You are on page 1of 17

Physics of Fluids ARTICLE scitation.

org/journal/phf

Piston driven converging shock waves in nonideal


magnetogasdynamics of variable density
Cite as: Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883
Submitted: 28 July 2021 . Accepted: 30 October 2021 .
Published Online: 22 November 2021

Antim Chauhan,1,a) Rajan Arora,1,b) and Amit Tomar2,c)

AFFILIATIONS
1
Department of Applied Science and Engineering, Indian Institute of Technology Roorkee, Roorkee 247667, India
2
Department of Mathematics (AIAS), Amity University, Noida, Uttar Pradesh 201303, India

a)
Electronic-mail: achauhan@as.iitr.ac.in
b)
Electronic-mail: rajana100@gmail.com. Tel.: þ91 1332286563
c)
Author to whom correspondence should be addressed: atomar@amity.edu

ABSTRACT
In this article, we analyze an imploding strong shock wave problem collapsing at the axis of cylindrical piston filled with a nonideal gas of
nonuniform density that is decreasing toward the axis of symmetry according to a power law. The magnetic field is considered to be present
in the axial direction, and the electrical resistance is assumed to be zero. The perturbation series technique applied to the system of
hyperbolic partial differential equations governing the one-dimensional adiabatic cylindrically symmetric flow of a nonideal gas in the pres-
ence of an axial magnetic field provides us a global solution and also recovers Guderley’s local solution, which holds only in the neighbor-
hood of shock collapse. All possible similarity exponents and corresponding amplitudes are found by expanding all the flow variables and
shock location in powers of time. A comparison has been made between the computed values of similarity exponents with published results
in the literature, and the results are in good agreement. The flow parameters and shock position have been analyzed graphically.
Published under an exclusive license by AIP Publishing. https://doi.org/10.1063/5.0064883

I. INTRODUCTION by Guderley.2 The important aspects and results of self-similar solu-


The shock wave phenomenon is interesting from both mathe- tions to the flow problems of a strong converging shock wave was dis-
matical and physical points of view and has been a very important cussed by Lazarus and Richtmyer.3 Hafner4 has formulated a strong
research topic for many years. The topic of converging shock waves is converging shock wave problem for a zero-pressure gas using
quite useful in inertial confinement fusion. The problem of imploding Lagrangian coordinates and solved it by using the power series
shock waves is a field of continuing research interest as it generates method. The study of shock wave phenomena can also be seen in the
high temperature and pressure at the center/axis of convergence. following work by Sari et al.,5 Ponchaut et al.,6 Sachdev et al.,7 Boyd
Moreover, these waves are important from a fundamental gas dynami- et al.,8 Chisnell,9 Whitham,10 Ramsey et al.,11 Ram and Pandey,12
cal point of view. High temperatures and pressures at the center of Singh et al.,13 Taniguchi and Ruggeri,14 Zhai et al.,15 Modelevsky and
convergence of spherical or cylindrical shock waves give rise to various Sari,44 and Shah and Singh.16
interesting applications such as detonation and fusion initiation. Many In a nonideal gas, shock waves exhibit a more complex behavior
young mathematicians and researchers around the world are working than the ideal gas model predicts. Thus, the study of shock waves in a
together to investigate all the possibilities to make the best utilization nonideal gas becomes significant and must be involved for theoretical
of shock waves for various industrial applications. In IISc Bengaluru, study.17 The van der Waals equation of state exemplifies the nonideal
remarkable work has been performed in this direction as researchers gas in particular. The study of shock wave-related events through a
have developed many applications of shock waves such as shock wave- nonideal gas is found to be relevant in a variety of practical applica-
assisted gene transfer, oil extraction, drug delivery, and preservative tions such as chemical processes, nuclear reactions, aerospace engi-
injection into wood slats.1 neering, and mechanical engineering systems. Wu and Roberts18
The theoretical study of cylindrically/spherically symmetric shock found the self-similar solutions for a shock wave propagating in a van
waves converging near the axis/center of symmetry was first published der Waals gas and also investigated its stability. A complete

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-1


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

classification of shock waves in a van der Waals fluid was given by


Zhao et al.19 Recently, Ambika et al.20 studied the effects of nonlinear
attenuation, convection, and geometrical spreading in a van der Waals
gas. Concerning the study of converging shock waves in a nonideal
gas, we mention the work by different authors, namely, Arora and
Sharma,21 Tomar et al.,22 Arora et al.,23 Chauhan and Arora,24 and
Sahu.25
Magnetogasdynamics play a significant role in plasmas and con-
ductive fluid flows experienced in nature. In the dynamics of interstel-
lar media, the presence of a strong magnetic field plays an important
role. Lock and Mestel26 and Bira et al.27 found self-similar solutions
for ideal and nonideal gases in the presence of a magnetic field, respec-
tively. A study on shock waves in nonideal magnetogasdynamics can
be found in the works of Ram et al.,29 Singh et al.,30 Shyam et al.,31
Singh et al.,32,33 Sharma,28 Singh and Arora,42 etc.
The situation in which strong converging waves can propagate
through a nonuniform medium of decreasing density and finally
achieve their destination, where the density is almost zero, is of great
interest in astrophysics because it is highly relevant to the problem of
the origin of cosmic rays.4,34,35 The propagation of shock waves
through a medium of variable density has been treated by Rogers,36 FIG. 1. The cylindrical piston is initially of radius R0. At time t ¼ 0, the piston begins
Hafner,4 and Madhumita and Sharma.37 Further, their results are to contract at high constant velocity V, driving ahead of it a cylindrical shock of
radius R(t) collapsing at the axis of implosion o, where R(t) is to be determined.
more applicable to shocks formed in the interior of stars. Roger36 stud-
ied the adiabatic motion of a perfect gas behind an infinitely strong
shock wave and obtained analytical solution to the problem with the
magnetic field together with zero electrical resistance is assumed. The
similarity method, whereas, Hafner4 presented the Lagrangian formu-
gaseous medium is taken to be inhomogeneous by means of variable
lation of the self-similar convergent shock problem for a quiescent per-
density. The density of the medium initially varies according to a
fect gas of zero pressure. Using an alternative approach introduced by
power law. The piston begins contracting with very high velocity V
Lee,38 Van Dyke and Guttmann,39 an implosion problem in an ideal
(this velocity V is greater than the acoustic speed of the medium) at
gas has been solved globally and all possible real similarity exponents
time t ¼ 0 and driving a cylindrically symmetric strong shock wave
and their corresponding amplitudes have been determined. The prob-
through a quiescent (v ¼ 0) nonideal gas of variable density
lem of strong converging shock waves in an ideal gas of variable den-
[q ¼ q0 ¼ qc ðr=R0 Þd ] in which the initial pressure (p ¼ p0 ) and mag-
sity and in a nonideal gas of constant density has been analyzed by
netic pressure (h ¼ h0 ) are assumed to be constant. The aim of this
Madhumita and Sharma37 and Arora and Sharma,21 respectively, by
study is to determine the position R(t) of a converging shock wave and
considering the problem of a strong converging cylindrical/spherical
the functional form of flow variables throughout the flow field between
shock collapsing at the axis/center driven by a piston, where the cylin-
the piston and the shock wave.
drical/spherical piston is assumed to be filled with the gas and the pis- The system of partial differential equations describing the one-
ton begins contracting at a very high constant speed, greater than the dimensional motion of a cylindrically symmetric, unsteady, adiabatic,
acoustic speed of the medium. Furthermore, Chauhan et al.40 extended nonideal gas under the magnetic field effect can be written as
the work to nonideal gases of constant density under the effect of mag- follows:40
netic fields. The problem of converging strong shock waves in a van
der Waals gas by taking the Mie–Gruneisen type equation of state was @q @q @v qv
þv þ q þ ¼ 0; (1)
analyzed by Tomar et al.22 More recently, Ramsay and Baty41 worked @t @r @r r
on the driving one-dimensional planar piston problem filled with stiff  
@v @v 1 @p @h
gas. Giron et al.43 obtained the solution for the imploding shock prob- þv þ þ ¼ 0; (2)
@t @r q @r @r
lem in a medium with varying density. In the present work, we deal  
with the generalized problem of converging shock waves propagating @p @p @q @q
þ v  a2 þv ¼ 0; (3)
through a nonuniform gaseous medium of variable density filled in a @t @r @t @r
 
cylindrical piston under the presence of an axial magnetic field that @h @h @q @q
begins contracting with a velocity greater than the acoustic speed of þ v  c2 þv ¼ 0; (4)
@t @r @t @r
the medium and generates a strong shock inside the piston, and deter-
mine the global solution for the implosion problem. where v; q; p, h, t, and r denote the outward radial velocity, density, pres-
sure, magnetic pressure, time, and spatial coordinate, respectively. The
II. THE PISTON PROBLEM AND SOLUTIONS IN TAYLOR magnetic pressure h, the speed of sound a, and the Alfven speed c are
SERIES EXPANSION defined as h ¼ lH 2 =2; a ¼ ðcp=qð1  bqÞÞ1=2 , and c ¼ ð2h=qÞ1=2 ,
A cylindrical piston (see Fig. 1) of initial radius R0 filled with a respectively, where l is the magnetic permeability and H is the magnetic
quiescent nonideal gas is considered. The presence of an axial field strength.

