let ϵ3 + ϵ5 = log(ωc4/z4), ϵ4 = 0 (0, 0, 0, c4, 0) V4 
z1 = z2 = z3 = z4 = 0, z5 = ωc5 ≠ 0 (0, 0, 0, 0, c5V5 
z3 = z4 = z5 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0. Let ϵ1 = 0, z1/ω = c1, z2/ω = c2 (c1, c2, 0, 0, 0) c1V1+c2V2 
z2 = z4 = z5 = 0, z1 ≠ 0, z3 = ωc3 ≠ 0. Let ϵ1 = 0, z1/ω = c1, z3/ω = c3 (c1, 0, c3, 0, 0) c1V1+c3V3 
z2 = z3 = z5 = 0, z1 = ωc1 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ3 + ϵ5 = log(ωc4/z4(c1, 0, 0, c4, 0) c1V1+c4V4 
z2 = z3 = z4 = 0, z5 = ωc5 ≠ 0, z1 ≠ 0. Let ϵ1 = 0, z1/ω = c1 (c1, 0, 0, 0, c5c1V1+c5V5 
10 z1 = z4 = z5 = 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0 (0, c2, c3, 0, 0) c2V2+c3V3 
11 z1 = z3 = z5 = 0, z2 = ωc2 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, c2, 0, c4, 0) c2V2+c4V4 
12 z1 = z3 = z4 = 0, z2 = ωc2 ≠ 0, z5 = ωc5 ≠ 0 (0, c2, 0, 0, c5c2V2+c5V5 
13 z1 = z2 = z5 = 0, z3 = ωc3 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, 0, c3, c4, 0) c3V3+c4V4 
14 z1 = z2 = z4 = 0, z3 = ωc3 ≠ 0, z5 = ωc3 ≠ 0 (0, 0, c3, 0, c5c3V3+c5V5 
15 z1 = 0, z2 = 0, z4 ≠ 0, z3 = 0, z5 = ωc5 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, 0, 0, c4, c5c4V4+c5V5 
16 z4 = z5 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0. Let ϵ1 = 0, z1/ω = c1 (c1, c2, c3, 0, 0) c1V1+c2V2+c3V3 
17 z3 = z5 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, c2, 0, c4, 0) c1V1+c2V2+c4V4 
18 z3 = z4 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0, z5 = ωc5 ≠ 0. Let ϵ1 = 0, z1/ω = c1 (c1, c2, 0, 0, c5c1V1+c2V2+c5V5 
19 z2 = z4 = 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0, z1 ≠ 0. Let ϵ1 = 0 = ϵ4 = ϵ5, ϵ3 = log(ωc4/z4(c1, 0, c3, c4, 0) c1V1+c3V3+c4V4 
20 z2 = z4 = 0, z1 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0. Let ϵ1 = 0 (c1, 0, c3, 0, c5c1V1+c3V3+c5V5 
21 z2 = z3 = 0, z1 = ωc1 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, 0, 0, c4, c5c1V1+c4V4+c5V5 
22 z1 = z5 = 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, c2, c3, c4, 0) c2V2+c3V3+c4V4 
23 z1 = z4 = 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0. (0, c2, c3, 0, c5c2V2+c3V3+c5V5 
24 z1 = z3 = 0, z2 = ωc2 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, c2, 0, c4, c5c2V2+c4V4+c5V5 
25 z5 = ωc5 ≠ 0, z1 = 0 = z2, z3 = ωc3 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, 0, c3, c4, c5c3V3+c4V4+c5V5 
26 z1 = 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, c2, c3, c4, c5c2V2+c3V3+c4V4+c5V5 
27 z2 = 0, z1 = ωc1 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, 0, c3, c4, c5c1V1+c3V3+c4V4+c5V5 
28 z1 ≠ 0, z2 = ωc2 ≠ 0, z4 ≠ 0, z5 = ωc5 ≠ 0, z3 = 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, c2, 0, c4, c5c1V1+c2V2+c4V4+c5V5 
29 z1 ≠ 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z4 = 0, z5 = ωc5 ≠ 0. Let ϵ1 = 0 (c1, c2, c3, 0, c5c1V1+c2V2+c3V3+c5V5 
30 z5 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, c2, c3, c4, 0) c1V1+c2V2+c3V3+c4V4 
31 z1 ≠ 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, c2, c3, c4, c5c1V1+c2V2+c3V3+c4V4+c5V5 
S. no.CaseBasis element (y)Sub-algebra Φd
z1 ≠ 0, z2 = z3 = z4 = z5 = 0. Let ϵ1 = 0 and z1/ω = c1 (c1, 0, 0, 0, 0) V1 
z1 = z3 = z4 = 0 = z5, z2 = ωc2 ≠ 0 (0, c2, 0, 0, 0) V2 
z1 = z2 = z4 = 0 = z5, z3 = ωc3 ≠ 0 (0, 0, c3, 0, 0) V3 
z1 = z2 = z3 = z5 = 0, z4 ≠ 0. Let ϵ3 + ϵ5 = log(ωc4/z4), ϵ4 = 0 (0, 0, 0, c4, 0) V4 
z1 = z2 = z3 = z4 = 0, z5 = ωc5 ≠ 0 (0, 0, 0, 0, c5V5 
z3 = z4 = z5 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0. Let ϵ1 = 0, z1/ω = c1, z2/ω = c2 (c1, c2, 0, 0, 0) c1V1+c2V2 
z2 = z4 = z5 = 0, z1 ≠ 0, z3 = ωc3 ≠ 0. Let ϵ1 = 0, z1/ω = c1, z3/ω = c3 (c1, 0, c3, 0, 0) c1V1+c3V3 
z2 = z3 = z5 = 0, z1 = ωc1 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ3 + ϵ5 = log(ωc4/z4(c1, 0, 0, c4, 0) c1V1+c4V4 
z2 = z3 = z4 = 0, z5 = ωc5 ≠ 0, z1 ≠ 0. Let ϵ1 = 0, z1/ω = c1 (c1, 0, 0, 0, c5c1V1+c5V5 
10 z1 = z4 = z5 = 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0 (0, c2, c3, 0, 0) c2V2+c3V3 
11 z1 = z3 = z5 = 0, z2 = ωc2 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, c2, 0, c4, 0) c2V2+c4V4 
12 z1 = z3 = z4 = 0, z2 = ωc2 ≠ 0, z5 = ωc5 ≠ 0 (0, c2, 0, 0, c5c2V2+c5V5 
13 z1 = z2 = z5 = 0, z3 = ωc3 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, 0, c3, c4, 0) c3V3+c4V4 
14 z1 = z2 = z4 = 0, z3 = ωc3 ≠ 0, z5 = ωc3 ≠ 0 (0, 0, c3, 0, c5c3V3+c5V5 
15 z1 = 0, z2 = 0, z4 ≠ 0, z3 = 0, z5 = ωc5 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, 0, 0, c4, c5c4V4+c5V5 
16 z4 = z5 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0. Let ϵ1 = 0, z1/ω = c1 (c1, c2, c3, 0, 0) c1V1+c2V2+c3V3 
17 z3 = z5 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, c2, 0, c4, 0) c1V1+c2V2+c4V4 
18 z3 = z4 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0, z5 = ωc5 ≠ 0. Let ϵ1 = 0, z1/ω = c1 (c1, c2, 0, 0, c5c1V1+c2V2+c5V5 
19 z2 = z4 = 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0, z1 ≠ 0. Let ϵ1 = 0 = ϵ4 = ϵ5, ϵ3 = log(ωc4/z4(c1, 0, c3, c4, 0) c1V1+c3V3+c4V4 
20 z2 = z4 = 0, z1 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0. Let ϵ1 = 0 (c1, 0, c3, 0, c5c1V1+c3V3+c5V5 
21 z2 = z3 = 0, z1 = ωc1 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, 0, 0, c4, c5c1V1+c4V4+c5V5 
22 z1 = z5 = 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, c2, c3, c4, 0) c2V2+c3V3+c4V4 
23 z1 = z4 = 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0. (0, c2, c3, 0, c5c2V2+c3V3+c5V5 
24 z1 = z3 = 0, z2 = ωc2 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, c2, 0, c4, c5c2V2+c4V4+c5V5 
25 z5 = ωc5 ≠ 0, z1 = 0 = z2, z3 = ωc3 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, 0, c3, c4, c5c3V3+c4V4+c5V5 
26 z1 = 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ4 = 0, ϵ3 + ϵ5 = log(ωc4/z4(0, c2, c3, c4, c5c2V2+c3V3+c4V4+c5V5 
27 z2 = 0, z1 = ωc1 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, 0, c3, c4, c5c1V1+c3V3+c4V4+c5V5 
28 z1 ≠ 0, z2 = ωc2 ≠ 0, z4 ≠ 0, z5 = ωc5 ≠ 0, z3 = 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, c2, 0, c4, c5c1V1+c2V2+c4V4+c5V5 
29 z1 ≠ 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z4 = 0, z5 = ωc5 ≠ 0. Let ϵ1 = 0 (c1, c2, c3, 0, c5c1V1+c2V2+c3V3+c5V5 
30 z5 = 0, z1 ≠ 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, c2, c3, c4, 0) c1V1+c2V2+c3V3+c4V4 
31 z1 ≠ 0, z2 = ωc2 ≠ 0, z3 = ωc3 ≠ 0, z5 = ωc5 ≠ 0, z4 ≠ 0. Let ϵ1 = 0 = ϵ4, ϵ5 + ϵ3 = log(ωc4/z4(c1, c2, c3, c4, c5c1V1+c2V2+c3V3+c4V4+c5V5 
Thus, we have 14 different optimal classes of subalgebras, namely,
(29)
After obtaining these 14 different subalgebras, we have to check for the existence of similarity solutions for each of them. The collection of all the obtained similarity solutions discussed in Sec. V constitute the optimal solutions to the considered problem.

