Skip to main content
Log in

Magnetogasdynamic exponential shock wave in a self-gravitating, rotational axisymmetric non-ideal gas under the influence of heat-conduction and radiation heat-flux

  • Published:
Ricerche di Matematica Aims and scope Submit manuscript

Abstract

Similarity solutions are obtained for one-dimensional cylindrical shock wave in a self-gravitating, rotational axisymmetric non-ideal gas with azimuthal or axial magnetic field in the presence of conductive and radiative heat fluxes. The total energy of the wave is non-constant. It is obtained that the increase in the Cowling number, in the parameters of radiative as well as conductive heat transfer and the parameter of the non-idealness of the gas have a decaying effect on the shock wave however increase in the value of gravitational parameter has reverse effect on the shock strength. It is manifested that the presence of azimuthal magnetic field removes the singularities which arise in some cases of the presence of axial magnetic field. Also, it is observed that the effect of the parameter of non-idealness of the gas is diminished by increasing the value of the gravitational parameter.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  1. Sedov, L. I. (1959), Similarity and Dimensional Methods in Mechanics, Academic Press, New York

    Google Scholar 

  2. Marshak, R. E. (1958), Effect of radiation on shock wave behavior, Phys. Fluids, 1, 24–29

    ADS  MathSciNet  Google Scholar 

  3. Elliott, L.A.: Similarity methods in radiation hydrodynamics. Proc. Roy. Soc. Lond. Series A 258, 287–301 (1960)

    ADS  MathSciNet  Google Scholar 

  4. Wang, K. C. (1964) , The ‘piston problem’ with thermal radiation, J. Fluid Mech., 20, 447–455

    ADS  MathSciNet  Google Scholar 

  5. Helliwell, J. B. (1969), Self-similar piston problems with radiative heat transfer, J. Fluid Mech., 37, 497–512

    ADS  Google Scholar 

  6. NiCastro, J. R. (1970), Similarity analysis of the radiative gas dynamic equations with spherical symmetry, Phys. Fluids, 13, 2000–2006

    ADS  Google Scholar 

  7. Ghoniem, A. F., Kamel, M. M., Berger, S. A., and Oppenheim, A. K. (1982), Effect of internal heat transfer on the structure of self-similar blast waves, J. Fluid Mech., 117, 473–491

    ADS  Google Scholar 

  8. Nath, G., Sahu, P.K.: Unsteady adiabatic flow behind a cylindrical shock in a rotational axisymmetric non-ideal gas under the action of monochromatic radiation. Proc. Eng. 144, 1226–1233 (2016)

    Google Scholar 

  9. Nath, G., Sahu, P.K.: Flow behind an exponential shock wave in a rotational axisymmetric non-ideal gas with conduction and radiation heat flux. Int. J. Appl. Comput. Math. 3, 2785–2801 (2017)

    MathSciNet  Google Scholar 

  10. Nath, G., Sahu, P.K.: Propagation of a cylindrical shock wave in a mixture of a non-ideal gas and small solid particles under the action of monochromatic radiation. Combust. Explos. Shock Waves 53(3), 298–308 (2017)

    Google Scholar 

  11. Nath, G., Sahu, P.K., Chaurasia, S.: An exact solution for the propagation of cylindrical shock waves in a rotational axisymmetric nonideal gas with axial magnetic field and radiative heat flux. Modell. Meas. Control B. 87(4), 236–243 (2018)

    Google Scholar 

  12. Sahu, P. K. (2020). Spherical and cylindrical shocks in a non-ideal dusty gas with magnetic field under the action of heat conduction and radiation heat flux. Physics of Fluids, 32(6), 066104

    ADS  CAS  Google Scholar 

  13. Hartmann, L. (1998), Accretion Processes in Star Formation, Cambridge University Press, Cambridge

    Google Scholar 

  14. Balick, B. and Frank, A. (2002), Shapes and shaping of planetary nebulae, Annu. Rev. Astron. Astrophys. 40, 439

    ADS  Google Scholar 

  15. Nath, G., Sahu, P.K., Dutta, M.: Magnetohydrodynamic cylindrical shock in a rotational axisymmetric non-ideal gas under the action of monochromatic radiation. Proc. Eng. 127, 1126–1133 (2015)

    Google Scholar 

  16. Nath, G., & Sahu, P. K. (2016). Flow behind an exponential shock wave in a rotational axisymmetric perfect gas with magnetic field and variable density. SpringerPlus, 5(1), 1509

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Nath, G., Sahu, P.K.: Similarity solution for the flow behind a cylindrical shock wave in a rotational axisymmetric gas with magnetic field and monochromatic radiation. Ain Shams Eng. J. 9(4), 1151–1159 (2018)