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-2


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

The gas follows an equation of state of the form (Roberts and 2ð1  bqc ð1  X Þd Þ
18 ¼
u _ ;
X
Wu ) cþ1
qRT ¼ pð1  bqÞ; (5) cþ1  Þd ;

q q ð1  X
ðc  1 þ 2bqc ð1  X Þd Þ c
where R is the gas constant and T is the absolute temperature. !
Let r ¼ RðtÞ be the shock front at which the shock jump condi-  Þd Þ 2C0 ððc  1Þbqc ð1  X
2ð1  bqc ð1  X  Þd  cÞ
p ¼ þ (13)
tions are given as follows: cþ1 ðc  1 þ 2bqc ð1  X Þd Þ2
2ð1  ~bÞ _ _ ;
 Þd X
 qc ð1  X
2
v¼ R;
cþ1 !2
 C0 ðc þ 1Þ 2
_ ;
 Þd X
cþ1 h¼ qc ð1  X
q¼ q0 ;  Þd Þ
2 ðc  1 þ 2bqc ð1  X
c  1 þ 2~b  
! 
2ð1  ~bÞ 2C0 ððc  1Þ~b  cÞ 2
(6) at y ¼ k Xt  k ; and
p¼ þ 2
q0 R_ ;
cþ1 ~
ðc  1 þ 2bÞ ¼1
u at y ¼ kðk  1Þ: (14)
!2
C0 ðc þ 1Þ 2 Furthermore, for convenience, all bar signs will be dropped from
h¼ q0 R_ ;
2 ðc  1 þ 2~bÞ Eqs. (8)–(14).
2 We consider the solutions to be analytic in time, then we can
_
where RðtÞ denotes the speed of the shock while C0 ¼ 2h0 =q0 R_ is write the shock position X and all the flow variables in powers of time
the shock cowling number representing the effect of magnetic field in as follows:
the present problem and ~b ¼ q0 b.
When there is no flow through the piston, we have the following X
1
X¼ Xn t n ; (15)
condition: n¼1

v ¼ V at r ¼ R0  Vt: (7) X
1 X
1
u¼ Un ðyÞt n1 ; q¼ Rn ðyÞt n1 ;
Here, v denotes the gas velocity relative to fixed coordinates in n¼1 n¼1
(16)
which the gas is at rest in the piston with radius R0. For convenience, X
1 X
1
n1 n1
we take the distance x ¼ R0  r and the corresponding velocity p¼ Pn ðyÞt ; h¼ Hn ðyÞt :
n¼1 n¼1
u ¼ v directed inward.
We introduce the dimensionless variables as r ! r=R0 ; v ! v=V; By substituting Eq. (16) into Eq. (13) and using Eqs. (14) and
 ! q=q0 ; p ! p=ðq0 V 2 Þ; h
q  ! h=h0 ; b ! b=q ; t ! tV=R0 , and (15), we determined the coefficients for the first-order approximations
0
_R ! R=V.
 _ In the dimensionless form, the above system of partial differ- U1 ; R1 ; P1 , and H1 by equating the coefficients of like powers of t as
ential equations (1)–(4) together with x þr ¼ 1 and v ¼ 
u becomes follows:

@q @q @u q


u ðc þ 1Þ
þu þq   ¼ 0; (8) U1 ¼ 1; R1 ¼ ;
@t @x @x 1  x ðc  1 þ 2bÞ
!
  2ð1  bÞ 2C0 ððc  1Þb  cÞ 2
u
@ u 1 @ p @ h
@  P1 ¼ þ X1 ;
þu  þ þ ¼ 0; (9) ðc þ 1Þ ðc  1 þ 2bÞ2
@t @x q  @x @x (17)
 2
  C0 cþ1 ðc þ 1Þ
@ p @ p cp @q q
@ H1 ¼ X12 ; X1 ¼ ;
þu
  þ 
u ¼ 0; (10) 2 c  1 þ 2b 2ð1  bÞ
@t @x q  ð1  bq Þ @t @x
2 þ bðc  1Þ
  k¼ :

@h 
@h q
@ q
@ 2ð1  bÞ
þu
 
 2h þu
 ¼ 0: (11)
@t @x @t @x Moreover, the nth (n  2) order approximations in the Taylor
By considering a dimensionless formulation, we choose an inde- series expansion are assumed to be of the following form:
pendent variable as X
n X
n
  Un ðyÞ ¼ Unl yl1 ; Rn ðyÞ ¼ Rnl yl1 ;
1  r
y ¼ k k ; (12) l¼1 l¼1
(18)
t
X
n X
n

2 Pn ðyÞ ¼ Pnl yl1 ; Hn ðyÞ ¼ Hnl yl1 :


where k is an unknown parameter and k ¼ ðc1Þ . l¼1 l¼1
Let the shock front be at a distance X, measured inward from the
initial piston location, i.e., X ¼ R0  R. Thus, following are the We substitute the Eqs. (15) and (16) together with Eq. (18) into
boundary conditions at the shock and piston, respectively: the transformed system given by Eqs. (8)–(11), to obtain the following:

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-3


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

" # !1 n
X
1 X
n X
n X
1 X
n XX
l2 l1 n2 l1 n1
½k  ðy þ kkÞt  ðy þ kkÞ ðl  1ÞRnl y þ ðn  1Þ Rnl y t þ k½k  ðy þ kkÞt  1 þ Unl y t ðl  1ÞRnl yl2 t n2
n¼2 l¼1 l¼1 n¼2 l¼1 n¼2 l¼1
!1 n ! !
X
1 X
n XX X
1 X
n X
1 X
n
l1 n1 l2 n2 l1 n1 l1 n1
þk½k  ðy þ kkÞt  R1 þ Rnl y t ðl  1ÞUnl y t  k R1 þ Rnl y t 1þ Unl y t ¼ 0; (19)
n¼2 l¼1 n¼2 l¼1 n¼2 l¼1 n¼2 l¼1
!1" #
X
1 X
n X X
n X
n
R1 þ Rnl yl1 t n1 ðy þ kkÞ ðl  1ÞUnl yl2 þ ðn  1Þ Unl yl1 t n2
n¼2 l¼1 n¼2 l¼1 l¼1
! !1 n
X
1 X
n X
1 X
n XX
þk R1 þ Rnl yl1 t n1 1þ l1 n1
Unl y t ðl  1ÞUnl yl2 t n2
n¼2 l¼1 n¼2 l¼1 n¼2 l¼1
X
1 X
n X
1 X
n
þk ðl  1ÞPnl yl2 t n2 þ k ðl  1ÞHnl yl2 t n2 ¼ 0; (20)
n¼2 l¼1 n¼2 l¼1
!! ! " !
X
1 X
n X
1 X
n X
1 X
n X
n
1  b R1 þ Rnl yl1 t n1 R1 þ Rnl yl1 t n1  l2
ðy þ kkÞ ðl  1ÞPnl y þ ðn  1Þ Pnl y l1
t n2
n¼2 l¼1 n¼2 l¼1 n¼2 l¼1 l¼1
! !
1 X
X n X
1 X
n X
1 X
n
þk 1 þ Unl yl1 t n1 ðl  1ÞPnl yl2 t n2   c P1 þ Pnl yl1 t n1
n¼2 l¼1 n¼2 l¼1 n¼2 l¼1
" ! !1 n #
X
1 X
n X
n X
1 X
n XX
 ðy þ kkÞ ðl  1ÞRnl yl2 þ ðn  1Þ Rnl yl1 t n2 þk 1 þ Unl yl1 t n1 ðl  1ÞRnl yl2 t n2 ¼ 0; (21)
n¼2 l¼1 l¼1 n¼2 l¼1 n¼2 l¼1
! " !
X
1 X
n X
1 X
n X
n
R1 þ Rnl yl1 t n1  l2
ðy þ kkÞ ðl  1ÞHnl y þ ðn  1Þ Hnl y l1
t n2
n¼2 l¼1 n¼2 l¼1 l¼1
! # !
1 X
X n X
1 X
n X
1 X
n
þk 1 þ Unl yl1 t n1 l2 n2
ðl  1ÞHnl y t  2 H1 þ l1 n1
Hnl y t
n¼2 l¼1 n¼2 l¼1 n¼2 l¼1
" ! !1 n #
X
1 X
n X
n X
1 X
n XX
l2 l1 n2 l1 n1 l2 n2
 ðy þ kkÞ ðl  1ÞRnl y þ ðn  1Þ Rnl y t þk 1 þ Unl y t ðl  1ÞRnl y t ¼ 0: (22)
n¼2 l¼1 l¼1 n¼2 l¼1 n¼2 l¼1