The set of fundamental Eqs. (2)(8) may admit a similarity solution by an appropriate choice of the subalgebra Φd, d = 1, 2, 3, …, 31, from Table III.

The subalgebras Φ23=c2V2+c3V3+c5V5, Φ29=c1V1+c2V2+c3V3+c5V5, and Φ26=c2V2+c3V3+c4V4+c5V5 or Φ31=c1V1+c2V2+c3V3+c4V4+c5V5 generate the same invariant subalgebra under translational transformation. In this case, the associated Lagrange auxiliary equation is expressed as
(30)
From integration of this equation, we obtain
(31)
and the similarity transformations
(32)
where θ = (σ22 + 2a)/a is a real constant and A is a dimensional constant.
The shock wave moves in an outward direction from the axis of the cylinder, and if we denote the distance of the shock front from the axis of symmetry by R(t), which is the shock radius, and the shock velocity by Us, then we have
(33)
From Eqs. (3), (8), and (14), the flow variables just ahead of the shock front vary as
(34)
where p*=(1/σ2)[(χσ2)μh*2/2+Gm*ρ*/2v*2ρ*/2], m* = 2πρ*/(σ1 + 2), ρ*, v*, w*, h* are dimensional constants, σ1, σ2, σ3, σ4 are real constants, and σ1, σ2, σ3 satisfy the relations σ1 + σ2 + 1 = 0 and 2σ3 + σ1 + 2σ2 = 0 in the presence of a magnetic field. Also, in the absence of a magnetic field (i.e., h* = 0), pa=p*R2(σ1+1), where p*=12(σ1+1)/[(v*2ρ*Gm*ρ*)] and σ1, σ3 satisfy the relation 2σ3σ1 − 2 = 0.
The boundary conditions at the shock front (η = 1) are obtained as
(35)
provided that σ1 = −2/θ, σ3 = (θ − 1)/θ, σ4 = (σ44 + a)/ = (θ − 1)/θ, and σ2 = −(θ − 2)/θ.
For the existence of a similarity solution, the Alfvén Mach number AM and shock Mach number SM must be constant, and therefore the expressions for AM and SM are obtained as
(36)
The relationship between SM and AM in the presence of a magnetic field (i.e., for AM20) is obtained as
(37)
and in the absence of magnetic field (i.e., for AM2=0), this reduces to
(38)
where Rp=v*Aσ31/θ and Gp=Gm*Aσ1 are taken as the nondimensional parameters for rotation and gravitation, respectively.
Using Eqs. (31) and (33), the similarity transformations (32) can be written as
(39)
where
(40)
Using Eq. (31) and the transformations (39) in the fundamental Eqs. (2)(8), we obtain a system of ordinary differential equations (ODEs) as
(41)
(42)
(43)
(44)
(45)
(46)
(47)
(48)
where Γp=[16ϰθ/3κo(R*)2](ρ*Aσ1)1/2 is the radiative dimensionless heat transfer parameter. It is necessary to take τ1=12 and τ2 = 2 for the existence of a similarity solution.
Using Eqs. (35) and (40), the jump conditions across the shock front are obtained as
(49)
The components of the vorticity vector in nondimensional form are expressed as lr = LrR/Us, lθ = LθR/Us, and lz = LzR/Us. With the help of Eqs. (13), (33), (39), (42), and (43), the expressions to calculate lr, lθ, lz are obtained as
(50)
where So=w*Aσ41/θ is a nondimensional constant.
The subalgebras Φ16=c1V1+c2V2+c3V3 and Φ30=c1V1+c2V2+c3V3+c4V4 in Table III generate the same invariant subspace, which can be prove by using translation invariance. Thus, the associated characteristic equation corresponding to Φ16 or Φ30 is given by
(51)
From integration of this equation, we obtain the similarity variable as
(52)
and the similarity transformations as
(53)
where δ = σ22/c and B are dimensional constants. The shock radius and shock front velocity are expressed as
(54)
Using Eqs. (34) and (54), the Alfvén–Mach number AM and shock Mach number SM are expressed as
(55)

In the presence of a magnetic field, the flow variables just ahead of the shock front vary according to Eq. (34), where p*=(1/σ2)[(χσ2)μh*2/2+Gm*ρ*/2v*2ρ*/2], ρ*, v*, w*, h* are dimensional constants, σ1, σ2, σ3, σ4 are real constants, and σ1, σ2, σ3 satisfy the relations σ1 + σ2 + 1 = 0 and 2σ3 + σ1 + 2σ2 = 0.

The shock Mach number SM and Alfvén Mach number AM in the presence of a magnetic field (i.e., AM20) are related by
(56)
In the absence of a magnetic field (i.e., AM2=0), this expression reduces to
(57)
and the initial values of the flow variables are given by Eq. (34) with h* = 0; in this case, pa=p*R2(σ1+1), where p*=v*2ρ*Gm*ρ*/[2(σ1+1)] and σ1, σ3 satisfy the relation 2σ3σ1 − 2 = 0, whereas the shock Mach number SM becomes nondimensional with the condition σ1 = 0.
In Eqs. (56) and (57), Re = v*/δ and Ge = Gπρ*/δ2 are taken as the nondimensional parameters for rotation and gravitation, respectively, in case of an exponential-law shock path. From Eq. (56), it is clear that Re = 0 cannot hold, because, if it did, then the shock Mach number would be imaginary i.e., the shock wave would not exist. Thus, for the existence of a shock wave, a necessary condition is that Re2>(1+χ)AM2+Ge in Eq. (56) and Re2>Ge in Eq. (57). Also, our problem can be reduced to the nonmagnetic and nongravitating case (i.e., AM2=0 and Ge = 0), or the nonmagnetic but gravitating case (i.e., AM2=0 and Ge ≠ 0), or the magnetic but nongravitating case (i.e., AM20 and Ge = 0). It cannot, however, be reduced to the nonrotating case (i.e., Re = 0). The jump conditions across the shock front (i.e., at η* = 1) are obtained by using the transformations (53) and Eqs. (16), (34), and (54), as
(58)
where m* = πρ*, σ1 = 0, σ2 = −1, σ3 = 1, and σ4 = σ44/.
With the help of Eqs. (34) and (54), the similarity transformations (53) are expressed as
(59)
where
(60)
Using Eq. (52) and the similarity transformations (59) in the fundamental Eqs. (2)(8) gives a system of ODEs that, after simplification, can be expressed as
(61)
(62)
(63)
(64)
(65)
(66)
(67)
(68)
where
and Γe=16ϰδ3κ0(R*)2 is the nondimensional radiation parameter, and it has been necessary to take τ1 = −1 and τ2 = −2 for the existence of a similarity solution.
By using Eqs. (58) and (60), we obtain the jump conditions at the shock front as
(69)
The vorticity components in nondimensional form are lr = LrR/Us, lθ = LθR/Us, and lz = LzR/Us, and the expressions to calculate these are obtained with the use of Eqs. (13), (52), (59), (62), and (63) as
(70)
where So*=w*/δ is a nondimensional constant, and σ44 = 1.
Here, the subalgebra is Φ6=c1V1+c2V2 in Table III, and the associate characteristic equation is given by
(71)

Proceeding as in case II, we ontain the same similarity variable and similarity transformations as given in Eqs. (31) and (59), respectively, and the shock jump condition as given in Eq. (58) with σ4 = 0 for the initial values of variables as given in Eq. (34). In this case, by using the similarity transformations (59) in the system of fundamental Eqs. (2)(8) and jump conditions (58), we obtain the same system of ODEs (61)(68) and shock jump conditions (69) with σ4 = 0.

In the present case, from Eq. (70), for σ4 = 0, the vorticity components in nondimensional form lr = lrR/Us, lθ = lθR/Us, and lz = LzR/Us are obtained as
(72)

Here, the subalgebras are Φ5, Φ9, Φ12, Φ14, Φ15, Φ17, Φ18, Φ20, Φ21, Φ24, Φ25, Φ27, and Φ28 in Table III. The similarity variable and the similarity transformations can be obtained, but the shock jump condition is not invariant for these subalgebras. Thus, a similarity solution does not exist in this case.

Here, the subalgebras are Φ1, Φ2, Φ3, Φ4, Φ7, Φ8, Φ10, Φ11, Φ13, Φ19, Φ22 in Table III. For these, a similarity variable and similarity transformations cannot be obtained, and hence a similarity solution does not exist in this case.

The systems of ODEs (41)(48) and (61)(68) for the cases of a power-law shock path (PLSP) and an exponential-law shock path (ELSP), respectively, are integrated numerically with respect to the similarity variables η and η* with boundary conditions (49) and (69), respectively. A fourth-order Runge–Kutta method is used for the integrations. For these numerical calculations, we have taken the values of the physical parameters as follows:39,56,60,70,72

  • PLSP: γ=43, 53; Gp = 0, 0.01, 0.025; AM2=0.08, 0.085; Rp = 0, 0.06, 0.08; So = 0, 1; Γp = 2, 4.