    Google Scholar 

  18. Nath, G., Sahu, P.K., Chaurasia, S.: Self-similar solution for the flow behind an exponential shock wave in a rotational axisymmetric non-ideal gas with magnetic field. Chin. J. Phys. 58, 280–293 (2019)

    CAS  Google Scholar 

  19. Sahu, P. K. (2020). Shock wave driven out by a piston in a mixture of a non-ideal gas and small solid particles under the influence of azimuthal or axial magnetic field. Brazilian Journal of Physics, 50(5), 548–565

    ADS  Google Scholar 

  20. Shiota, D., Kataoka, R., (2016), Magnetohydrodynamic simulation of interplanetary propagation of multiple coronal mass ejections with internal magnetic flux rope (SUSANOO-CME), Space Weather, 14: 56–75. DOI : 10.1002/2015SW001308

    Article  ADS  Google Scholar 

  21. Verma, M. K. (2004). Statistical theory of magnetohydrodynamic turbulence: recent results. Physics Reports, 401(5–6), 229–380

    ADS  MathSciNet  CAS  Google Scholar 

  22. Wu, P. and van der Wal, W., (2003), Postglacial sealevels on a spherical, self-gravitating viscoelastic earth: effects of lateral viscosity variations in the upper mantle on the inference of viscosity contrasts in the lower mantle. Earth and Planetary Science Letters, 211(1–2), pp. 57–68

    ADS  CAS  Google Scholar 

  23. Colwell, J.E., Esposito, L.W., Sremčević, M.: Self-gravity wakes in Saturn’s A ring measured by stellar occultations from Cassini. Geophys. Res. Lett. 33(7), (2006)

  24. Mitrovica, J., Tamisiea, M., Davis, J. and Milne, G. (2001), Recent mass balance of polar ice sheets inferred from patterns of global sea-level change. Nature 409, 1026–1029

    ADS  CAS  PubMed  Google Scholar 

  25. Carrus, P., Fox, P., Hass, F., Kopal, Z. (1951), The propagation of shock waves in a stellar model with continuous density distribution, Astrophys. J. 113, 496

    ADS  MathSciNet  Google Scholar 

  26. Purohit, S.C. (1974), Self-similar homothermal flow of self-gravitating case behind shock wave, J. Phys. Soc. Jpn. J. Phys. Soc. Jpn. 36, 288

    ADS  Google Scholar 

  27. Singh, J.B., Vishwakarma, P.R. (1983), Self-similar solutions in the theory of flare-ups in novae, I, Astrophys. Space Sci. 95, 99

    ADS  Google Scholar 

  28. Sahu, P. K. (2018). Self-similar solution of spherical shock wave propagation in a mixture of a gas and small solid particles with increasing energy under the influence of gravitational field and monochromatic radiation. Communications in Theoretical Physics, 70(2), 197

    ADS  MathSciNet  CAS  Google Scholar 

  29. Sahu, P.K.: Similarity solution for a spherical shock wave in a non-ideal gas under the influence of gravitational field and monochromatic radiation with increasing energy. Math. Methods Appl. Sci. 42(14), 4734–4746 (2019)

    ADS  MathSciNet  Google Scholar 

  30. Sahu, P.K.: Propagation of an exponential shock wave in a rotational axisymmetric isothermal or adiabatic flow of a self-gravitating non-ideal gas under the influence of axial or azimuthal magnetic field. Chaos Solitons Fractals 135, 109739 (2020)

    MathSciNet  Google Scholar 

  31. Zel’dovich, Ya., B. and Raizer, Yu. P. : Physics of shock waves and high temperature hydrodynamic phenomena, vol. II. Academic Press, New York (1967)

    Google Scholar 

  32. Lee, T. S., Chen, T. (1968), Hydrodynamic interplanetary shock waves, Planet. Space Sci. 16:1483–502

    ADS  Google Scholar 

  33. Summers, D. (1972), An idealized model of a magnetohydrodynamic spherical blast wave applied to a flare produced shock in the solar wind, Astron. Astophys. 45:151–158

    ADS  Google Scholar 

  34. Sagdeev, R. Z., (1966) Reviews of Plasma Physics, Consultants Bureau, New York, 4:23–93

    Google Scholar 

  35. Chen, F.F.: Introduction to plasma physics. Plenum, New York (1974).. ([chapter 8])

  36. Chaturani, P. (1970), Strong cylindrical shocks in a rotating gas, Appl. Sci. Res., 23, 197–211

    MathSciNet  Google Scholar 

  37. Ganguly, A., and Jana, M. (1998), Propagation of shock wave in self-gravitating radiative magnetohydrodynamic non-uniform rotating atmosphere, Bull. Cal. Math. Soc., 90, 77–82