In a similar manner, Eq. (13) can also be transformed as follows:


! 0 !d 1
X1 X n
2 @ X1 X1
1þ l1 n1
Unl ys t  1  bqc 1  Xn t n A nXn t n1 ¼ 0; (23)
n¼2 l¼1
c þ 1 n¼1 n¼1
!   !1 !d
X
1 X
n
cþ1 2bqc X1 X1
n d
R1 þ Rnl ysl1 t n1  1þ ð1  Xn t Þ qc 1  Xn t n
¼ 0; (24)
n¼2 l¼1
c1 c1 n¼1 n¼1
! 2 ! !2 3
X
1 X
n
2 X1
2C0 2bqc X1
P1 þ Pnl ysl1 t n1  qc 4 1  bqc ð1  Xn t n Þd þ 1þ ð1  Xn t n Þd 5
n¼2 l¼1
cþ1 n¼1 ðc  1Þ2 ðc  1Þ n¼1
!d !2
X
1 X
1
 1 Xn t n nXn t n1 ¼ 0; (25)
n¼1 n¼1
!   !2 !d !2
X
1 X
n
C 0 qc c þ 1 2 2bqc X1 X
1 X
1
n d
H1 þ Hnl ysl1 t n1  1þ ð1  Xn t Þ  1 Xn t n nXn t n1 ¼ 0: (26)
n¼2 l¼1
2 c1 c1 n¼1 n¼1 n¼1

From the above, it is clear that Eqs. (23)–(26) are valid only at y ¼ ys , where the y-position of the cylindrical shock wave can easily be
obtained from Eqs. (12) and (15) in the following form:

1 X 1
ys ¼ yðr ¼ RÞ ) ys ¼ 1 þ Xn t n1 : (27)
X1  k n¼2

In order to solve the problem up a second-order approximation, we set n ¼ 2 in the above Eqs. (19)–(26), which yields the following:

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-4


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

2
½k  ðy þ kkÞt ½ðkkR22 Þ þ R21 þ kð1 þ U21 t þ U22 ytÞR22 U21 þ U22 ¼ ð2ð1  bÞX2 þ bdX12 Þ; (42)
cþ1
þ kðR1 þ R21 t þ R22 ytÞU22   k½R1 þ R21 t þ R22 yt  !
2ð1  bÞ 2C0 ððc  1Þb  cÞ
 ½1 þ U21 t þ U22 yt  ¼ 0; (28) P21 þ P22 ¼ þ ð4X1 X2  dX13 Þ
cþ1 ðc  1 þ 2bÞ2
½R1 þ R21 t þ R22 yt ½kkU22 þ U21  þ kU22 ½R1 þ R21 t þ R22 yt  !!
 ½1 þ U21 t þ U22 yt  þ k½P22 þ H22  ¼ 0; (29) 2bd 4dbððc  1Þb  cÞ bðc  1Þd
þ þ 2C0 3  X13 ;
cþ1 ðc  1 þ 2bÞ ðc  1 þ 2bÞ2
½R1 þ R21 t þ R22 yt ½1  bðR1 þ R21 t þ R22 ytÞ
 ½ðkkP22 þ P21 Þ þ kP22 ð1 þ U21 t þ U22 ytÞ (43)
 2 !
3
c½P1 þ P21 t þ P22 yt ½ðkkR22 þ R21 Þ C0 cþ1 3 4dX 1 b
H21 þ H22 ¼ 4X1 X2  dX1 þ :
þ kR22 ð1 þ U21 t þ U22 ytÞ ¼ 0; (30) 2 c  1 þ 2b ðc  1 þ 2bÞ

½R1 þ R21 t þ R22 yt ½ðkkH22 þ H21 Þ þ kð1 þ U21 t þ U22 ytÞH22  (44)
2½H1 þ H21 t þ H22 yt ½ðkkR22 þ R21 Þ On the piston, the boundary condition (14) becomes
þ kð1 þ U21 t þ U22 ytÞR22  ¼ 0; (31) 2ðk  1Þ
Un ðyÞ ¼ 0; at y¼ ; for n ¼ 2; 3; 4;   : (45)
2 d c1
ð1 þ U21 t þ U22 ys tÞ  ð1  bqc ð1  X1 t  X2 t 2 Þ Þ
cþ1 On solving Eqs. (37)–(40) together with Eqs. (41)–(45), we deter-
 ðX1 þ 2X2 tÞ ¼ 0; (32) mined R21 ; R22 ; U21; U22 ; P21 ; P22 ; H21 , and H22; therefore, the coeffi-
  cients for the second-order approximations U2 ; R2, P2, and H2 are
cþ1
ðR1 þ R21 t þ R22 ys tÞ  qc calculated as follows:
c1
 1  
2bqc ð2X2 ð1  bÞ þ bdX22 Þ ðc  1Þ
 1þ ð1  X1 t  X2 t 2 Þd ð1  X1 t  X2 t 2 Þd ¼ 0; (33) U2 ¼ 2b þ y ; (46)
c1 ðc  1 þ 2bÞ X1
  
2 ðc  1ÞR21 ðy  1Þ
ðP1 þ P21 t þ P22 ys tÞ  qc 1  bqc ð1  X1 t  X2 t 2 Þd R2 ¼ ð2bdX13 þ 2bdX12  ðc  1 þ 2bÞX1
cþ1 2ðc  1 þ 2bÞX12
 2
2C0 2bqc 2 d
dðc  1ÞR1 X1 y
þ 1 þ ð1  X 1 t  X 2 t Þ þ 4ð1  bÞX2 Þ  ; (47)
ðc  1Þ2 c1 2ðc  1 þ 2bÞ
 i  
yðc  1Þ
 ðc  1Þbqc ð1  X1 t  X2 t 2 Þd  c P2 ¼ b þ R1 A þ ð1  yÞcP1 B; (48)
2X1
!
 ð1  X1 t  X2 t 2 Þd ðX1 þ 2X2 tÞ2 ¼ 0; (34)
  C0 R21 4dbX12
C0 qc c þ 1 2 H2 ¼ 4X1 X2  dX13 þ
ðH1 þ H21 t þ H22 ys tÞ  2 ðc  1 þ 2bÞ
2 c1  
 2 ðc  1ÞR1 A
2bqc þ ð1  yÞ  cP1 B ; (49)
 1þ ð1  X1 t  X2 t 2 Þd ð1  X1 t  X2 t 2 Þd 2X1
c1
C
 ðX1 þ 2X2 tÞ2 ¼ 0; (35) X2 ¼ ;
D
(50)
where ys is obtained from Eq. (27) in the following form: where
X2 t P1 1 4ðc þ bðc  1ÞÞ
ys ¼ 1 þ : (36) A¼ ð4X2  dX12 Þ þ 2bdX13 þ C0
X1  k X1 cþ1 ðc  1 þ 2bÞ3
Coefficients of t0 in Eqs. (28)–(31) are, respectively, !!
ðc  1Þ
kR22 ð1  kÞ þ R21 þ R1 ðkU22  1Þ ¼ 0; (37)  ;
ðc  1 þ 2bÞ2
kR1 U22 ð1  kÞ þ R1 U21 þ kðP22 þ H22 Þ ¼ 0; (38)
bR1 dX1 2ð2X2 ð1  bÞ þ bdX12 Þ 2bdX12
B¼1þ   ;
ð1  bR1 Þ½kP22 ð1  kÞ þ P21   cP1 ½kR22 ð1  kÞ þ R21  ¼ 0; (39) ð1  bÞ X1 ðc  1 þ 2bÞ ðc  1 þ 2bÞ
 