  • ELSP: γ=43, 53; Ge = 0, 0.04, 0.06; AM2=0, 0.02, 0.03, 0.035, 0.04; Re = 0.55, 0.58, 0.6; So*=1; Γe = 0.5, 0.7.

For the PLSP, the value Gp = 0 represents the nongravitating case, for which our solution corresponds to that obtained by Singh.70 The values Rp = 0, So = 0, and Gp = 0 for the PLSP represent the nongravitating and nonrotating case, for which our solution corresponds to that obtained by Vishwakarma et al.39 for a perfect gas. The values AM2=0, Rp = 0, So = 0 and Gp = 0 represent the nonmagnetic, nonrotating, and nongravitating case, for which our solution corresponds to that obtained by Singh.59 Thus, our present work for the PLSP represents a generalization of that in Refs. 39, 59, and 70. For the ELSP, our work represents a generalization of that by Singh70 in that it additionally considers the effect of gravitation. Also, Singh70 assumed an azimuthal magnetic field, but we have considered both axial and azimuthal magnetic fields. The value γ=43 corresponds to a relativistic gas and γ=53 to a fully ionized gas, both of which are appropriate in an the astrophysical context.69 These values of γ characterize the most general ranges that are seen in real stars. Also, the value of γ (γ < 1 or γ > 1) is in the general range corresponding to different fluids present in the interstellar medium.79,80 Low values of γ are expected to lead to the formation of dense clusters of low-mass stars, whereas γ > 1 probably results in the formation of isolated and massive stars.81 In our problem, the value of the parameter AM2 is taken in the range 0.01<AM2<1 because Rosenau and Frankenthal43,82 have shown that the magnetic field has a remarkable effect on flow variables and on shocks when AM20.01. Also, for AM2=1, the shock wave degenerates into a weak discontinuity that does not compress the gas. The choices of the nondimensional gravitational parameter Gp=Gm*Aσ1 for the PLSP and Ge = Gπρ*/δ2 for the ELSP are typical values, because all quantities in this parameter are positive and no restrictions on the values of m*, A, ρ* and δ are obtained, whereas π is a fixed constant and the value of G depends on the mass of the object. Thus, the values Gp = 0, 0.01, 0.025 and Ge = 0, 0.04, 0.06 are taken such that these parameters have significant effects on the flow variables and on shocks. The values of the rotational parameters Rp and Re are taken so that the conditions for the existence of shock waves as obtained from Eqs. (37) and (56), respectively, are satisfied: Rp2>(χσ2)AM2+Gp/θ2σ2/γ for the PLSP and Re2>(χ+1)AM2+Ge+2/γ for the ELSP. The choices of the values of the nondimensional radiative heat transfer parameters Γp=[16ϰθ/3κo(R*)2](ρ*Aσ1)1/2 for the PLSP and Γe=16ϰδ/3κo(R*)2 for the ELSP are typical. The values of the radiative heat transfer parameters Γp and Γe are inversely related to the absorption coefficient or opacity of the medium.

Tables IV and V show the values of the density ratio at the shock front and the position of the inner boundary surface (IBS) for the PLSP, and Tables VI and VII show the corresponding values for the ELSP. Physically, the IBS may result from an exploding wire in a swirling detonation engine, the surface of a stellar corona, condensed explosives, or a diaphragm retaining a very high-pressure driving gas. With the help of the data in these tables, we have plotted the flow variables between the IBS and shock front in Figs. 1 and 2 for a PLSP and in Figs. 3 and 4 for the ELSP.

TABLE IV.

For the PLSP, the values of β and ηp for different values of Rp, Gp, and AM2, with χ = 1, θ = 1.4, Γp = 4, and γ=53.

RpGpAM2βηp
0.08 0.286 354 0.928 926 
0.085 0.298 305 0.921 712 6 
0.01 0.08 0.294 428 0.913 388 
0.085 0.306 431 0.905 639 
0.025 0.08 0.306 805 0.887 264 
0.085 0.318 876 0.878 726 
0.06 0.08 0.280 745 0.939 016 
0.085 0.292 655 0.932 190 
0.01 0.08 0.288 716 0.924 503 
0.085 0.300 683 0.917 729 
0.025 0.08 0.300 944 8 0.899 981 
0.085 0.312 985 0.891 813 
0.08 0.08 0.276 434 0.946 367 
0.085 0.288 311 0.939 853 
0.01 0.08 0.284 323 0.932 646 
0.085 0.296 26 0.925 569 
0.025 0.08 0.296 433 0.909 342 
0.085 0.384 48 0.901 462 
2.305 9.9 0.5 0.661 646 0.993 885 3 
0.6 0.829 01 0.849 039 5 
10 0.5 0.740 471 0.889 249 1 
0.6 0.914 301 0.791 873 
RpGpAM2βηp
0.08 0.286 354 0.928 926 
0.085 0.298 305 0.921 712 6 
0.01 0.08 0.294 428 0.913 388 
0.085 0.306 431 0.905 639 
0.025 0.08 0.306 805 0.887 264 
0.085 0.318 876 0.878 726 
0.06 0.08 0.280 745 0.939 016 
0.085 0.292 655 0.932 190 
0.01 0.08 0.288 716 0.924 503 
0.085 0.300 683 0.917 729 
0.025 0.08 0.300 944 8 0.899 981 
0.085 0.312 985 0.891 813 
0.08 0.08 0.276 434 0.946 367 
0.085 0.288 311 0.939 853 
0.01 0.08 0.284 323 0.932 646 
0.085 0.296 26 0.925 569 
0.025 0.08 0.296 433 0.909 342 
0.085 0.384 48 0.901 462 
2.305 9.9 0.5 0.661 646 0.993 885 3 
0.6 0.829 01 0.849 039 5 
10 0.5 0.740 471 0.889 249 1 
0.6 0.914 301 0.791 873 
TABLE V.

For the PLSP, the values of β and ηp for different values of χ, γ, θ, and Γp, with Rp = 0.08, AM2=0.08, and Gp = 0.2.

χγθβΓpηp
43 1.3 0.300 851 0.955 091 
0.903 736 
1.5 0.329 034 0.926 761 
0.812 226 
53 1.3 0.300 851 0.950 934 
0.895 308 
1.5 0.329 034 0.919 319 
0.804 934 
10 0.728 979 7 
50 0.691 769 4 
43 1.3 0.413 258 0.838 811 
0.828 904 
1.35 0.433 122 0.810 663 
0.798 072 
53 1.3 0.413 258 0.814 824 
0.803 618 
1.35 0.433 122 0.782 794 
0.768 757 
10 0.118 781 
50 0.082 388 21 
χγθβΓpηp
43 1.3 0.300 851 0.955 091 
0.903 736 
1.5 0.329 034 0.926 761 
0.812 226 
53 1.3 0.300 851 0.950 934 
0.895 308 
1.5 0.329 034 0.919 319 
0.804 934 
10 0.728 979 7 
50 0.691 769 4 
43 1.3 0.413 258 0.838 811 
0.828 904 
1.35 0.433 122 0.810 663 
0.798 072 
53 1.3 0.413 258 0.814 824 
0.803 618 
1.35 0.433 122 0.782 794 
0.768 757 
10 0.118 781 
50 0.082 388 21 
TABLE VI.

For the ELSP, the values of β and ηp* for different values of Ge, Re, and Γe, with γ=53, AM2=0.02, and χ = 1.

GeReβΓeηp*
0.55 0.305 259 0.5 0.890 711 
0.7 0.881 794 
0.58 0.336 149 0.5 0.882 916 
0.7 0.872 791 
0.04 0.55 0.269 593 0.5 0.900 045 
0.7 0.892 412 
0.6 0.321 139 0.5 0.886 751 
0.7 0.877 251 
0.06 0.55 0.252 128 0.5 0.904 725 
0.7 0.897 657 
0.6 0.303 003 0.5 0.893 928 
0.7 0.882 596 
10 0.858 657 
50 0.857 041 
100 0.856 84 
GeReβΓeηp*
0.55 0.305 259 0.5 0.890 711 
0.7 0.881 794 
0.58 0.336 149 0.5 0.882 916 
0.7 0.872 791 
0.04 0.55 0.269 593 0.5 0.900 045 
0.7 0.892 412 
0.6 0.321 139 0.5 0.886 751 
0.7 0.877 251 
0.06 0.55 0.252 128 0.5 0.904 725 
0.7 0.897 657 
0.6 0.303 003 0.5 0.893 928 
0.7 0.882 596 
10 0.858 657 
50 0.857 041 
100 0.856 84 
TABLE VII.

For the ELSP, the values of β and ηp* for different values of χ, γ, and AM2, with SM = 2.3, Γe = 0.7, and Ge = 0.06.