    Google Scholar 

  38. Nath, O., Ojha, S. N., and Takhar, H. S. (1999), Propagation of a shock wave in a rotating interplanetary atmosphere with increasing energy, J. Mhd. Plasma Res., 8, 269–282

    Google Scholar 

  39. Nath, G., & Sahu, P. K. (2017). Self-similar solution of a cylindrical shock wave under the action of monochromatic radiation in a rotational axisymmetric dusty gas. Communications in Theoretical Physics, 67(3), 327

    ADS  CAS  Google Scholar 

  40. Sahu, P. K. (2017). Cylindrical shock waves in rotational axisymmetric non-ideal dusty gas with increasing energy under the action of monochromatic radiation. Physics of Fluids, 29(8), 086102

    ADS  CAS  Google Scholar 

  41. Sahu, P.K.: Unsteady flow behind an MHD exponential shock wave in a rotational axisymmetric non-ideal gas with conductive and radiative heat fluxes. Intelligent Techniques and Applications in Science and Technology, vol. 12, pp. 1049–1059. Springer, Berlin (2020)

    Google Scholar 

  42. Landau, L.D., Lifshitz, E.M.: Course of theoretical physics. Elsevier, Amsterdam (2013)

    Google Scholar 

  43. Henderson, L.F.: General laws for propagation of shock waves through matter. Handb. Shock Waves 1, 144–183 (2001)

    Google Scholar 

  44. Zhao, N., Mentrelli, A., Ruggeri, T., & Sugiyama, M. (2011). Admissible shock waves and shock-induced phase transitions in a van der Waals fluid. Physics of fluids, 23(8), 086101

    ADS  Google Scholar 

  45. Levin, V. A. and Skopina, G. A. (2004), Detonation wave propagation in rotational gas flows, J. Appl. Mech. Tech. Phys., 45, 457–460

    ADS  MathSciNet  Google Scholar 

  46. Rosenau, P., Frankenthal, S.: Shock disturbances in a thermally conducting solar wind. Astrophys. J. 208, 633–637 (1976a)

    ADS  Google Scholar 

  47. Rosenau, P., and Frankenthal, S., (1978) Propagation of magnetohydrodynamic shocks in a thermally conducting medium Phys. Fluids 21, 559–566

    ADS  Google Scholar 

  48. Kamel, M.M., Khater, H.A., Siefien, H.G., Rafat, N.M., Oppenheim, A.K. (1977), A self similar solution for blast waves with transport properties. Acta Astronaut. 4, 425–437

    ADS  Google Scholar 

  49. Rao, M. P. R. and Ramana, B. V. (1976), Unsteady flow of a gas behind an exponential shock, J. Math. Phys. Sci., 10, 465–476

    Google Scholar 

  50. Sahu, P.K.: Similarity solution for the flow behind an exponential shock wave in a rotational axisymmetric non-ideal gas under the influence of gravitational field with conductive and radiative heat fluxes. Intelligent Techniques and Applications in Science and Technology, vol. 12, pp. 1060–1070. Springer, Berlin (2020)

    Google Scholar 

  51. Laumbach, D. D. and Probstein, R. F. (1970), A point explosion in a cold exponential atmosphere. Part 2. Radiating flow, J. Fluid Mech., 40, 833–858

    ADS  Google Scholar 

  52. Freeman R. A. and Craggs J. D., (1969) Shock waves from spark discharges, J. Phys. D: Appl. Phys. 2, 421–427

    ADS  Google Scholar 

  53. Onsi M, Przysiezniak H, Pearson JM (1994) Equation of state of homogeneous nuclear matter and the symmetry coefficient. Phys. Rev. C 50: 460–468

    ADS  CAS  Google Scholar 

  54. Casali RH, Menezes DP (2010) Adiabatic index of hot and cold compact objects. Braz. J. Phys. 40: 166–171

    CAS  Google Scholar 

  55. Rosenau, P. and Frankenthal, S. (1976b), Equatorial propagation of axisymmetric magnetohydrodynamic shocks. Physics of Fluids 19, 1889–1899

    ADS  Google Scholar 

  56. Korolev, A.S., Pushkar, E.A.: ollision of an interplanetary shock wave with theearth’s bow shock. Hydrodyn. Parameters Mag. Field Fluid Dyn. 49, 270 (2014). https://doi.org/10.1134/S001546281402015X

    Article  CAS  Google Scholar 

  57. Lin, Shao-Chi (1954), Cylindrical shock waves produced by instantaneous energy release, J. Appl. Phys. 25, 54

    ADS  Google Scholar 

  58. Higashino, F. (1983) Characteristic method applied to blast waves in a dusty gas. Z Naturforsch 38:399–406

    ADS  Google Scholar 

  59. Liberman, M. A., Velikovich, A. L., (1989) Self-similar spherical expansion of a laser plasma or detonation products into a low-density ambient gas. Phys Fluids 1 :1271–1276