R1 ½kð1  kÞH22 þ H21   2H1 ½kð1  kÞR22 þ R21  ¼ 0: (40) 2bdX1 ð1 þ X1 Þ ðc  1Þð1  bÞdX1 bR1 dX1
C ¼cP1 X1 1  þ þ
Equations (37)–(40) do not form a closed system for c  1 þ 2b cðc  1 þ 2bÞ ð1  bÞ
!
R21 ; R22 ; U22 ; P21 ; P22 ; H21 , and H22 cannot be solved independently. R1 dX12 X1 bdR1 ð1 þ X1 Þ ðc  1ÞdX1 R1
This system is closed with t1 powers of Eqs. (32)–(35) as follows: þ2H1 X1 þ  
2 ðc  1 þ 2bÞ ðc  1 þ 2bÞ 2ðc  1 þ 2bÞ
ðc þ 1Þðc  1Þ ðc  1ÞR31 C0 dX12 ðX1 ðc  1 þ 2bÞ  4bÞ
R21 þ R22 ¼  dX1 ; (41) þ ;
ðc  1 þ 2bÞ2 4ðc  1 þ 2bÞ

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-5


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

2ð3c  1ÞR1 P1 ðc  1ÞR21 ðC0 R1 X13 þ 2H1 Þ the hand calculations become very monotonous, so we computed the
D¼ þ :
X1 X12 succeeding terms of the series (15) by writing the program in software
package “Mathematica 7.” The program has two parts for the compu-
Up to second-order approximations, the shock wave location is
tation of coefficients for n  3; Xn ; Un ; Rn ; and Pn . The first part is
given by
devoted to the generation of the system of algebraic equations by using
ðc þ 1Þ C Eqs. (16) and (18) while in the second we evaluate the coefficients
XðtÞ ¼ t þ t 2 þ   : (51) Xnl ; Unl ; Rnl ; and Pnl by solving the algebraic equations obtained in
2ð1  bÞ D
the first part. We calculated the first 31 coefficients of X(t) for different
From Eq. (51), we note that when d ¼ 0 (i.e., in the absence of sets of parameters, which are listed in Table I.
density stratification), the above result reduces to the results obtained
by Chauhan et al.40 in nonideal magnetogasdynamics while for a non- III. GUDERLEY’S EXACT SIMILARITY SOLUTION
ideal gas (b ¼6 0) of constant density (d ¼ 0), it reproduces all the In this section, we determined the Guderley solution for an ideal
results obtained by Arora and Sharma21 in the absence of a magnetic gas of variable density. For the ranges 1 < t  0 and 1  n < 1, we
field. We also clearly see that the results obtained by Madhumita and take the similarity variable n of the following form:
Sharma37 for an ideal gas (b ¼ 0) of varying density (d ¼ 6 0) and Van
r
dyke and Guttmann39 for an ideal gas (b ¼ 0) of constant density n¼ ; (52)
(d ¼ 0) can be easily recovered from the above obtained results. As we ta
can see from Eqs. (46)–(50), after the second-order approximations, where a is the similarity exponent.

TABLE I. Coefficients in the Taylor series expansion given by Eq. (15).

n c ¼ 7=5; b ¼ 0:1; d ¼ 1; C0 ¼ 1=50 c ¼ 5=3; b ¼ 0:01; d ¼ 2; C0 ¼ 1=20 c ¼ 7=5; b ¼ 0; d ¼ 2; C0 ¼ 1=50

1 1.333 333 333 33 1.346 801 346 801 1.200 000 000 0
2 0.185 195 468 110 0.408 223 578 423 0.236 724 409 736
3 0.248 160 744 607 0.540 234 523 130 0.299 693 856 426
4 0.248 080 030 581 0.692 081 665 411 0.332 654 291 800
5 0.280 535 219 015 0.977 445 166 655 0.373 500 413 237
6 0.351 606 876 775 1.522 191 876 90 0.469 866 022 602
7 0.470 325 928 010 2.507 025 497 70 0.649 854 967 067
8 0.653 099 120 412 4.269 008 088 98 0.929 729 041 897
9 0.932 726 264 483 7.463 217 175 20 1.349 073 313 24
10 1.363 571 484 65 13.336 224 375 7 1.998 296 677 07
11 2.031 596 775 67 24.250 574 202 5 3.027 273 441 91
12 3.073 563 389 34 44.720 211 077 1 4.660 350 878 02
13 4.709 324 105 15 83.439 578 410 4 7.250 532 855 90
14 7.293 952 659 00 157.246 786 566 11.385 229 118 8
15 11.402 537 509 6 298.897 152 347 18.043 049 711 0
16 17.969 598 093 5 572.390 245 619 28.827 780 934 1
17 28.519 129 024 8 1103.294 604 29 46.363 974 242 1
18 45.544 605 468 9 2138.903 946 77 74.989 907 715 5
19 73.137 833 343 6 4167.892 348 61 121.924 661 239
20 118.031 998 386 8158.951 085 89 199.188 416 810
21 191.335 442 138 16 037.878 016 4 326.802 578 070
22 311.419 025 739 31 643.509 894 6 538.207 783 946
23 508.733 121 799 62 647.091 695 8 889.438 169 775
24 833.857 400 123 124 413.640 866 1474.580 544 72
25 1370.974 634 50 247 784.466 542 2451.845 420 37
26 2260.457 518 92 494 789.728 258 4087.719 526 41
27 3736.786 454 37 990 422.769 240 6831.910 336 23
28 6192.291 079 91 1:986 993 055 15  106 11 444.591 417 7
29 10284.441 476 0 3:994 648 655 91  106 19 212.585 357 1
30 17116.665 558 1 8:046 465 875 84  106 32 317.208 659 6
31 28543.436 940 1 1:623 753 626 27  107 54 461.095 416 4

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-6


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

All the flow variables density (q), velocity (v), gas pressure (p), n4 ðE  1Þ2
and magnetic pressure (h) in terms of the new independent variable n H¼ ððE  1Þ2 F  cG  2IÞ: (58)
F2
can be rewritten as follows:
The determinants Hn ðn ¼ 1; 2; 3; 4Þ are identified from H by
   2  2 replacing the nth column of L by the column vector N.
ar ar ar
v¼ EðnÞ; q ¼ q0 EðnÞ; p ¼ q0 GðnÞ; h ¼ q0 IðnÞ: At the instant of collapse, t ¼ 0 (i.e., n ¼ 0), all the flow variables,
t t t
speed of sound, and Alfven speed are bounded for any finite radius.
(53) Therefore, for the boundedness of variables defined in Eq. (53), we
Here, the initial density q0 varies according to power law, i.e., must have
q0 ¼ qc ðRr0 Þd , where d is the density exponent.
Gð1Þ Ið1Þ
Under the transformation given by Eq. (53), the system of equa- Eð1Þ ¼ 0; ¼ 0; ¼ 0: (59)
tions (1)–(4) get transformed into a new system of equations as Fð1Þ Fð1Þ
follows: Equations (57)–(59) together with boundary conditions given by Eq.
0 0 (55) constitute a boundary value problem that can be solved for the
nðE  1ÞF þ nFE þ FEðd þ 2Þ ¼ 0;
  flow variable behind the shock. However, for this purpose, we first
1 1 need to determine the unknown parameter a appearing in Eq. (57). It
nðE  1ÞE0 þ E E  þ ððd þ 2ÞðG þ IÞ þ nðG0 þ I 0 ÞÞ ¼ 0;
a F can be seen that H > 0 at n ¼ 1 and H < 0 at n ¼ 1; this shows that
 