χγAM2βηp*
43 0.283 554 0.907 305 
0.02 0.324 382 0.893 521 
0.04 0.359 229 0.882 947 
53 0.226 843 0.923 303 
0.02 0.273 417 0.904 876 
0.03 0.293 032 0.897 967 
43 0.283 554 0.907 305 
0.035 0.350 922 0.867 267 
0.04 0.359 229 0.863 317 
53 0.226 843 0.922 330 
0.02 0.273 417 0.891 318 
0.04 0.311 126 0.871 090 1 
χγAM2βηp*
43 0.283 554 0.907 305 
0.02 0.324 382 0.893 521 
0.04 0.359 229 0.882 947 
53 0.226 843 0.923 303 
0.02 0.273 417 0.904 876 
0.03 0.293 032 0.897 967 
43 0.283 554 0.907 305 
0.035 0.350 922 0.867 267 
0.04 0.359 229 0.863 317 
53 0.226 843 0.922 330 
0.02 0.273 417 0.891 318 
0.04 0.311 126 0.871 090 1 
FIG. 1.

Flow variable distribution between shock front and IBS for the PLSP with azimuthal magnetic field (χ = 1), with γ = 5/3, θ = 1.4, and Γp = 4: (a) radial velocity; (b) density; (c) azimuthal velocity; (d) axial velocity; (e) pressure; (f) magnetic field; (g) mass; (h) radiation heat flux; (i) azimuthal vorticity component; (j) axial vorticity component; 1, Rp=0,Gp=0,AM2=0.08; 2, Rp=0,Gp=0,AM2=0.085; 3, Rp=0,Gp=0.01,AM2=0.08; 4, Rp=0,Gp=0.01,AM2=0.085; 5, Rp=0,Gp=0.025,AM2=0.08; 6, Rp=0.06,Gp=0,AM2=0.08; 7, Rp=0.06,Gp=0,AM2=0.085; 8, Rp=0.06,Gp=0.01,AM2=0.08; 9, Rp=0.06,Gp=0.01,AM2=0.085; 10, Rp=0.06,Gp=0.025,AM2=0.08; 11, Rp=0.08,Gp=0,AM2=0.08; 12, Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 1.

Flow variable distribution between shock front and IBS for the PLSP with azimuthal magnetic field (χ = 1), with γ = 5/3, θ = 1.4, and Γp = 4: (a) radial velocity; (b) density; (c) azimuthal velocity; (d) axial velocity; (e) pressure; (f) magnetic field; (g) mass; (h) radiation heat flux; (i) azimuthal vorticity component; (j) axial vorticity component; 1, Rp=0,Gp=0,AM2=0.08; 2, Rp=0,Gp=0,AM2=0.085; 3, Rp=0,Gp=0.01,AM2=0.08; 4, Rp=0,Gp=0.01,AM2=0.085; 5, Rp=0,Gp=0.025,AM2=0.08; 6, Rp=0.06,Gp=0,AM2=0.08; 7, Rp=0.06,Gp=0,AM2=0.085; 8, Rp=0.06,Gp=0.01,AM2=0.08; 9, Rp=0.06,Gp=0.01,AM2=0.085; 10, Rp=0.06,Gp=0.025,AM2=0.08; 11, Rp=0.08,Gp=0,AM2=0.08; 12, Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 2.

Flow variable distribution between shock front and IBS for the PLSP, with Rp = 0.08, AM2=0.08, and Gp = 0.2: (a) radial velocity; (b) density; (c) azimuthal velocity; (d) axial velocity; (e) pressure; (f) magnetic field; (g) mass; (h) radiation heat flux; (i) azimuthal vorticity component; (j) axial vorticity component; 1, χ=0,γ=43,θ=1.3,Γp=2; 2, χ=0,γ=43,θ=1.3,Γp=4; 3, χ=0,γ=43,θ=1.5,Γp=2; 4, χ=0,γ=53,θ=1.3,Γp=2; 5, χ=0,γ=53,θ=1.3,Γp=4; 6, χ=0,γ=53,θ=1.5,Γp=2; 7, χ=1,γ=43,θ=1.3,Γp=2; 8, χ=1,γ=43,θ=1.3,Γp=4; 9, χ=1,γ=43,θ=1.35,Γp=2; 10, χ=1,γ=53,θ=1.3,Γp=2; 11, χ=1,γ=53,θ=1.3,Γp=4; 12, χ=1,γ=53,θ=1.35,Γp=2.

FIG. 2.

Flow variable distribution between shock front and IBS for the PLSP, with Rp = 0.08, AM2=0.08, and Gp = 0.2: (a) radial velocity; (b) density; (c) azimuthal velocity; (d) axial velocity; (e) pressure; (f) magnetic field; (g) mass; (h) radiation heat flux; (i) azimuthal vorticity component; (j) axial vorticity component; 1, χ=0,γ=43,θ=1.3,Γp=2; 2, χ=0,γ=43,θ=1.3,Γp=4; 3, χ=0,γ=43,θ=1.5,Γp=2; 4, χ=0,γ=53,θ=1.3,Γp=2; 5, χ=0,γ=53,θ=1.3,Γp=4; 6, χ=0,γ=53,θ=1.5,Γp=2; 7, χ=1,γ=43,θ=1.3,Γp=2; 8, χ=1,γ=43,θ=1.3,Γp=4; 9, χ=1,γ=43,θ=1.35,Γp=2; 10, χ=1,γ=53,θ=1.3,Γp=2; 11, χ=1,γ=53,θ=1.3,Γp=4; 12, χ=1,γ=53,θ=1.35,Γp=2.

Close modal
FIG. 3.

Flow variable distribution between shock front and IBS for the ELSP with azimuthal magnetic field (χ = 1), with γ = 5/3 and AM2=0.02: (a) radial velocity; (b) density; (c) azimuthal velocity; (d) axial velocity; (e) pressure; (f) magnetic field; (g) mass; (h) radiation flux; (i) azimuthal vorticity component; (j) axial vorticity component; 1, Ge = 0, Re = 0.55, Γe = 0.5; 2, Ge = 0, Re = 0.55, Γe = 0.7; 3, Ge = 0, Re = 0.58, Γe = 0.5; 4, Ge = 0.04, Re = 0.55, Γe = 0.5; 5, Ge = 0.04, Re = 0.55, Γe = 0.7; 6, Ge = 0.04, Re = 0.6, Γe = 0.7; 7, Ge = 0.06, Re = 0.55, Γe = 0.5.

FIG. 3.

Flow variable distribution between shock front and IBS for the ELSP with azimuthal magnetic field (χ = 1), with γ = 5/3 and AM2=0.02: (a) radial velocity; (b) density; (c) azimuthal velocity; (d) axial velocity; (e) pressure; (f) magnetic field; (g) mass; (h) radiation flux; (i) azimuthal vorticity component; (j) axial vorticity component; 1, Ge = 0, Re = 0.55, Γe = 0.5; 2, Ge = 0, Re = 0.55, Γe = 0.7; 3, Ge = 0, Re = 0.58, Γe = 0.5; 4, Ge = 0.04, Re = 0.55, Γe = 0.5; 5, Ge = 0.04, Re = 0.55, Γe = 0.7; 6, Ge = 0.04, Re = 0.6, Γe = 0.7; 7, Ge = 0.06, Re = 0.55, Γe = 0.5.

Close modal
FIG. 4.

Flow variable distribution between shock front and IBS for the ELSP, with SM = 2.3, Γe = 0.7, and Ge = 0.06: (a) radial velocity; (b) density; (c) azimuthal velocity; (d) axial velocity; (e) pressure; (f) magnetic field; (g) mass; (h) radiation flux; (i) azimuthal vorticity component; (j) axial vorticity component; 1, χ=0,γ=43,AM2=0; 2, χ=0,γ=43,AM2=0.02; 3, χ=0,γ=43,AM2=0.04; 4, χ=0,γ=53,AM2=0; 5, χ=0,γ=53,AM2=0.02; 6, χ=0,γ=53,AM2=0.03; 7, χ=1,γ=43,AM2=0; 8, χ=1,γ=43,AM2=0.035; 9, χ=1,γ=43,AM2=0.04; 10, χ=1,γ=53,AM2=0; 11, χ=1,γ=53,AM2=0.02; 12, χ=1,γ=53,AM2=0.04.

FIG. 4.

Flow variable distribution between shock front and IBS for the ELSP, with SM = 2.3, Γe = 0.7, and Ge = 0.06: (a) radial velocity; (b) density; (c) azimuthal velocity; (d) axial velocity; (e) pressure; (f) magnetic field; (g) mass; (h) radiation flux; (i) azimuthal vorticity component; (j) axial vorticity component; 1, χ=0,γ=43,AM2=0; 2, χ=0,γ=43,AM2=0.02; 3, χ=0,γ=43,AM2=0.04; 4, χ=0,γ=53,AM2=0; 5, χ=0,γ=53,AM2=0.02; 6, χ=0,γ=53,AM2=0.03; 7, χ=1,γ=43,AM2=0; 8, χ=1,γ=43,AM2=0.035; 9, χ=1,γ=43,AM2=0.04; 10, χ=1,γ=53,AM2=0; 11, χ=1,γ=53,AM2=0.02; 12, χ=1,γ=53,AM2=0.04.