    Google Scholar 

  60. Smirnov, N.N., Nikitin, V.F.: Modeling and simulation of hydrogen combustion in engines. Int. J. Hydrog. Energy 39, 1122–1136 (2014)

    CAS  Google Scholar 

  61. Smirnov, N.N., Nikitin, V.F., Dushin, V.R., Filippov, Y.G., Nerchenko, V.A., Khadem, J.: Combustion onset in non-uniform dispersed mixtures. Acta Astron. 115, 94–101 (2015)

    CAS  Google Scholar 

  62. Pomroning, G. C. (1973), The equations of radiation hydrodynamics, International Series of Monographs in Natural Philosophy, vol. 54, Pergaman Press, Oxford

    Google Scholar 

  63. Anisimov, S. I. and Spiner, O. M. (1972), Motion of an almost ideal gas in the presence of a strong point explosion, J. Appl. Math. Mech., 36, 883–887

    Google Scholar 

  64. Chandrasekhar, S. (1939), An introduction to the study of Stellar structure. University Chicago Press, Chicago

    Google Scholar 

  65. Moelwyn-Hughes, E. A. (1961), Physical chemistry, Pergamon Press, London

    Google Scholar 

Download references

Acknowledgements

The author is thankful to Prof. M. K. Verma, Department of Physics, Indian Institute of Technology Kanpur, Kanpur–208016, India for fruitful discussions. This work was supported by research grant no. TAR/2018/000150 under Teachers Associateship for Research Excellence (TARE) scheme from the Science and Engineering Research Board (SERB), India. The author gratefully acknowledges financial support from SERB.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to P. K. Sahu.

Ethics declarations

Conflict of interests

The author reports no conflict of interest at this time. If a conflict of interest is identified after publication, a correction will be submitted.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendices

For interested readers, a detailed description of the mathematical model of the problem under consideration is presented in the Appendix 1, as well as the detailed solution procedure, is reported in the Appendix 2.

Mathematical model and problem description

The motion of piston is assumed to obey an exponential law, namely ( Sahu [30], Rao and Ramana [49])

$$\begin{aligned} r_{p} = A exp(\xi \, t),\quad \xi > 0 , \end{aligned}$$
(A.1)

where \(r_{p}\) is the radius of the piston, A and \(\xi \) are dimensional constants, and t is the time. ‘A’ represents the initial radius of the piston. It may be, physically, the radius of the stellar corona or the condensed explosive or the diaphragm containing a very high-pressure driver gas, at \(t = 0\). By sudden expansion of the stellar corona or the detonation products or the driver gas into the undisturbed ambient gas, a shock wave is produced in the ambient gas. The shocked gas is separated from the expanding surface which is a contact discontinuity. This contact surface acts as a ‘piston’ for the shock wave in the ambient medium (Rosenau and Frankenthal [55], Higashino [58], Liberman and Velikovich [59], Sahu [30]). Also, another justification for the exponential growth of the expanding piston refers to the internal energy source in simulating the ignition process. Ignition delay is a notation characterizing in experiments the time interval between mixture is placed under definite conditions and active energy release beginning accompanied by temperature as well as pressure growth. Ignition delay are often measured in shock tube experiments behind reflected shock waves. The dependence of ignition delay on temperature is always monotonous for constant pressure. Ignition delay decreases exponentially on linear increase of temperature (Smirnov and Nikitin [60], Smirnov et al. [61]). Thus, the energy release following chemical kinetics and turbulence increases after ignition, which brings to the formation of strong accelerating shock waves, which can finally result in detonation (Smirnov and Nikitin [60], Smirnov et al. [61]).

The law of piston motion (A.1) implies a boundary condition on the gas speed at the piston, which is required in the determination of the problem. Since we have assumed self-similarity, we may postulate that the shock propagation follows the exponential law

$$\begin{aligned} R = B exp(\xi \, t), \end{aligned}$$
(A.2)

where R is shock radius, and B is a dimensional constant. B depends on A and non-dimensional position of the piston [see equation (B.1)]. As is often the case in problems of this type, it is more convenient to solve for the piston motion in terms of the shock motion, rather than vice-versa. We shall therefore adopt this point of view forthwith, and consider B a known parameter of the problem, rather than A (Rosenau and Frankenthal [55]). The ‘piston’ is used to replicate blast waves as well as other similar phenomena in a model to simulate actual explosions and their effects, usually on a smaller scale. Thus ‘piston’ problem can be applied to quantify an estimate for the outcome from supernova explosions, sudden expansion of the stellar corona or detonation products, central part of starburst galaxies etc.