1 dE (54) there must exist a n ¼ nc in ð1; 1Þ at which H ¼ 1 and solution of
nðE  1ÞG0 þ c n GF 0 þ 2G  þ þ E þ cE ¼ 0; Eq. (57) becomes singular. For the nonsingular solution of the above
a 2
  boundary value problem, we need to choose the value of a such that H
1 dE and all Hn ; n ¼ 1; 2; 3; 4 become zero at nc. For this purpose, we
nðE  1ÞI 0 þ 2nIE0 þ 2I  þ þ 3E ¼ 0;
a 2 introduced a new variable Z of the following form:
where prime denotes the derivative with respect to the similarity vari- Z ¼ FðE2 þ 1  2EÞ  cG  2I; (60)
able n.
At the strong shock front, the jump conditions are given as whose derivative is given as follows:
follows:
dZ 1
cþ1 2 ¼ ð2FðE  1ÞH1 þ ðE  1Þ2 H2  cH3  2H4 Þ: (61)
Eð1Þ ¼ ; Fð1Þ ¼ ; dn H
c1 cþ1
!   (55) Therefore, the system given by Eq. (57) in view of Eq. (61)
2 2cC0 C0 c þ 1 2 becomes
Gð1Þ ¼  ; Ið1Þ ¼ :
c þ 1 ðc  1Þ2 2 c1
dE H1 dF H2 dG H3 dI H4
¼ ; ¼ ; ¼ ; ¼ ; (62)
The matrix notation for the system given by Eq. (54) is given by dZ H5 dZ H5 dZ H5 dZ H5

LM 0 ¼ N; (56) where H5 ¼ 2FðE  1ÞH1 þ ðE  1Þ2 H2  cH3  2H4 :


We obtained the numerical solution to the system of ordinary
where M ¼ ðE; F; G; IÞtr , and the matrix L and the column vector N differential equations given by Eq. (62) together with the boundary
can be obtained from the system given by Eq. (54). The entity a conditions given by Eq. (55) by using the explicit Runge–Kutta
appearing in the system given by Eq. (54) is unknown and can be method of order four. For the trial value of a, we integrate the system
obtained by solving the nonlinear problem for a system of ordinary of Eq. (62) from the shock, Z ¼ Z½1, to the singular point Z ¼ 0, and
differential equations. The system given by Eq. (54) is solved by using compute the values of E; F; G; I, and H1 at Z ¼ 0. If the determinant
Cramer’s rule for E0 ; F 0 ; G0 , and I 0 as follows: H1 becomes zero corresponding to this trial value of a, it means we
get the correct value of a; we further correct this value of a by succes-
H1 H2 H3 H4
E0 ¼ ; F0 ¼ ; G0 ¼ ; I0 ¼ ; (57) sive approximations in such a manner that corresponds to these val-
H H H H ues, the determinant H1 vanishes at Z ¼ 0. We calculated the value of
where the determinant H of the matrix L is as follows: a for different values of parameters c; C0 ; d (listed in Table II).

TABLE II. Neville table for similarity exponents a1 for d ¼ 2; c ¼ 5=3; b ¼ 0:01; C0 ¼ 1=20.

n e0n Linear Quadratic Cubic Quartic

27 0.598 853 96 0.597 872 58 0.597 486 22 0.597 051 69 0.608 647 13
28 0.598 818 00 0.597 847 01 0.597 514 65 0.597 751 52 0.601 950 52
29 0.598 783 59 0.597 820 16 0.597 457 61 0.596 963 29 0.592 036 86
30 0.598 750 47 0.597 790 07 0.597 368 90 0.596 570 49 0.594 017 31
31 0.598 718 48 0.597 758 86 0.597 306 33 0.596 722 39 0.597 747 67

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-7


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

IV. ANALYSIS OF THE COEFFICIENTS OF SERIES TABLE III. Calculated similarity exponents and corresponding amplitudes.
FOR SHOCK WAVE POSITION
c d b C0 Exponents Amplitudes
By following the similar analysis as in Refs. 40 and 47, we see that
at time t ¼ 1, the power series solution breaks down because at this 7/5 1 0.1 1/50 a1 ¼ 0:707 935 71 A1 ¼ 0:976 702 12
time the piston itself reaches the axis. Now, we will show that this sin- a2 ¼ 2:175 128 12 A2 ¼ 0:032 497 28
gularity is the Guderley singularity, whose position is given by 7/5 1 0.1 1/20 a1 ¼ 0:707 521 27 A1 ¼ 0:974 150 77
P
RðtÞ ¼ 1  XðtÞ¼ 1  Xn t n a2 ¼ 2:131 539 06 A2 ¼ 0:033 769 20
a1
t (63) 7/5 2 0.1 1/50 a1 ¼ 0:632 163 79 A1 ¼ 0:992 347 61
A1 1  as t ! tc ;
tc a2 ¼ 2:887 003 56 A2 ¼ 0:009 902 33
7/5 2 0.1 1/20 a1 ¼ 0:629 794 46 A1 ¼ 0:986 570 08
where, a1 is the similarity exponent and A1 is the corresponding
a2 ¼ 2:444 226 36 A2 ¼ 0:022 873 67
amplitude.
7/5 1 0.01 1/50 a1 ¼ 0:711 171 49 A1 ¼ 0:980 289 91
Therefore,
  a2 ¼ 2:202 687 43 A2 ¼ 0:028 008 92
Xn 1 1 þ a1 7/5 1 0.01 1/20 a1 ¼ 0:706 537 06 A1 ¼ 0:975 428 10
 1 as n ! 1: (64)
Xn1 tc n a2 ¼ 2:178 080 99 A2 ¼ 0:033 052 841
a3 ¼ 3:991 298 16 A3 ¼ 0:000 612 151
From Eq. (64), we see that the value of 1=tc can be approximated
5/3 1 0.01 1/20 a1 ¼ 0:685 748 16 A1 ¼ 0:986 981 94
by the ratio Xn =Xn1 , for large values of n. Now, we have constructed
a sequence of ratios Xn =Xn1 for n ¼ 22, 23, …, 31, and refined this a2 ¼ 2:097 255 43 A2 ¼ 0:021 738 63
estimate of 1=tc by forming a Neville Table VI (see the Appendix) for 5/3 2 0.01 1/20 a1 ¼ 0:599 879 48 A1 ¼ 0:989 355 20
the radius of convergence tc and Neville Table VII (see the Appendix) a2 ¼ 2:201 596 4 A2 ¼ 0:021 430 31
for estimating following a procedure similar to Gaunt and
Guttmann.46 We also estimated the leading similarity exponents a1,
which are shown in Table II. In this section, all the real exponents and their corresponding ampli-
tudes are determined by following the procedure given in Refs. 39 and
V. STRUCTURE OF LOCAL SINGULAR SOLUTION 47 and are listed in Table III.
Van Dyke and Guttmann39 found all the real exponents together
with the corresponding amplitudes for an ideal gas of constant density VI. RESULTS AND DISCUSSION
by using the theory of Baker and Hunter.45 The latter have devised a In the present work, the problem of an imploding shock wave in
technique for detecting confluent singularities when an accurate esti- a nonideal gas of varying density is studied under the effect of an axial
mate of the location of the dominant singularity is available. This magnetic field. The initial density of the gas is considered to be
involves the creation of an auxiliary function from the given function. decreasing toward the axis of collapse followed by a power law. All the

TABLE IV. Comparison of leading similarity exponents with Madhumita and Sharma37 and Hafner.4

Madhumita and Error between a1 and Error between a1 and


c d B C0 Computed a1 Sharma37 Madhumita and Sharma37 Hafner4 Hafnar4

7/5 1 0.1 0.02 0.707 935 71   


7/5 2 0.1 0.02 0.629 794 46   
7/5 1 0.01 0.02 0.711 171 49   
7/5 2 0.01 0.02 0.630 579 46   
7/5 1 0.01 0.05 0.706 537 06   
7/5 2 0.01 0.05 0.635 296 05   
7/5 0 0.00 0.00 0.835 818 06   0.835 323 19 0:4948  103
7/5 1 0.00 0.00 0.715 359 02 0.714 922 78 0:4362  103 
7/5 2 0.00 0.00 0.628 716 76 0.628 341 36 0:3754  103 0.628 341 71 0:3750  103
5/3 1 0.01 0.05 0.685 748 16  
5/3 2 0.01 0.05 0.599 879 48  
5/3 0 0.00 0.00 0.815 997 45   0.815 624 90 0:3725  103
5/3 1 0.00 0.00 0.688 876 48 0.688 373 55 0:5029  103  
5/3 2 0.00 0.00 0.600 104 84 0.599 446 01 0:6588  103 0.599 446 00 0:6588  103
6/5 0 0.00 0.00 0.863 744 89   0.861 163 02 0:2581  102
6/5 2 0.00 0.00 0.675 921 73   0.671 763 20 0:4158  102

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-8


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

TABLE V. Comparison of leading similarity exponents with Guderley’s similarity exponents (ideal gas b ¼ 0).

c d C0 Computed a1 Guderley’s2 a Error between a1 and Guderley’s2 a

7/5 1 0.02 0.711 489 15 0.710 62 0:8691  103


7/5 1 0.05 0.707 528 30 0.704 45 0:3078  102
7/5 2 0.02 0.631 429 69 0.630 55 0:8796  103
7/5 2 0.05 0.638 724 86 0.632 55 0:6174  102
5/3 1 0.02 0.687 598 97 0.686 80 0:7989  103
5/3 1 0.05 0.685 608 33 0.684 55 0:1058  102
5/3 2 0.02 0.600 095 75 0.599 25 0:8457  103
5/3 2 0.05 0.599 894 00 0.599 05 0:8440  103

FIG. 2. Comparison of leading similarity exponents with Madhumita and Sharma37 and Hafner.4

FIG. 3. Comparison of leading similarity exponents with Guderley’s similarity exponents (ideal gas b ¼ 0).