Close modal

From Figs. 1 and 2, we observe that on moving from the shock front toward the IBS, the flow variables u/ub, ρ/ρb, h/hb, J/Jb, and lz increase, but v/vb, w/wb, p/pb, m/mb, and lθ decrease for the PLSP. (These variables exhibit a similar flow field nature in the contour plots for the PLSP, as can be seen in Figs. 514.) Also, for the ELSP, Figs. 3(a), 3(c)3(e), 3(g)3(i), and 3(j) and Figs. 4(a)4(e) and 4(g)4(j) show that the variables u/ub, p/pb, and J/Jb increase, but v/vb, w/wb, m/mb, and lθ decrease.The variables ρ/ρb and lz increase with increasing axial magnetic field, but with increasing azimuthal magnetic field they first increase and then decrease rapidly after attaining a maximum on moving from the shock front to the IBS [see panels (b) and (j) of Figs. 3 and 4]. Figures 3(f) and 4(f) show that the variable h/hb first increases and then decreases after reaching a maximum in the flow field region behind the shock.

FIG. 5.

Contour plots of reduced radial velocity behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, and Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 5.

Contour plots of reduced radial velocity behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, and Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 6.

Contour plots of reduced density behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, and Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 6.

Contour plots of reduced density behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, and Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 7.

Contour plots of reduced azimuthal velocity behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 7.

Contour plots of reduced azimuthal velocity behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 8.

Contour plots of reduced axial velocity behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 8.

Contour plots of reduced axial velocity behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 9.

Contour plots of reduced pressure behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 9.

Contour plots of reduced pressure behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 10.

Contour plots of reduced magnetic field behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 10.

Contour plots of reduced magnetic field behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 11.

Contour plots of reduced mass behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 11.

Contour plots of reduced mass behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 12.

Contour plots of reduced radiation flux behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 12.

Contour plots of reduced radiation flux behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 13.

Contour plots of reduced axial vorticity vector behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 13.

Contour plots of reduced axial vorticity vector behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal
FIG. 14.

Contour plots of reduced azimuthal vorticity vector behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

FIG. 14.

Contour plots of reduced azimuthal vorticity vector behind cylindrical shock front for the PLSP with azimuthal magnetic field, with γ=53, θ = 1.4, Γp = 4: (a) Rp=0.06,Gp=0.01,AM2=0.06; (b) Rp=0.06,Gp=0.01,AM2=0.085; (c) Rp=0.06,Gp=0.025,AM2=0.085; (d) Rp=0.08,Gp=0.01,AM2=0.085.

Close modal

In Figs. 514, we present contour plots to provide a visual representation of the flow variables such as density and velocity, and to see the effect of the rotational parameter, gravitational parameter, and magnetic field parameter behind the shock front for the PLSP. In these contour plots, dark blue color represents lower values of η and light blue represents higher values. We have analyzed the distribution of flow variables from higher to lower values of η. The horizontal lines in the contour plots represent the values of the flow variables at different values of η.

Comparison of panels (a) and (b) in Figs. 514 illustrates the variation in the flow variables when the strength of the magnetic field is increased, comparison of panels (b) and (c) in each figure illustrates the variation in flow variables when the value of the gravitational parameter is increased, and comparison of panels (b) and (d) illustrates the variation in flow variables when the value of the rotational parameter is increased. To reveal these variations in the flow variables, we select a value of η = 0.96, and we find that when the magnetic field strength and the value of the gravitational parameter are increased, u/ub, ρ/ρb, h/hb, J/Jb, and lθ decrease, whereas v/vb, w/wb, p/pb, m/mb, and lz increase. The rotational parameter has the reverse effects on these flow variables compared with the effects of the magnetic field and the gravitational parameter. Similar effect of these parameters on the flow variables are also found from Fig. 1 (see Secs. VI AVI C). Similarly, we can plot contours of the flow variables for different values of the physical parameters to compare with the results obtained from Figs. 14.

The distance between the positions of the IBS and the shock front increases with increasing Gp, whereas it decreases with increasing Ge, i.e., the shock decays with Gp for the PLSP, and grows with Ge for the ELSP (see Tables IV and VI). The flow variables v/vb, w/wb, p/pb, m/mb, and lθ increase, whereas u/ub, ρ/ρb, h/hb, J/Jb, and lz decrease in the whole flow field region behind the shock front for the PLSP with Gp (see Fig. 1). The effects of Gp on u/ub, ρ/ρb, m/mb, p/pb, and J/Jb are similar to the results obtained by Singh and Vishwakarma60 in the magnetic and nonmagnetic cases. The flow variables u/ub, v/vb, w/wb, m/mb, J/Jb, and lθ decrease in general with increasing Ge [see Figs. 3(a), 3(c), 3(d), and 3(g)3(i)]. The variables ρ/ρb, h/hb, and lz increase near the shock front and decrease near the IBS with increasing Ge, but Ge has the reverse effects on p/pb behind the shock front [see Figs. 3(b), 3(e), 3(f), and 3(j)].

From the above results, we observe that with an increase in the value of the gravitational parameter Gp or Ge, the flow field region between the shock front and the IBS expands for the PLSP, whereas it contracts for the ELSP. For the PLSP, this reduces the compression of the shocked region, and the mass flux m/mb decreases; however, for the ELSP, the shocked region is highly compressed, and so the mass flux m/mb increases.

The distance between the position of the IBS and shock front decreases with increasing Rp, whereas it increases with increasing Re, i.e., the shock strength grows with Rp for the PLSP and decays with Re for the ELSP (see Tables IV and VI). The flow variables u/ub, ρ/ρb, h/hb, J/Jb and lz increase, whereas v/vb, w/wb, p/pb, m/mb, and lθ decrease behind the shock front for the PLSP with Rp (see Fig. 1). The flow variables u/ub, v/vb, w/wb, m/mb, J/Jb, and lθ increase in general with increasing Re [see Figs. 3(a), 3(c), 3(d), and 3(g)3(i)]. The variables ρ/ρb, h/hb and lz decrease near the shock front and increase near the IBS with increasing Re [see Figs. 3(b), 3(f), and 3(j)]. The variable p/pb increases near the shock front and decreases near the IBS with increasing Re in the nongravitating case, but in the gravitating case, it decreases in the whole flow field with increasing Re [see Fig. 3(e)].

Physically, this means that with an increase in the value of the rotational parameter Rp or Re, the width of the flow field region between the shock front and the IBS is reduced owing to a decrease in compressibility for the PLSP, and so the shock become stronger. Also, the width of the flow field region between the shock front and the IBS is enlarged owing to an increase in compressibility for the ELSP, and so the shock wave decay with increasing Re becomes weaker.

The distance between the position of the IBS and the shock front increases with increasing AM2, i.e., the shock strength decays in the presence of a magnetic field for both the PLSP and ELSP (see Tables V and VII). For the PLSP, the flow variables u/ub, ρ/ρb, h/hb, and J/Jb, lz decrease; however, lθ, m/mb, p/pb, v/vb, and w/wb increase with increasing AM2 in the whole flow field region behind the shock front (see Fig. 1). The effects of AM2 on u/ub, ρ/ρb, h/hb, J/Jb, and p/pb are similar to the results obtained by Vishwakarma et al.39 for a nonrotating nongravitating perfect gas. For the ELSP, the variables u/ub, v/vb, w/wb, m/mb, and lθ increase, but p/pb decreases with increasing AM2 in the whole flow field region [see Figs. 4(a), 4(c)4(e), 4(g), and 4(i)]. The variables ρ/ρb and lz decrease and J/Jb and h/hb increase in the presence of an axial magnetic field [see Figs. 4(b), 4(f), 4(h), and 4(j)]. The variables ρ/ρb, h/hb, and lz decrease near the shock front and increase near the IBS, but J/Jb decreases in the presence of an azimuthal magnetic field [see Figs. 4(b), 4(f), 4(h), and 4(j)]. For the ELSP, there has yet to be any examination of the effects of a magnetic field on the flow variables or on the strength of a shock wave in the presence of a radiation heat flux using the Lie symmetry method.

Physically, this means that an increase in the strength of the magnetic field AM2 has the effect of increasing the compressibility, and hence the shock wave strength decreases and the flow field region between the shock front and the IBS increases for both the PLSP and the ELSP.

The distance between the positions of the IBS and the shock front increases with Γp and Γe, i.e., the shock wave decays with an increase in the value of the radiation parameter for both the PLSP and ELSP (see Tables V and VI). For the PLSP, the flow variables u/ub, ρ/ρb, h/hb, J/Jb, and lz decrease with increasing Γp, whereas p/pb, v/vb, w/wb, and m/mb, lθ increase (see Fig. 2). In this case, the effects of Γp on u/ub, u/ub, ρ/ρb, m/mb, p/pb, and J/Jb are similar to the results obtained by Singh and Vishwakarma60 in both the nonrotating and nonmagnetic and the magnetic and rotating cases. For the ELSP, u/ub and J/Jb, decrease with increasing Γe, but v/vb, w/wb, m/mb and lθ increase [see Figs. 3(a), 3(c), 3(d), and 3(g)3(i)]. The variables ρ/ρb, h/hb, and lz decrease near the shock front and increase near the IBS with increasing Γe, but Γe has the reverse effect on p/pb behind the shock front [see Figs. 3(b), 3(e), 3(f), and 3(j)].