We have considered the medium to be a non–ideal gas, which is rotating about an axis of symmetry. We have taken r and t as independent space and time coordinates; uv,  and w as the radial, azimuthal and axial components of the fluid velocity \(\overrightarrow{X} \) in the cylindrical coordinates \((r,\theta ,z)\). Also, the relation between the angular velocity ‘\(C^{*}\)’ of the medium at radial distance r from the axis of symmetry and the azimuthal component of velocity is given by

$$\begin{aligned} v = C^{*} r, \end{aligned}$$
(A.3)

where ‘\(C^{*}\)’ is the angular velocity of the medium at radial distance r from the axis of symmetry. In this case the vorticity vector

$$\begin{aligned} \overrightarrow{\zeta } = \dfrac{1}{2} Curl \overrightarrow{X}, \end{aligned}$$

has the following components

$$\begin{aligned} \zeta _{r} = 0, \quad \zeta _{\theta } = -\dfrac{1}{2} \dfrac{\partial w}{\partial r}, \quad \zeta _{z} = \dfrac{1}{2 r} \dfrac{\partial }{\partial r} (rv). \end{aligned}$$
(A.4)

The total heat-flux Q,  which appear in the energy equation (5) may be decomposed as

$$\begin{aligned} Q = Q_{C} + Q_{R}, \end{aligned}$$
(A.5)

where \(Q_{C}\) is conduction heat flux and \( Q_{R} \) is radiation heat flux.

According to Fourier’s law of heat conduction

$$\begin{aligned} Q_{C} = - K \, \dfrac{\partial T}{\partial r}, \end{aligned}$$
(A.6)

where ‘K’ is the coefficient of the thermal conductivity of the gas and ‘T’ is the absolute temperature.

Assuming local thermodynamic equilibrium and using the radiative diffusion model for an optically thick grey gas Pomroning [62], the radiative heat flux \( Q_{R} \) may be obtain from the differential approximation of the radiation transport equation in the diffusion limit as

$$\begin{aligned} Q_{R} = - \dfrac{4}{3} \left( \dfrac{\sigma }{\alpha _{R}}\right) \dfrac{\partial T^{4}}{\partial r}, \end{aligned}$$
(A.7)

where \( \sigma \) is the Stefan–Boltzman constant and \( \alpha _{R} \) is the Rosseland mean absorption coefficient.

The radiation energy flux in local equilibrium (when the radiation at each point of a medium with a non-uniform temperature is close to equilibrium, then the medium is spoken of as being in a state of local thermodynamic equilibrium between the radiation and the fluid. The necessary condition for the existence of local equilibrium-small gradients in an extended, optically thick medium - serves simultaneously as a justification for the use of the diffusion approximation when considering radiative transfer) is proportional to the temperature gradient, and the radiative transfer is similar to heat conduction and is termed radiation heat conduction. The coefficient of thermal conductivity is equal to \(\dfrac{16 \, \sigma }{ 3 \, \alpha _{R}} T^{3}\) and is a function of temperature. In many cases it is possible to consider the radiation mean free path \( \left( = \dfrac{1}{\alpha _{R}} \right) \) proportional to some power of temperature (It is assumed that the density of the medium is constant) and for the power-law variation of the radiation mean free path the coefficient of the thermal conductivity is also proportional to a power of temperature (Zel’dovich and Raizer [31]). Thus, the above mentioned thermal conductivity ‘K’ and the absorption coefficient \( \alpha _{R} \) of the medium are assumed to vary with temperature only and these can be written in the form of power laws, namely (Ghoniem et al. [7], Nath and Sahu [9])

$$\begin{aligned} K = K_{0} \left( \dfrac{T}{T_{0}}\right) ^{\beta _{c}} \left( \dfrac{\rho }{\rho _{0}}\right) ^{\delta _{c}} \quad and \quad \alpha _{R} = \alpha _{R_{0}} \left( \dfrac{T}{T_{0}}\right) ^{\beta _{R}} \left( \dfrac{\rho }{\rho _{0}}\right) ^{\delta _{R}}, \end{aligned}$$
(A.8)

where the subscript ‘0’ denotes a reference state. The exponents in the above equations should satisfy the similarity requirements if a self similar solution is sought.