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-9


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

flow variables and shock location are analyzed using the perturbation following values of the parameters: c ¼ 7=5; 5=3; b ¼ 0:1; 0:01;
series approach.39 The perturbation series approach gives the global d ¼ 1; 2; and C0 ¼ 1=20; 1=50. Although, Guderley’s local self-similar
solution to the converging problem with which the Guderley local solution yields only the first dominant similarity exponent, it provides
solution in the neighborhood of shock collapse can be extracted easily. other less dominant similarity exponents and their corresponding
From the coefficients in the perturbation series, the first three similar- amplitudes as well. The values of the leading similarity exponents cal-
ity exponents and their corresponding amplitudes have been extracted culated by the perturbation series approach are compared with the
in Guderley’s local expansion, which are depicted in Table III for the similarity exponents calculated by Madhumita and Sharma37 and

FIG. 4. Flow profiles of (a) velocity, (b) density, (c) gas pressure, (d) magnetic pressure, and (e) shock trajectory behind a strong cylindrical shock wave in a nonideal gas for
d ¼ 2, c ¼ 7=5, and C0 ¼ 0:02.

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-10


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

Hafner,4 which are listed in Table IV for various values of c; b; C0 , and and IV(b) are represented graphically (see Figs. 2 and 3), clearly showing
d. It can be clearly seen from Table IV that the calculated results match that the similarity exponents calculated by the perturbation series tech-
well with those obtained by other authors. In the case of an ideal gas of nique match well with those obtained by other authors. From Tables IV
variable density under the presence of an axial magnetic field, the results and V, it can be seen that the value of a1 is always less than unity; this is
of the leading similarity exponent a1 obtained by the present method due to the fact that the shock accelerated continuously and became
are compared with those obtained with Guderley’s exact similarity unbounded as t ! tc , but less rapidly than ðt  tc Þ1 .
method, which are listed in Table V; it can be seen that results obtained We have shown the Neville Table II in Sec. IV and the Neville
show competence. Moreover, the results obtained in Tables IV Tables VI and VII (in the Appendix) for estimating the similarity

FIG. 5. Flow profiles of (a) velocity, (b) density, (c) gas pressure, (d) magnetic pressure, and (e) shock trajectory behind a strong cylindrical shock wave in a nonideal gas for
d ¼ 1, c ¼ 5=3, and C0 ¼ 0:05.

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-11


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

exponent a1, the reciprocal radius of convergence 1=tc , and the time tc Table III by the Pade approximation. The value of a1 (i.e., leading sim-
of shock wave to collapse at the axis of symmetry, respectively, for the ilarity exponent) is listed in Tables IV and V for various values of
following set of parameters c ¼ 5=3; d ¼ 2; C0 ¼ 1=20, and b ¼ 0.01. c; d; b; and C0 and it can be seen that an increment in the values
The values of tc seen in Tables VI and VII are almost the same; in fact c; d; b and decrement in the value of C0 cause a1 to decrease, due to
Table VII gives a more accurate value of tc as compared to that which the shock gets accelerated toward the axis of symmetry. For dif-
obtained from Table VI. The value of the leading similarity exponent ferent sets of values of parameters c, b, C0, and d, the time of shock col-
a1 obtained from Table II also matches well with that obtained from lapse is calculated and listed in Table VIII (see the Appendix). It can

FIG. 6. Flow profiles of (a) velocity, (b) density, (c) gas pressure, (d) magnetic pressure, and (e) shock trajectory behind a strong cylindrical shock wave in a nonideal gas for
b ¼ 0.01, c ¼ 7=5, and C0 ¼ 0:02.

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-12


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

be observed from Table VIII that an increase in C0 causes the time of Figures 4(a)–8(a) and 4(b)–8(b) show that on moving toward
shock collapse tc to increase, whereas tc decreases with increase in the the piston, there is a monotonic decrement in velocity while density
values of b; d; and c by keeping other parameters fixed. shows an increasing trend in the region behind the shock. This hap-
The profiles for the flow variables, viz., the velocity, density, pres- pens because of the contraction of area or due to geometrical con-
sure, magnetic pressure, and shock path for various sets of parameters vergence. In the region behind the shock wave, the nonideal gas,
have been plotted, showing the effects of varying the nonideal parame- which is highly compressed by the shock, gets cooled down and
ter b, the density stratification d, and the shock cowling number C0 of causes the density to decrease ahead of the shock front at the same
gas on flow parameters. rate at which the square of the front velocity increases. The gas

FIG. 7. Flow profiles of (a) velocity, (b) density, (c) gas pressure, (d) magnetic pressure, and (e) shock trajectory behind a strong cylindrical shock wave in a nonideal gas for
b ¼ 0.01, c ¼ 5=3, and C0 ¼ 0:05.

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-13


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

pressure [see Figs. 4(c)–8(c)] and magnetic pressure [see trajectory show an increase with increase in time t. From
Figs. 4(d)–8(d)] remain bounded in the region behind the shock Figs. 4(e)–8(e), it is observed that in the neighborhood of the pis-
wave and are even stationary in most of the region, except in the ton, the shock trajectories coincide, indicating that the geometrical
vicinity of collapse. This behavior may also exist because of the convergence effects are initially small at large radii. In fact, shock
high compression of the nonideal gas by shock waves in the region accelerates slowly in the intermediate region, while near the axis of
behind the shock. In Figs. 4(e)–8(e), the shock trajectory X(t) has symmetry it accelerates rapidly due to adiabatic compression of the
been plotted with respect to time t. The profiles of the shock shocked state because of flow area convergence.

FIG. 8. Flow profiles of (a) velocity, (b) density, (c) gas pressure, (d) magnetic pressure, and (e) shock trajectory behind a strong cylindrical shock wave in a nonideal gas for
c ¼ 7=5; b ¼ 0:01; and d ¼ 2.