Physically, this means that with an increase in the value of the radiation parameter Γp or Γe, the shock strength decreases, because the shocked region between the shock front and IBS expands for both the PLSP and ELSP. Owing to the reduced transfer of heat energy, the radiation flux J/Jb decreases in the whole shocked region for both the PLSP and ELSP. For the PLSP, the shock strength is strongly influenced by small values of the radiation parameter in the presence of an axial magnetic field in comparison with the case of an azimuthal magnetic field. At higher values of the radiation parameter, the shock strength is influenced by the presence of an azimuthal magnetic field (see Table V). For the ELSP, the shocked region is more compressed by an increase in the value of Γe in a self-gravitating gas than in a nongravitating gas (see Table VI).

The distance between the positions of the IBS and shock front increases with γ for the PLSP but decreases for the ELSP, i.e., with increasing γ, the shock strength decays for the PLSP but grows for the ELSP (see Tables V and VII). With increasing γ, the flow variables ρ/ρb, u/ub, J/Jb, h/hb, and lz decrease, whereas v/vb, w/wb, m/mb, p/pb, and lθ increase in the whole flow field region for PLSP (see Fig. 2). For the ELSP, u/ub, v/vb, w/wb, m/mb, J/Jb, and lθ decrease with increasing γ, whereas p/pb increase in general behind the shock front [see Figs. 4(a), 4(c)4(e), and 4(g)4(i)]. In the presence of an azimuthal magnetic field, ρ/ρb and h/hb decrease, but ρ/ρb increases in the nonmagnetic case [see Figs. 4(b) and 4(f)]. With increasing γ, h/hb decreases, whereas ρ/ρb increases near the shock front and decreases near the IBS in the presence of an axial magnetic field [see Figs. 4(b) and 4(j)]. The variable lz increases with increasing γ in the absence of magnetic field or in the presence of an axial magnetic field, but in the presence of an azimuthal magnetic field, it increases near the shock front but decreases near the IBS [see Fig. 4(j)].

The shock strength decreases with increasing θ (see Table V). The flow variables ρ/ρb, J/Jb, w/wb, m/mb, h/hb, and lz decrease with increasing θ, but v/vb and p/pb increase [see Figs. 2(b)2(h) and 2(j)]. The variables u/ub and lθ decrease throughout in the presence of an axial magnetic field. With increasing θ, in the presence of an azimuthal magnetic field, u/ub increases near the shock front and decreases near the IBS, but θ has the reverse effect on lθ [see Figs. 2(a) and 2(i)].

The propagation of a shock wave in a self-gravitating ideal gas in the presence of a magnetic field, radiation heat flux, and rotation of the medium has been studied here using the Lie symmetry method. The advantage of this method is that it provides all possible similarity solutions (optimal solutions) of the problem under consideration. We have discussed the effects on the shock wave and the flow variables between the shock front and IBS of different values of physical parameters, such as those associated with gravitation, rotation, radiation, ambient magnetic field strength, adiabatic index, and similarity exponent. Understanding the propagation of a magnetogasdynamic shock wave in self-gravitating gas with radiative heat flux is essential for the study of a variety of astrophysical phenomena. Explosions in astrophysical and geophysical environments generate shock waves in which the shock may follow a power law shock path (PLSP) or an exponential law shock path (ELSP).78,83 Gravity is the dominant force involved in interactions at small scales in the interstellar medium and governs the formation, shape, and trajectory of astronomical objects, such as in the formation of stars and planetary systems and in the creation of supernova remnants.28,29,45,81 The present study has also included an analysis of data from exploding wire experiments in a rotating electrically conducting medium, and cylindrically symmetric hypersonic flow problems associated with meteors or reentry vehicles.9,37

The following conclusions of this study are summarized from the results shown in Tables IVVII and Figs. 14:

  1. The shock strength decays in the presence of a radiation heat flux and of a magnetic field in for both a PLSP and an ELSP. Physically, energy transformation through shock waves is less in the presence of a magnetic field.

  2. With an increase in the value of the rotational parameter, the shock strength is increased, whereas it is reduced with an increase in the value of the adiabatic index or the gravitational parameter for a PLSP. For an ELSP, the rotation and gravitational parameters and the adiabatic exponent have effects on the shock strength that are the reverse of those for a PLSP.

  3. The effects of the rotational parameter on shock strength are the reverse of those of the adiabatic index and the gravitational parameter.

  4. When the magnetic field is changed from axial to azimuthal, the flow field region between the shock front and the IBS increases for both PLSP and ELSP cases, i.e., the shock strength is decreased in the presence of an azimuthal magnetic field compared with an axial magnetic field.

  5. When the value of the rotational parameter increases, the pressure decreases near the shock front and increases near the IBS in the nongravitating case, but it decreases everywhere in the flow field in the gravitating case for a PLSP. Also, with an increase in the value of the rotational parameter, the pressure decreases in general for an ELSP.

The results of the current study may be of particular interest in the following contexts.

The interstellar medium, which consists of dust, gas, and charged particles, makes up around 10%–15% of the galactic disk’s total mass.84 It is widely believed that new stars and planetary systems form primarily within interstellar molecular clouds. An anomalous dissipation mechanism generates shock waves in these clouds. These shock waves propagate in the clouds and trigger transportation phenomena, leading to regions of increased density, enabling gravitational self-compression (or Jeans instability).85 Observations reveal a strong correlation between the presence of molecular clouds and star formation. The main question here is to determine which factor plays a more critical role: condensation processes influenced by the strength of shock propagation or gravitational self-compression driven by the shock waves.

Stellar explosions leads to supernovae, after which about 1015 erg of kinetic energy is dispersed in the ambient medium. The interaction of the supersonically (∼5000–10 000 km/s) moving ejecta (stellar material) with the circumstellar/interstellar medium (CSM/ISM) results in the formation of a shock wave, which creates a shell of shock-heated plasma. The interaction of the outermost ejecta with the CSM/ISM marks the beginning of the formation of a supernova remnant.86,87 During a supernova explosion, high amounts of heat radiation are emitted in the ISM, which influences the propagation of shock waves and ejected stellar material.

The authors have no conflicts to disclose.

G. Nath: Formal analysis (equal); Investigation (equal); Methodology (equal); Project administration (lead); Resources (lead); Software (lead); Supervision (lead); Validation (equal); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (equal). P. Upadhyay: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article.

For interested readers, a complete derivation of the optimal system for the subalgebras are given as follows:

As V1, V2, V3, V4, and V5 are independent vector fields of a five-dimensional Lie algebra, every vector of that Lie algebra can be expressed as a linear combination of independent vectors Vα, α = 1, 2, …, 5. Thus, a general vector V of the Lie algebra can be written as
(A1)
where c1 = c, c2 = σ22, c3 = σ44, c4 = k, and c5 = a.
The adjoint action of V1 on V can be expressed with the help of Table II as
where superscript Tr indicates the transpose of the matrix, and
(A2)
The adjoint action of V2 on V can be expressed with the help of Table II as
where the matrix
(A3)
The adjoint action of V3 on V can be expressed with the help of Table II as
where the matrix
(A4)
The adjoint action of V4 on V can be expressed with the help of Table II as
where the matrix
(A5)
The adjoint action of V5 on V can be expressed with the help of Table II as
where the matrix
(A6)
We can then express the general adjoint matrix D with the help of the matrices A1, A2, A3, A4, and A5 from Eqs. (A2)(A6) as
(A7)
1.
W. W.
Roberts
, “
Large-scale shock formation in spiral galaxies and its implications on star formation
,”
Astrophys. J.
158
,
123
143
(
1969
).
2.
A. A.
Rumyantsev
, “
Propagation of shock waves in a rotating star
,”
Sov. Astron. AJ
16
(
4
),
610
614
(
1973
), available at https://adsabs.harvard.edu/full/1973SvA....16..610R.
3.
G. S.
Bisnovatyi-Kogan
and
S. A.
Silich
, “
Shock-wave propagation in the nonuniform interstellar medium
,”
Rev. Mod. Phys.
67
,
661
712
(
1995
).
4.
T.
Yamasaki
and
S.
Yamada
, “
Effects of rotation on the revival of a stalled shock in supernova explosions
,”
Astrophys. J.
623
,
1000
1010
(
2005
).
5.
F.
Alzahrani
and
M. I.
Khan
, “
Entropy generation and joule heating applications for Darcy Forchheimer flow of Ree-Eyring nanofluid due to double rotating disks with artificial neural network
,”
Alexandria Eng. J.
61
(
5
),
3679
3689
(
2022
).
6.
P.
Chaturani
, “
Strong cylindrical shocks in a rotating gas
,”
Appl. Sci. Res.
23
,
197
211
(
1971
).
7.
L.
Sedov
,
Similarity and Dimensional Methods in Mechanics
(
Academic Press
,
New York
,
1959
).
8.
V. A.
Levin
and
G. A.
Skopina
, “
Detonation wave propagation in rotational gas flows
,”
J. Appl. Mech. Tech. Phys.
45
,
457
460
(
2004
).
9.
G.
Nath
, “
Self-similar solutions for unsteady flow behind an exponential shock in an axisymmetric rotating dusty gas
,”
Shock Waves
24
,
415
428
(
2014
).
10.
S. I.
Pai
,
Magnetogasdynamics and Plasma Dynamics
(
Springer
,
1963
).
11.
W.
Baumjohann
and
R. A.
Treumann
,
Basic Space Plasma Physics
(
Imperial College Press
,
London
,
1999
).
12.
J. S.
Shang
, “
Recent research in magneto-aerodynamics
,”
Prog. Aerosp. Sci.
37
(
1
),
1
20
(
2001
).
13.
D.
Grasso
and
H. R.
Rubinstein
, “
Magnetic fields in the early universe
,”
Phys. Rep.
348
,
163
266
(
2001
).
14.
R. I.
Sujith
, “
Magnetohydrodynamic shock wave formation: Effect of area and density variation
,”
Phys. Plasmas
12
,
052116
(
2005
).
15.
S. I.
Popel
and
A. A.
Gisko
, “
Charged dust and shock phenomena in the Solar System
,”
Nonlinear Processes Geophys.
13
,
223
229
(
2006
).
16.
K.
Fujisawa
,
H.
Okawa
,
Y.
Yamamoto
, and
S.
Yamada
, “
Effects of rotation and magnetic field on the revival of a stalled shock in supernova explosions
,”
Astrophys. J.
872
,
155
172
(
2019
).
17.
S. I.
Popel
,
L. M.
Zelenyi
, and
A. V.
Zakharov
, “
Dusty plasma in the solar system: Celestial bodies without atmosphere
,”
Plasma Phys. Rep.
49
,
1006
1013
(
2023
).
18.
J. M.
Rax
,
R.
Gueroult
, and
N. J.
Fisch
, “
Rotating Alfvén waves in rotating plasmas
,”
J. Plasma Phys.
89
,
905890613
(
2023
).
19.
M. A. Z.
Raza
,
R.
Tabassum
,
E. R.
EI-Zahar
,
M.
Shoaib
,
M. I.
Khan
,
M. Y.
Malik
, and
S.
Qayyum
, “
Intelligent computing through neural networks for entropy generation in MHD third-grade nanofluid under chemical reaction and viscous dissipation
,”
Waves Random Complex Media
35
,
2502
2526
(
2025
).
20.
S. R.
Mishra
,
T. C.
Sun
,
B. C.
Rout
,
M.
Ijaz Khan
,
M.
Kbiri Alaoui
, and
S.
Ullah Khan
, “
Control of dusty nanofluid due to the interaction on dust particles in a conducting medium: Numerical investigation
,”
Alexandria Eng. J.
61
(
4
),
3341
3349
(
2022
).
21.
R. M.
Lock
and
A. J.
Mestel
, “
Annular self-similar solutions in ideal magnetogasdynamics
,”
J. Plasma Phys.
74
,
531
554
(
2008
).
22.
D. I.
Pullin
,
W.
Mostert
,
V.
Wheatley
, and
R.
Samtaney
, “
Converging cylindrical shocks in ideal magnetohydrodynamics
,”
Phys. Fluids
26
,
097103
(
2014
).
23.
W.
Mostert
,
V.
Wheatley
,
R.
Samtaney
, and
D. I.
Pullin
, “
Effects of magnetic fields on magnetohydrodynamic cylindrical and spherical Richtmyer-Meshkov instability
,”
Phys. Fluids
27
,
104102
(
2015
).
24.
S. D.
Ramsey
,
E. M.
Schmidt
,
Z. M.
Boyd
,
J. F.
Lilieholm
, and
R. S.
Baty
, “
Converging shock flows for a Mie-Grüneisen equation of state
,”
Phys. Fluids
30
,
046101
(
2018
).
25.
I.
Giron
,
S.
Balberg
, and
M.
Krief
, “
Solutions of the imploding shock problem in a medium with varying density
,”
Phys. Fluids
33
,
066105
(
2021
).
26.
A. N.
Golubyatnikov
and
S. D.
Kovalevskaya
, “
Shock wave acceleration in a magnetic field
,”
Fluid Dyn.
49
,
844
848
(
2014
).
27.
G.
Nath
, “
Magnetogasdynamic shock wave generated by a moving piston in a rotational axisymmetric isothermal flow of perfect gas with variable density
,”
Adv. Space Res.
47
(
9
),
1463
1471
(
2011
).
28.
Y. B.
Zel’dovich
and
Y. P.
Raizer
,
Physics of Shock Waves and High Temperature Hydrodynamic Phenomena
(
Academic Press
,
New York
,
1967
), Vol.
II
.
29.
S. I.
Pai
,
Radiation Gas Dynamics
(
Springer
,
New York
,
1966
).
30.
R. E.
Marshak
, “
Effect of radiation on shock wave behavior
,”
Phys. Fluids
1
,
24
29
(
1958
).
31.
A.
Raza
,
S. U.
Khan
,
M. I.
Khan
,
S.
Farid
,
T.
Muhammad
,
M. I.
Khan
, and
A. M.
Galal
, “
Fractional order simulations for the thermal determination of graphene oxide (GO) and molybdenum disulphide (MoS2) nanoparticles with slip effects
,”
Case Stud. Therm. Eng.
28
,
101453
(
2021
).
32.
G.
Sulochana
,
C. V.
Prasad
,
S. K.
Bhatti
,
V. v.
Venu Madhav
,
K. K.
Saxena
,
M. I.
Khan
,
Z.
Aloui
,
C.
Prakash
, and
M. I.
Khan
, “
Impact of multi-walled carbon nanotubes (MWCNTs) on hybrid biodiesel blends for cleaner combustion in ci engines
,”
Energy
303
,
131911
(
2024
).
33.
M. I.
Khan
,
Usman
,
A.
Ghaffari
,
S. U.
Khan
,
Y. M.
Chu
, and
S.
Qayyum
, “
Optimized frame work for Reiner–Philippoff nanofluid with improved thermal sources and Cattaneo–Christov modifications: A numerical thermal analysis
,”
Int. J. Mod. Phys. B
35
(
6
),
2150083
(
2021
).
34.
M. W.
Nazir
,
T.
Javed
,
N.
Ali
,
M.
Nazeer
, and
M. I.
Khan
, “
Theoretical investigation of thermal analysis in aluminum and titanium alloys filled in nanofluid through a square cavity having the uniform thermal condition
,”
Int. J. Mod. Phys. B
36
(
22
),
2250140
(
2022
).
35.
J. R. A. J.
Nicastro
, “
Similarity analysis of the radiative gas dynamic equations with spherical symmetry
,”
Phys. Fluids
13
,
2000
2006
(
1970
).
36.
D.
Summers
, “
An idealized model of a magnetohydrodynamic spherical blast wave applied to a flare produced shock in the solar wind
,”
Astron. Astrophys.
45
,
151
158
(1975), available at https://ui.adsabs.harvard.edu/abs/1975A%26A....45..151S/abstract.
37.
G. J.
Hutchens
, “
Approximate cylindrical blast theory: Near-field solutions
,”
J. Appl. Phys.
77
,
2912
2915
(
1995
).
38.
Y. M.
Chu
,
M. S.
Hashmi
,
N.
Khan
,
M. I.
Khan
,
S.
Kadry
, and
Z.
Abdelmalek
, “
Thermophoretic particles deposition features in thermally developed flow of Maxwell fluid between two infinite stretched disks
,”
J. Mater. Res. Technol.
9
(
6
),
12889
12898
(
2020
).
39.
J. P.
Vishwakarma
,
A. K.
Maurya
, and
A. K.
Singh
, “
Cylindrical shock waves in a non-ideal gas with radiation heat-flux and magnetic field
,”
Modell. Meas. Control B
80
,
35
52
(2011), available at https://scholar.google.com/scholar?hl=en&as_sdt=0%2C5&q=Cylindrical+shock+waves+in+a+non-ideal+gas+with+radiation+heat-flux+and+magnetic+field%2C%E2%80%9D+Model&btnG.
40.
J. P.
Vishwakarma
and
N.
Patel
, “
Magnetogasdynamic cylindrical shock waves in a rotating nonideal gas with radiation heat flux
,”
J. Eng. Phys. Thermophys.
88
,
521
530
(
2015
).
41.
J. P.
Vishwakarma
and
G.
Nath
, “