The electrical conductivity of the gas is assumed to be infinite. Therefore the diffusion term from the magnetic field equation is omitted, and the electrical resistivity is ignored. Also, the effect of viscosity on the flow of the gas is assumed to be negligible. The above system of equations (1) - (7) should be supplemented with an equation of state. In most of the cases, the propagation of shock waves arises in extreme conditions under which the assumption that the gas is ideal is not a sufficiently accurate description. The equation of state for a non-ideal gas is obtained by considering an expansion of the pressure p in powers of the density \( \rho \) (Anisimov and Spiner [63], Sahu [29, 30])

$$\begin{aligned} p = \Gamma \rho T \left[ 1 + \rho C_{1}(T) + \rho ^{2} C_{2}(T) + ... \right] , \end{aligned}$$
(A.9)

where \( \Gamma \) is the gas constant and \( C_{1}(T)\), \( C_{2}(T) \), ... are virial coefficients. The first term in the expansion corresponds to an ideal gas. The second term is obtained by taking into account the interaction between pairs of molecules, and subsequent terms must involve the interactions between the groups of three, four, etc. molecules. In the high-temperature range the. coefficients \( C_{1}(T)\) and \( C_{2}(T) \) tend to constant values equal to b and \( \frac{5}{8} \, b^{2} \), respectively. For gases \( b \rho<< 1 \), b being the internal volume of the molecules, and therefore it is sufficient to consider the equation of state in the form (Anisimov and Spiner [63], Sahu [29, 30] )

$$\begin{aligned} p&= \Gamma \rho T (1 + \rho b). \end{aligned}$$
(A.10)

In this equation the correction to pressure is missing due to neglect of second and higher powers of \( b \rho \), i.e. due to neglect of interactions between groups of three, four, etc. molecules of the gas. The internal energy \( U_{m} \) per unit mass of the non-ideal gas is given by (Anisimov and Spiner [63], Sahu [29, 30] )

$$\begin{aligned} U_{m} = \dfrac{p}{(\gamma - 1) \rho \, (1 + \rho b)}, \end{aligned}$$
(A.11)

where \( \gamma \) is the adiabatic index.

Real gas effects can be expressed in the fundamental equations according to Chandrasekhar [64], by two thermodynamic variables, namely by the sound velocity factor (the isotropic exponent) \( \Gamma ^{*} \) and a factor \(K^{*}\), which contains internal energy as follows,

$$\begin{aligned} \Gamma ^{*}&= \left( \dfrac{\partial \, ln \, p }{\partial \, ln \,\rho } \right) _{S} \quad and \quad K^{*} = - \dfrac{p}{\rho } \, \left( \dfrac{\partial \, U_{m} }{\partial \, ln \,\rho } \right) _{P}, \end{aligned}$$
(A.12)

where the subscripts ‘S’ and ‘P’ refers to the process of constant entropy and constant pressure.

Using the first law of thermodynamics and the Equations (A.10) and (A.11), we obtain

$$\begin{aligned} \Gamma ^{*}&= \dfrac{\gamma \, (1 + 2 b \rho )}{1 + b \rho } \quad and \quad K^{*} = \dfrac{1}{\gamma - 1}, \end{aligned}$$
(A.13)

neglecting the second and higher powers of b. This shows that the isentropic exponent \( \Gamma ^{*} \) is non-constant in the shocked gas, but the factor \(K^{*}\) is constant for the simplified equation of state of the non-ideal gas in the form equation (A.10).

The isentropic velocity of sound \( a_{non} \) is given by

$$\begin{aligned} a_{non}^{2}&= \Gamma ^{*} \, \dfrac{p}{\rho }. \end{aligned}$$
(A.14)

Ahead of the shock, the components of the vorticity vector, therefore vary as

$$\begin{aligned} \zeta _{r_{a}}&= 0, \end{aligned}$$
(A.15)
$$\begin{aligned} \zeta _{\theta _{a}}&= - \dfrac{w^{*} \alpha }{2 \, \xi R} exp(\alpha \, t), \end{aligned}$$
(A.16)
$$\begin{aligned} \zeta _{z_{a}}&= \dfrac{v^{*} (\delta + \xi )}{2 \, \xi R} exp(\delta \, t). \end{aligned}$$
(A.17)

The initial angular velocity of the medium at radial distance R is given by, from (A.3),

$$\begin{aligned} C^{*}_{a} = \dfrac{v_{a}}{R}. \end{aligned}$$
(A.18)

From equations (10) and (A.18), we find that the initial angular velocity vary as

$$\begin{aligned} C^{*}_{a} = \dfrac{v^{*}}{R} exp(\delta \, t) = \dfrac{v^{*}}{B} exp \left\{ (\delta - \xi )\, t\right\} . \end{aligned}$$
(A.19)

In addition to equilibrium sound speed of non-ideal gas (A.14), the isothermal speed of sound may also play a role, where thermal radiation is taken into account. The isothermal sound speed in the non-ideal gas is

$$\begin{aligned} a_{iso} = \left( \dfrac{\partial p}{\partial \rho } \right) _{T}^{\frac{1}{2}} = \left[ \dfrac{ (1 + 2 b \, \rho ) p}{\rho \, (1 + b \rho )} \right] ^{\frac{1}{2}}, \end{aligned}$$
(A.20)

where the subscript ‘T’ refers to the process of constant temperature.