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-14


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

The effects of variation in the nonideal parameter (b), density while the perturbation series method is an approach that provides
exponent (d) and shock cowling number (C0) on the flow variables, all possible similarity exponents together with their corresponding
and shock trajectory are discussed below. amplitudes. From the coefficients in the perturbation series, the first
The effect of parameter b can be seen on the profiles of the flow three similarity exponents and their corresponding amplitudes are
variables and shock trajectory in Figs. 4 and 5 for d ¼ 2; c ¼ 7=5; extracted in Guderley’s local expansion, which are depicted in
C0 ¼ 0:02 and d ¼ 1; c ¼ 5=3; C0 ¼ 0:05, respectively. Figures 4(a), Table III. Also, for an ideal gas of variable density under the pres-
4(c) and 5(a), 5(c) show that as the value of the nonideal parameter b ence of axial magnetic field, the similarity exponents are calculated
increases, the velocity and pressure increase, whereas the density and by Guderley’s method (see Sec. III). The Neville Table II (see
magnetic pressure decrease for the given values of c; d, and C0, as Sec. IV) is formed to estimate the similarity exponent a1 and
shown in Figs. 4(b), 4(d) and 5(b), 5(d). As the value of the nonideal Tables VI and VII (see the Appendix) are constructed to estimate
gas parameter b increases, the particles collide more frequently and the values of 1=tc and tc, respectively. The calculated results are
in turn generate high pressure. From Figs. 4(e) and 5(e), it can be compared for leading similarity exponents a1 with the already exist-
observed that an increase in the value of b for the given values of c; d, ing results by other authors and are listed in Table IV. The flow
and C0 causes a decrease in the time of shock collapse tc, i.e., the variables and shock position have been analyzed graphically with
shock approaches the axis much faster for large values of b. This can respect to the variation of different values of van der Waals
also be seen in Table VIII. The effect of density stratification d on the excluded volume b, density exponent d, and shock cowling number
flow variable and shock trajectory can be visualized in Figs. 6 and 7 C0. From the present study, the following can be concluded:
for the following set of parameters: b ¼ 0:01; c ¼ 7=5; C0 ¼ 0:02
and b ¼ 0:01; c ¼ 5=3; C0 ¼ 0:05, respectively. Furthermore, an (1) The leading similarity exponents obtained by this method for dif-
increment in the density index d causes the velocity to increase [see ferent values of the parameters c; b; d, and C0 match well with the
Figs. 6(a) and 7(a)] and the density, gas pressure, and magnetic pres- already existing results (Madhumita and Sharma,37 Hafner4 as
sure [see Figs. 6(b), 6(c), 6(d), 7(b), 7(c), and 7(d)] to decrease. well as with Guderley’s method2) [see Tables IV and V]. We have
Moreover, Figs. 6(e) and 7(e) indicate that as the value of parameters also shown these results graphically in Figs. 2 and 3.
d increases, the time of shock collapse decreases and the shock (2) The value of a1 is always less than unity [see Tables IV and V],
reaches the axis much faster. In Fig. 8, the effect of C0 is shown on showing the fact that the shock accelerated continuously and
the profiles of the flow variables and shock trajectory for became unbounded as t ! tc , but less rapidly than ðt  tc Þ1 .
c ¼ 7=5; d ¼ 2, and b ¼ 0.01. As the value of C0 increases, the veloc- (3) The value of the leading similarity exponent a1 shown in the
ity and pressure decrease [see Figs. 8(a) and 8(c)], whereas the den- Neville Table II matches well with that shown in Table III by
sity and magnetic pressure show an opposite trend on incrementing Pade approximation. Also, the values of tc shown in Tables VI
the value of C0 [see Figs. 8(b) and 8(d)] behind the shock. and VII are almost the same.
Decrement in the value of C0 causes an increment in time of (4) The time of shock collapse tc increases with increment in the
shock collapse tc, i.e., for a small value of C0; the shock collapses much value of C0 whereas it decreases with increment in the values
faster in comparison to the large value of C0. Due to the presence of an of b; d; c by keeping other parameters fixed (see Table VIII).
axial magnetic field in the gaseous medium, the charged gas particle These results are verified graphically in Figs. 4(e)–8(e).
transports toward the shock front, resulting in an increment in the Furthermore, for larger values of b and d, shock collapse
density of particles. occurs much faster [see Figs. 4(e)–7(e)], whereas for large val-
ues of C0, the shock wave takes a greater time to collapse to
VII. CONCLUSION the axis [see Fig. 8(e)].
In the present work, an imploding strong shock wave problem (5) As there is a movement toward the piston from the axis of col-
has been investigated for a cylindrical symmetric flow of a nonideal lapse, the velocity decreases monotonically while the density
gas of variable density under the significant presence of magnetic shows an increasing trend in the region behind the shock [see
field effect. To generate cylindrically symmetric shock waves, a Figs. 4(a)–8(a) and 4(b)–8(b)]. The results in Figs. 4(c)–4(d) and
cylindrical piston is considered, which is filled with a nonideal gas 8(d)–8(d) also indicate that the gas pressure and magnetic pres-
(van der Waals type) of variable density instead of constant density sure remain bounded in the region behind the shock; indeed, the
under the presence of axial magnetic field with the assumption that gas density, pressure, and magnetic pressure are stationary in
the density of the gas is decreasing toward the axis of collapse fol- most of the region except in the vicinity of the front where a max-
lowing a power-law and the piston starts contracting with a very imum is attained for the value of density exponent d ¼ 2.
high constant speed and producing strong cylindrical shock waves. (6) As the value of b increases, the velocity and pressure enhance
The unknown position of the shock as well as all the flow variables [see Figs. 4(a), 4(c) and 5(a), 5(c)] while the density and mag-
behind the shock are determined by using the perturbation series netic pressure are reinforced [see Figs. 4(b), 4(d) and 5(b),
technique suggested by Van Dyke and Guttmann39 and the global 5(d)].
solution is obtained. With the help of this technique, the whole flow (7) As the value of density exponent is increased d, the velocity
field extending from the piston to the axis of collapse is analyzed, increases [see Figs. 6(a) and 7(a)], whereas the density, gas pres-
where Guderley’s local analysis cannot provide the solution. From sure, and magnetic pressure decrease [see Figs. 6(b), 6(c), 6(d)
the obtained global solution, Guderley’s local solution is extracted, and 7(b), 7(c), 7(d)] behind the shock.
which is valid only in the neighborhood of shock collapse. (8) It can be observed that an increase in the value of C0 results in a
Guderley’s method provides only the first similarity exponents, decrease in the velocity and pressure [see Figs. 8(a) and 8(c)]

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-15


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

and an increase in density and magnetic pressure [see Figs. 8(b) DATA AVAILABILITY
and 8(d)] for fixed values of c; b, and d. The data that support the findings of this study are available
within the article.
ACKNOWLEDGMENTS
APPENDIX: ESTIMATION OF THE RADIUS OF
Antim Chauhan acknowledges the research support from
“University Grant Commission (Government of India).” [Sr. No. CONVERGENCE AND TIME TAKEN BY SHOCK TO
2121541039 with Ref No. 20/12/2015 (ii) EU-V]. COLLAPSE

AUTHOR DECLARATIONS Neville table for estimating 1=tc for c ¼ 5=3, d ¼ 2; b ¼ 0:01;
C0 ¼ 1=20. Neville table for estimating the radius of convergence tc
Conflict of Interest for c ¼ 5=3; d ¼ 2; b ¼ 0:01; C0 ¼ 1=20. Values of tc, the time
The authors declare that they have found no conflict of interest. taken by the shock to collapse.

TABLE VI. Neville table for estimating 1=tc for c ¼ 5=3, d ¼ 2; b ¼ 0:01; C0 ¼ 1=20.

n e0n Linear Quadratic Cubic Quartic

27 2.001 704 38 2.127 780 310 2.127 732 881 2.127 738 52 2.126 627 32
28 2.006 206 96 2.127 776 517 2.127 727 19 2.127 679 84 2.127 327 74
29 2.010 398 90 2.127 773 210 2.127 728 56 2.127 740 45 2.128 119 27
30 2.014 311 28 2.127 770 46 2.127 732 00 2.127 762 91 2.127 908 93
31 2.017 971 18 2.127 768 05 2.127 733 20 2.127 744 37 2.127 619 20

TABLE VII. Neville table for estimating the radius of convergence tc for c ¼ 5=3; d ¼ 2; b ¼ 0:01; C0 ¼ 1=20.

n e0n Linear Quadratic Cubic Quartic

27 0.470 034 624 0.469 969 312 0.469 984 799 0.469 982 40 0.470 246 39
28 0.470 032 33 0.469 970 505 0.469 986 017 0.469 996 17 0.470 078 81
29 0.470 030 23 0.469 971 545 0.469 985 592 0.469 981 90 0.469 892 74
30 0.470 028 31 0.469 972 423 0.469 984 703 0.466 823 85 0.469 942 83
31 0.470 026 53 0.469 973 192 0.469 984 352 0.466 827 24 0.470 010 67

TABLE VIII. Values of tc, the time taken by the shock to collapse.

c b C0 tc ðd ¼ 1Þ tc ðd ¼ 2Þ

7/5 0.01 1/20 0.628 071 07 0.571 391 81


7/5 0.01 1/50 0.627 744 91 0.565 423 54
7/5 0.001 1/50 0.633 498 70 0.571 358 97
5/3 0.01 1/20 0.538 999 92 0.479 248 08