Propagation of a cylindrical shock wave in a rotating dusty gas with heat conduction and radiation heat flux
,”
Phys. Scr.
81
,
045401
(
2010
).
42.
G.
Nath
, “
Self-similar flow of a rotating dusty gas behind the shock wave with increasing energy, conduction and radiation heat flux
,”
Adv. Space Res.
49
(
1
),
108
120
(
2012
).
43.
P.
Rosenau
and
S.
Frankenthal
, “
Equatorial propagation of axisymmetric magnetohydrodynamic shocks
,”
Phys. Fluids
19
,
1889
1899
(
1976
).
44.
P.
Rosenau
and
S.
Frankenthal
, “
Propagation of magnetohydrodynamic shocks in a thermally-conducting medium
,”
Phys. Fluids
21
,
559
566
(
1978
).
45.
D. F.
Cioffi
,
C. F.
Mckee
, and
E.
Bertschinger
, “
Dynamics of radiative supernova remnants
,”
Astrophys. J.
334
,
252
265
(
1988
).
46.
O.
Nath
,
S.
Ojha
, and
H. S.
Takhar
, “
A study of stellar point explosion in a self-gravitating radiative magneto-hydrodynamic medium
,”
Astrophys. Space Sci.
183
,
135
145
(
1991
).
47.
G.
Nath
and
J. P.
Vishwakarma
, “
Similarity solution for the flow behind a shock wave in a non-ideal gas with heat conduction and radiation heat-flux in magnetogasdynamics
,”
Commun. Nonlinear Sci. Numer. Simul.
19
(
5
),
1347
1365
(
2014
).
48.
G.
Nath
, “
Cylindrical shock wave generated by a moving piston in a rotational axisymmetric non-ideal gas with conductive and radiative heat-fluxes in the presence of azimuthal magnetic field
,”
Acta Astronaut.
156
,
100
112
(
2019
).
49.
P. A.
Carrus
,
P. A.
Fox
,
F.
Haas
, and
Z.
Kopal
, “
The propagation of shock waves in a stellar model with continuous density distribution
,”
Astrophys. J.
113
,
496
518
(
1951
).
50.
Z.
Kopal
, “
The propagation of shock waves in self-gravitating gas spheres
,”
Astrophys. J.
120
,
159
171
(
1954
).
51.
M. V.
Penston
, “
Dynamics of self-gravitating gaseous spheres—II: Collapses of gas spheres with cooling and the behaviour of polytropic gas spheres
,”
Mon. Not. R. Astron. Soc.
145
(
4
),
457
485
(
1969
).
52.
J. K.
Truelove
,
R. I.
Klein
,
C. F.
Mckee
,
J.
Holliman
II
,
L. H.
Howell
,
J. A.
Greenough
, and
D. T.
Woods
, “
Self-gravitational hydrodynamics with three-dimensional adaptive mesh refinement: Methodology and applications to molecular cloud collapse and fragmentation
,”
Astrophys. J.
495
(
2
),
821
852
(
1998
).
53.
R. B.
Larson
, “
Numerical calculations of the dynamics of a collapsing proto-star
,”
Mon. Not. R. Astron. Soc.
145
(
3
),
271
295
(
1969
).
54.
M. H.
Rogers
, “
Analytic solutions for the blast-wave problem with an atmosphere of varying density
,”
Astrophys. J.
125
,
478
493
(
1957
).
55.
A.
Sakurai
, “
Propagation of spherical shock waves in stars
,”
J. Fluid Mech.
1
,
436
453
(
1956
).
56.
G.
Nath
, “
Approximate analytical solution for the propagation of shock waves in self-gravitating perfect gas via power series method: Isothermal flow
,”
J. Astrophys. Astron.
41
,
21
(
2020
).
57.
G.
Nath
,
J. P.
Vishwakarma
,
V. K.
Shrivastava
, and
A. K.
Sinha
, “
Propagation of magnetogasdynamic shock waves in a self-gravitating gas with exponentially varying density
,”
J. Theor. Appl. Phys.
7
,
15
(
2013
).
58.
G.
Nath
and
S.
Singh
, “
Flow behind magnetogasdynamic exponential shock wave in self-gravitating gas
,”
Int. J. Non-Linear Mech.
88
,
102
108
(
2017
).
59.
J. B.
Singh
, “
Self-similar cylindrical shock waves with radiation heat flux
,”
Astrophys. Space Sci.
102
,
263
268
(
1984
).
60.
J. B.
Singh
and
P. R.
Vishwakarma
, “
A self-similar flow of self-gravitating gas behind a shock wave with radiation heat flux, I
,”
Astrophys. Space Sci.
90
,
153
159
(
1983
).
61.
J. P.
Vishwakarma
and
A. K.
Singh
, “
A self-similar flow behind a shock wave in a gravitating or non-gravitating gas with heat conduction and radiation heat-flux
,”
J. Astrophys. Astron.
30
,
53
69
(
2009
).
62.
G.
Nath
,
M.
Dutta
, and
R. P.
Pathak
, “
An exact solution for the propagation of shock waves in self-gravitating perfect gas in the presence of magnetic field and radiative heat flux
,”
Modell., Meas. Control, B
86
,
907
927
(
2017
).
63.
G. W.
Bluman
and
J. D.
Cole
,
Similarity Methods for Differential Equations
(
Springer
,
New York
,
1974
).
64.
G.
Nath
and
R. K.
Singh
, “
Lie symmetry method to discuss the optimal system of Lie subalgebras and similarity solutions for shock propagation in a perfectly conducting dusty gas with magnetic field in rotating medium
,”
Int. J. Geom. Methods Mod. Phys.
(published online)
(
2024
).
65.
G. W.
Bluman
and
S.
Kumei
,
Symmetries and Differential Equations
(
Springer Science & Business Media
,
1989
).
66.
L. V. E.
Ovsiannikkov
,
Group Analysis of Differential Equations
(
Academic Press
,
New York
,
1982
).
67.
P. J.
Olver
,
Application of Lie Group to Differential Equation
(
Springer
,
New York
,
1993
).
68.
K. O.
Wambura
,
M. E. O.
Okoya
, and
T. J. O.
Aminer
, “
Lie symmetry analysis and the optimal system of non-linear fourth order evolution equation
,”
Int. J. Sci. Eng. Appl. Sci.
5
,
1
6
(
2019
), available at https://ijseas.com/volume5/v5i3/ijseas20190301.pdf.
69.
G.
Nath
and
A.
Devi
, “
Optimal classification and similarity solution for unsteady flow behind a shock wave in a perfectly conducting dusty gas with magnetic field using group invariance method
,”
Int. J. Non-Linear Mech.
154
,
104443
(
2023
).
70.
S.
Singh
, “
Similarity solutions for strong magnetogasdynamic cylindrical shock wave in rotating axisymmetric ideal gas with radiation heat flux using Lie group theoretic method
,”
Ric. Mat.
73
,
1895
1917
(
2024
).
71.
L. V. E.
Ovsiannikkov
,
Group Analysis of Differential Equations
(
Academic Press
,
New York
,
1982
).
72.
G.
Nath
and
A.
Maurya
, “
Optimal system of solution using group invariance technique for shock wave in a non-ideal self-gravitating gas in rotating medium in presence of magnetic field
,”
Z. Naturforsch. A
78
(
8
),
721
742
(
2023
).
73.
G.
Nath
and
V. S.
Kadam
, “
Lie symmetry analysis and optimal system for shock wave in a self-gravitating rotating ideal gas under the effect of magnetic field and monochromatic radiation
,”
Int. J. Geom. Methods Mod. Phys.
21
(
3
),
2450058
(
2024
).
74.
X.
Hu
,
Y.
Li
, and
Y.
Chen
, “
A direct algorithm of one-dimensional optimal system for the group invariant solutions
,”
J. Math. Phys.
56
(
5
),
053504
(
2015
).
75.
S.
GhoshHajra
,
S.
Kandel
, and
S. P.
Pudasaini
, “
Optimal systems of Lie subalgebras for a two-phase mass flow
,”
Int. J. Non-Linear Mech.
88
,
109
121
(
2017
).
76.
S.
Mandal
,
S.
Sil
, and
S.
Ghosh
, “
Lie symmetries and optimal classifications with certain modal approaches for the three-dimensional gas-dynamical equations
,”
Chaos, Solitons Fractals
181
,
114672
(
2024
).
77.
M.
Nadjafikhah
and
A. H.
Zaeim
, “
Symmetry analysis of wave equation on hyperbolic space
,”
Int. J. Nonlinear Sci.
11
,
35
43
(
2011
).
78.
D. D.
Laumbach
and
R. F.
Probstein
, “
A point explosion in a cold exponential atmosphere. Part 2. Radiating flow
,”
J. Fluid Mech.
40
,
833
858
(
1970
).
79.
M.
Spaans
and
J.
Silk
, “
The polytropic equation of state of interstellar gas clouds
,”
Astrophys. J.
538
,
115
120
(
2000
).
80.
Y.
Li
,
R. S.
Klessen
, and
M.-M. M.
Low
, “
The formation of stellar clusters in turbulent molecular clouds: Effects of the equation of state
,”
Astrophys. J.
592
,
975
985
(
2003
).
81.
S. E.
Woosley
,
A.
Heger
, and
T. A.
Weaver
, “
The evolution and explosion of massive stars
,”
Rev. Mod. Phys.
74
,
1015
1071
(
2002
).
82.
P.
Padoan
and
A.
Nordlund
, “
A super-Alfvénic model of dark clouds
,”
Astrophys. J.
526
,
279
294
(
1999
).
83.
H. A.
Bethe
, “
Supernova shock. VIII
,”
Astrophys. J.
490
,
765
771
(
1997
).
84.
K. M.
Ferrière
, “
The interstellar environment of our galaxy
,”
Rev. Mod. Phys.
73
,
1031
(
2001
).
85.
G.
Nath
, “
Analytical solution for unsteady adiabatic and isothermal flows behind the shock wave in a rotational axisymmetric mixture of perfect gas and small solid particles
,”
Z. Naturforsch. A
76
,
853
(
2021
).
86.
D. C.
Ellison
,
S. P.
Reynolds
,
K.
Borkowski
,
R.
Chevalier
,
D. P.
Cox
,
J. R.
Dickel
,
R.
Pisarski
,
J.
Raymond
,
S. R.
Spangler
,
H. J.
Volk
, and
J. P.
Wefel
, “
Supernova remnants and the physics of strong shock waves
,”
Publ. Astron. Soc. Pac.
106
,
780
797
(
1994
).
87.
J.
Vink
,
Physics and Evolution of Supernova Remnants
(
Springer Nature
,
Switzerland
,
2020
).