The adiabatic compressibility of the non–ideal gas may be calculated as (Moelwyn-Hughes [65])

$$\begin{aligned} C_{adi} = \dfrac{1}{\rho } \left( \dfrac{\partial \rho }{\partial p} \right) _{S} = \dfrac{1}{\rho \, a_{non}^{2}}= \left[ \dfrac{ (1 + b \rho ) }{\gamma p (1 + 2 b \rho )} \right] . \end{aligned}$$
(A.21)

Following Levin and Skopina [45], Nath and Sahu [9]; we obtained the jump conditions for the components of vorticity vector across the shock front as

$$\begin{aligned} \zeta _{\theta _{s}} = \dfrac{\zeta _{\theta _{a}}}{\beta }, \quad \zeta _{z_{s}} = \dfrac{\zeta _{z_{a}}}{\beta }. \end{aligned}$$
(A.22)

Detailed solution procedure

Equations (A.1), (A.2) and (21) yields a relation between B and A in the form

$$\begin{aligned} B = \dfrac{A}{y_{p}}. \end{aligned}$$
(B.1)

Using the similarity transformations (21), the system of governing equations (1)–(7) can be transformed to the following system of ordinary differential equations:

$$\begin{aligned}&(U - y) \, \dfrac{dD}{dy} + D \, \dfrac{dU}{dy} + \dfrac{D \, U}{y} = 0, \end{aligned}$$
(B.2)
$$\begin{aligned}&(U - y) \, D \, \dfrac{dU}{dy} + \dfrac{dP}{dy} + H \, \dfrac{dH}{dy} + \dfrac{i \, H^{2}}{y} + U \,D - \dfrac{D \, V^{2} }{y} + \dfrac{ G^{*} \, N \, D}{y} = 0 , \end{aligned}$$
(B.3)
$$\begin{aligned}&(U - y) \, \dfrac{dV}{dy} + \dfrac{(U + y) \, V }{y} = 0 , \end{aligned}$$
(B.4)
$$\begin{aligned}&(U - y) \, \dfrac{dW}{dy} + W = 0 , \end{aligned}$$
(B.5)
$$\begin{aligned}&\dfrac{2\, P }{(\gamma - 1)\, D(1 + {\overline{b}} \,D)} + \dfrac{(U - y)}{(\gamma - 1)\, D^{2} \, (1 + {\overline{b}} \,D)^{2}} \left[ D \, (1 + {\overline{b}} \,D) \dfrac{dP}{dy} - P \, (1 + 2 {\overline{b}} \,D) \dfrac{dD}{dy} \right] \nonumber \\&\quad - \dfrac{P \, (U - y) }{D^{2}} \dfrac{dD}{dy} + \dfrac{F}{D \, y} + \dfrac{1}{ D} \dfrac{dF}{dy} = 0 \end{aligned}$$
(B.6)
$$\begin{aligned}&(U - y) \, \dfrac{dH}{dy} + H \, \dfrac{dU}{dy} + H + \dfrac{H \, U (1 - i) }{y} = 0, \end{aligned}$$
(B.7)
$$\begin{aligned}&\dfrac{dN}{dy} - 2 \, D \, y = 0, \end{aligned}$$
(B.8)

where \( G^{*} = \dfrac{G \pi \rho _{a}}{\xi ^{2}}\) is the gravitational parameter.

Using equations (A.6), (A.7) and (A.8) in equation (A.5), we get

$$\begin{aligned} Q = - \dfrac{K_{0}}{ T_{0}^{\beta _{c}} \rho _{0}^{\delta _{c}}} T^{\beta _{c}} \rho ^{\delta _{c}} \dfrac{\partial T}{\partial r} - \dfrac{16\, \sigma T_{0}^{\beta _{R}} \, \rho _{0}^{\delta _{R}}}{3\, \alpha _{R_{0}}} T^{ 3 - \beta _{R}} \rho ^{- \delta _{R}} \dfrac{\partial T}{\partial r}. \end{aligned}$$
(B.9)