6
REFERENCES N. F. Ponchaut, H. G. Hornung, D. I. Pullin, and C. A. Mouton, “On implod-
1 ing cylindrical and spherical shock waves in a perfect gas,” J. Fluid Mech. 560,
G. Jagadeesh, “Fascinating world of shock waves,” Resonance 13(8), 752–767 103–122 (2006).
(2008). 7
P. L. Sachdev, N. Gupta, and D. S. Ahluwalia, “Exact analytic solutions describ-
2
G. Guderley, “Starke kugelige und zylindrische Verdichtungsstosse in der Nahe ing unsteady plane gas flows with shocks of arbitrary strength,” Q. Appl. Math.
des Kugelmittelpunktes bzw der Zylinderachse,” Luftfahrtforschung 19, 50(4), 677–726 (1992).
302–312 (1942). 8
Z. M. Boyd, S. D. Ramsey, and R. S. Baty, “On the existence of self-similar con-
3
R. B. Lazarus and R. D. Richtmyer, “Similarity solutions for converging shocks,” verging shocks in non-ideal materials,” Q. J. Mech. Appl. Math. 70(4),
Report No. LA-6823-MS (Los Alamos Scientific Laboratory, NM, 1977). 401–417 (2017).
4 9
P. Hafner, “Strong convergent shock waves near the center of convergence: A R. F. Chisnell, “An analytic description of converging shock waves,” J. Fluid
power series solution,” SIAM J. Appl. Math. 48(6), 1244–1261 (1988). Mech. 354, 357–375 (1998).
5 10
R. Sari, N. Bode, A. Yalinewich, and A. MacFadyen, “Slightly two- or three- G. B. Whitham, Linear and Nonlinear Waves (Wiley-Interscience, New York,
dimensional self-similar solutions,” Phys. Fluids 24, 087102 (2012). 1974), Chap. 8.

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-16


Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

11 30
S. D. Ramsey, E. M. Schmidt, Z. M. Boyd, J. F. Lilieholm, and R. S. Baty, L. P. Singh, A. Husain, and M. Singh, “A self-similar solution of exponential
“Converging shock flows for a Mie-Gr€ uneisen equation of state,” Phys. Fluids shock waves in non-ideal magnetogasdynamics,” Meccanica 46, 437–445
30(4), 046101 (2018). (2011).
12 31
R. Ram and B. D. Pandey, “Propagation of weak MHD waves in steady hyper- R. Shyam, L. P. Singh, and V. D. Sharma, “Steepening of waves in radiative
sonic flows with radiation,” AIAA J. 18, 855–857 (1980). magnetohydrodynamics,” Acta Astronaut. 13, 95–100 (1986).
13 32
L. P. Singh, M. Singh, and B. D. Pandey, “Analytical solution of converging L. P. Singh, M. Singh, and A. Husain, “Similarity solutions of imploding shocks
shock wave in magnetogasdynamic,” AIAA J. 48, 2523–2528 (2010). in non-ideal magnetogasdynamics,” Astrophys. Space Sci. 331, 597–603
14
S. Taniguchi and T. Ruggeri, “On the sub-shock formation in extended (2011).
33
thermodynamics,” Int. J. Non Linear Mech. 99, 69–78 (2018). L. P. Singh, S. D. Ram, and D. B. Singh, “Quasi-similar solution of the strong
15
Z. Zhai, Y. Liang, L. Liu, J. Ding, X. Luo, and L. Zou, “Interaction of rippled shock wave problem in non-ideal gas dynamics,” Astrophys. Space Sci. 337(2),
shock wave with flat fast-slow interface,” Phys. Fluids 30, 046104 (2018). 597–604 (2012).
16 34
S. Shah and R. Singh, “Collision of a steepened wave with a blast wave in dusty A. Sakurai, “Propagation of spherical shock waves in stars,” J. Fluid Mech. 1(4),
real reacting gases,” Phys. Fluids 31, 076103 (2019). 436–453 (1956).
17 35
M. Liverts and N. Apazidis, “Limiting temperatures of spherical shock wave A. Sakurai, “On the problem of a shock wave arriving at edge of a gas,”
implosion,” Phys. Rev. Lett. 116, 014501 (2016). Commun. Pure Appl. Math. 13, 353–370 (1960).
18 36
C. C. Wu and P. H. Roberts, “Structure and stability of a spherical shock wave M. H. Rogers, “Analytic solutions for blast wave problem with an atmosphere
in a Van der Waals gas,” Q. J. Mech. Appl. Math. 49(4), 501–543 (1996). of varying density,” Astrophys. J. 125, 478–493 (1957).
19 37
N. Zhao, A. Mentrelli, T. Ruggeri, and M. Sugiyama, “Admissible shock waves G. Madhumita and V. D. Sharma, “Propagation of strong converging shock
and shock induced phase transitions in a Van der Waals fluid,” Phys. Fluids waves in a gas of variable density,” J. Eng. Math. 46, 55–68 (2003).
38
23(8), 86–101 (2011). B. H. K. Lee, “The initial phases of collapse of an imploding shock wave
20
K. Ambika, R. Radha, and V. D. Sharma, “Progressive waves in non-ideal and the application to hypersonic internal flow,” C.A.S.I. Trans. 1, 57–67
gases,” Int. J. Non Linear Mech. 67, 285–290 (2014). (1968).
21 39
R. Arora and V. D. Sharma, “Convergence of strong shock in a Van Der Waals M. Van Dyke and A. J. Guttmann, “The converging shock wave from a spheri-
gas,” SIAM J. Appl. Math. 66(5), 1825–1837 (2006). cal or cylindrical piston,” J. Fluid Mech. 120, 451–462 (1982).
22 40
A. Tomar, R. Arora, and A. Chauhan, “Propagation of strong shock waves in a A. Chauhan, R. Arora, and A. Tomar, “Convergence of strong shock waves in
non-ideal gas,” Acta Astronaut. 159, 96–104 (2019). a non-ideal magnetogasdynamics,” Phys. Fluids 30(11), 116105 (2018).
23 41
R. Arora, A. Tomar, and V. P. Singh, “Shock waves in reactive hydro- S. D. Ramsey and R. S. Baty, “Piston driven converging shock waves in a stiff-
dynamics,” Shock waves 19(2), 145–150 (2009). ened gas,” Phys. Fluids 31(8), 086106 (2019).
24 42
A. Chauhan and R. Arora, “Kinematics of spherical shock waves in an interstel- D. Singh and R. Arora, “Piston driven converging cylindrical shock waves in a
lar ideal gas clouds with dust particles,” Math. Methods Appl. Sci. 44(8), non-ideal gas with azimuthal magnetic field,” Phys. Fluids 32, 126116 (2020).
43
6282–6300 (2021). I. Giron, S. Balberg, and M. Krief, “Solutions of the imploding shock problem
25
P. K. Sahu, “Cylindrical shock waves in rotational axisymmetric non-ideal in a medium with varying density,” Phys. Fluids 33, 066105 (2021).
44
dusty gas with increasing energy under the action of monochromatic radi- E. Modelevsky and R. Sari, “Revisiting the strong shock problem: Converging
ation,” Phys. Fluids 29, 086102 (2017). and diverging shocks in different geometries,” Phys. Fluids 33, 056105
26
R. M. Lock and A. J. Mestel, “Annular self-similar solutions in ideal magneto- (2021).
45
gasdynamics,” J. Plasma Phys. 74(4), 531–554 (2008). G. A. Baker and D. L. Hunter, “Methods of series analysis II. Generalized and
27
B. Bira, T. R. Sekhar, and G. P. R. Sekhar, “Collision of characteristic shock extended methods with applications to the Ising model,” Phys. Rev. B 7,
with weak discontinuity in non-ideal magnetogasdynamics,” Comput. Math. 3377–3392 (1973).
46
Appl. 75(11), 3873–3883 (2018). D. S. Gaunt and A. J. Guttmann, Asymptotic Analysis of Coefficients in Phase
28
V. D. Sharma, “Development of jump discontinuities in radiative magnetogas- Transitions and Critical Phenomena, edited by C. Domb and M. S. Green
dynamics,” Int. J. Eng. Sci. 24(5), 813–818 (1986). (Academic, New York, 1974), Vol. 3, pp. 181–243.
29 47
S. D. Ram, R. Singh, and L. P. Singh, “An exact analytical solution of the strong A. Chauhan, R. Arora, and A. Tomar, “Converging shock waves in a van der
shock wave problem in non-ideal magnetogasdynamics,” J. Fluids 2013, Waals gas of variable density,” Q. J. Mech. Appl. Math. 73(2), 101–118
810206. (2020).

Phys. Fluids 33, 116110 (2021); doi: 10.1063/5.0064883 33, 116110-17


Published under an exclusive license by AIP Publishing

You might also like