Using equations (A.9) and (21) in equation (B.9), we get

$$\begin{aligned} F= & {} \Bigl [ \dfrac{- K_{0} \, \xi }{T_{0}^{\beta _{c}} \, \rho _{0}^{\delta _{c}} \, \Gamma ^{\beta _{c} + 1}} \, \dfrac{{\dot{R}}^{2 \beta _{c} - 2} P^{\beta _{c}} \, \rho _{a}^{\delta _{c} - 1 } \, D^{\delta _{c} - \beta _{c}}}{(1 + {\overline{b}} \, D)^{\beta _{c}} \,} \nonumber \\&\quad - \dfrac{16\, \sigma T_{0}^{\beta _{R}} \, \rho _{0}^{\delta _{R}} \, \xi }{3\, \alpha _{R_{0}} \, \Gamma ^{4 - \beta _{R}}} \, \dfrac{{\dot{R}}^{4 - 2 \beta _{R}} \, P^{3- \beta _{R}} \, \rho _{a}^{- 1 -\delta _{R} } \, D^{- 3 + \beta _{R} - \delta _{R} }}{(1 +{\overline{b}} \, D)^{3- \beta _{R}} \,} \Bigr ] \dfrac{d}{d y} \left( \dfrac{P }{D \, (1 + {\overline{b}} \, D)}\right) .\nonumber \\ \end{aligned}$$
(B.10)

Equation (B.10) shows that the similarity solution of the present problem exists only when

$$\begin{aligned} \beta _{c} = 1 \quad and \quad \beta _{R} = 2 . \end{aligned}$$
(B.11)

Therefore equation (B.10) becomes

$$\begin{aligned} F = - X \left[ \dfrac{1}{D \, (1 + {\overline{b}} \, D)} \dfrac{dP}{d y} - \dfrac{P \, (1 + 2 \, {\overline{b}} \, D)}{D^{2} \, (1 +{\overline{b}} \, D)^{2}} \dfrac{dD}{d y} \right] , \end{aligned}$$
(B.12)

where \( X = \left[ \Gamma _{C} \, D^{\delta _{c} - 1} + \Gamma _{R} \, D^{ - 1 - \delta _{R}}\right] \dfrac{ P}{(1 + {\overline{b}} \, D)} \), \( \Gamma _{C} \) and \( \Gamma _{R} \) are the conductive and radiative non-dimensional heat transfer parameters, respectively. The parameters \( \Gamma _{C} \) and \( \Gamma _{R} \) depend on the thermal conductivity K and the mean free path of radiation \( \dfrac{1}{\alpha _{R}} \) respectively and they are given by

$$\begin{aligned} \Gamma _{C} = \dfrac{ K_{0} \, \xi }{T_{0} \, \rho _{0}^{\delta _{c}} \, \Gamma ^{2}} \, \rho _{a}^{\delta _{c} - 1} \quad and\quad \Gamma _{R} = \dfrac{16\, \sigma T_{0}^{2} \, \rho _{0}^{\delta _{R}} \, \xi }{3\, \alpha _{R_{0}} \, \Gamma ^{2}} \,\rho _{a}^{- 1 - \delta _{R} }. \end{aligned}$$

Solving the set of differential equations (B.2)–(B.8) and equation (B.12), we have equations (23)–(30).

Applying the similarity transformations (21) on equation (A.4), we obtained the non-dimensional components of the vorticity vector \(l_{r} = \dfrac{\zeta _{r}}{{\dot{R}}/R},\)   \(l_{\theta } \dfrac{\zeta _{\theta }}{{\dot{R}}/R}\),   \(l_{z} = \dfrac{\zeta _{z}}{{\dot{R}}/R}\) in the flow-filed behind the shock as

$$\begin{aligned} l_{r}&= 0, \end{aligned}$$
(B.13)
$$\begin{aligned} l_{\theta }&= - \dfrac{W}{2 (y- U)}, \end{aligned}$$
(B.14)
$$\begin{aligned} l_{z}&= \dfrac{V}{(y- U)}. \end{aligned}$$
(B.15)

By using equation (21) the expression for reduce isothermal speed of sound (A.20) becomes

$$\begin{aligned} \dfrac{a_{iso}}{{\dot{R}}} = \left[ \dfrac{ (1 + 2\, {\overline{b}} \, D) P}{ D \, (1 + {\overline{b}} \, D)} \right] ^{\frac{1}{2}}. \end{aligned}$$
(B.16)

By using equation (21) in equation (A.21), we obtain the expression for the adiabatic compressibility \( C_{adi} \) as

$$\begin{aligned} \left( C_{adi}\right) \, p_{a} = \dfrac{ (1 + {\overline{b}} \,D)}{P \, \gamma ^{2} \, M^{2} \, (1 + 2 {\overline{b}} \, D)}. \end{aligned}$$
(B.17)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sahu, P.K. Magnetogasdynamic exponential shock wave in a self-gravitating, rotational axisymmetric non-ideal gas under the influence of heat-conduction and radiation heat-flux. Ricerche mat 73, 113–149 (2024). https://doi.org/10.1007/s11587-021-00563-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11587-021-00563-7

Keywords

Navigation