Vous êtes sur la page 1sur 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/260605219

Analytical Solutions for Prandtl-Meyer Wave – Oblique Shock Overtaking


Interaction

Article  in  Acta Astronautica · June 2014


DOI: 10.1016/j.actaastro.2014.02.025

CITATIONS READS

27 440

3 authors, including:

M.V. Chernyshov
Baltic State Technical University
40 PUBLICATIONS   165 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

shock waves View project

Blast mitigation devices View project

All content following this page was uploaded by M.V. Chernyshov on 20 November 2017.

The user has requested enhancement of the downloaded file.


This article appeared in a journal published by Elsevier. The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/authorsrights
Author's personal copy

Acta Astronautica 99 (2014) 175–183

Contents lists available at ScienceDirect

Acta Astronautica
journal homepage: www.elsevier.com/locate/actaastro

Analytical solutions for Prandtl–Meyer wave–oblique shock


overtaking interaction
M.V. Silnikov a,b, M.V. Chernyshov a,n, V.N. Uskov c
a
Saint Petersburg State Polytechnical University, 29 Politechnicheskaya Street,195251 St. Petersburg, Russia
b
Special Materials Corp., 28A Bolshoy Sampsonievsky Avenue, 194044 St. Petersburg, Russia
c
Baltic State Technical University “VoenMech”, 1 1st Krasnoarmeyskaya Street, 190005 St. Petersburg, Russia

a r t i c l e i n f o abstract

Article history: The interaction between oblique shock and preceding expansion or compression Prandtl–
Received 13 February 2014 Meyer wave of the same direction are studied theoretically. Two reliable analytical
Accepted 25 February 2014 solutions for overtaking Prandtl–Meyer wave–oblique shock interaction were obtained.
Available online 6 March 2014
The criteria of reflected wave type change, shock inflection and degeneration, occurrence
Keywords: of subsonic flow pockets and influence of ratio of gas specific heats are established
Supersonic flow analytically.
Shock–expansion interaction The results obtained can be used for the design of re-entry capsules, supersonic inlets,
Analytical solutions and supersonic nozzles, where wave interactions considered in the present paper
take place.
& 2014 IAA. Published by Elsevier Ltd. All rights reserved.

1. Introduction relations as well as the complexity of the flow (the


interaction is not localized at a single point, and even in
In aerodynamic streaming flows of bodies with com- any finite region).
plex geometry the problems of shock waves and Riemann In the case to be solved, uniform gas stream with Mach
waves interaction often take place. Despite of availability number M0 (Fig. 1) turns along the plane wall, and simple
of different numerical solvers the role of exact theoretical (non-centered, Fig. 1a and c) or centered (Fig. 1b and d)
solutions does not decrease, because verification basis for Prandtl–Meyer wave r1 originates. The oblique stationary
numerical tools in gas dynamics is very limited. Present shock j2 is situated downstream the isentropic wave. The
paper is aimed at receiving theoretical solution for the supersonic gas flow in the region 1 between the wave r1 and
problem of Prandtl–Meyer waves and shock waves inter- the shock j2 is characterized by its Mach number M1, and flow
action in streaming flows. angle θ1. Flow direction in the region 1 is horizontal  in the
The problems of interactions between Prandtl–Meyer chosen coordinate system: θ1 ¼ 0. The distance OA ¼ 1
waves and oblique stationary shocks are classical in the between the point O of the intersection of the last Prandtl–
theory of discontinuities interactions in gas dynamics. Meyer straight characteristic OB with the streamlined surface
Owing to the numerical methods, these problems can be and the corner point A is the length scale.
solved in each separate case. But, in spite of the long Straight shock j2 becomes curvilinear at its intersection
history of studies (e.g., see [1]), full theoretical analysis of with the last characteristic OB of the wave r 1 . The reflected
the solution has not been executed due to lack of analytical weak discontinuity BB1 (“weak” means the discontinuity of
not flow parameters, but of their first spatial derivatives)
comes out at the point B, if the flow behind the shock j2 is
n
Corresponding author. supersonic. This discontinuity is a border between the
E-mail address: mvcher@mail.ru (M.V. Chernyshov). uniform flow region 2 behind the shock, and the reflected

http://dx.doi.org/10.1016/j.actaastro.2014.02.025
0094-5765 & 2014 IAA. Published by Elsevier Ltd. All rights reserved.
Author's personal copy

176 M.V. Silnikov et al. / Acta Astronautica 99 (2014) 175–183

Fig. 1. Interaction of the Prandtl–Meyer wave with the proceeding oblique shock: (a) the asymptotic degeneration of the shock inside the expansion fan;
(b) shock interaction with the centered wave; (c) shock strengthening under the compression wave influence; and (d) the case of the subsonic flow
downstream the shock.

perturbations in the region r3. Due to contiguity theorem consider this type of interaction, except of some complemen-
[2], the region r3 is either a rarefaction or compression tary remarks.
Prandtl–Meyer wave. We should emphasize that the wave r3 is partly a result of
Weak tangential (contact) discontinuity τ1 originates the refraction of the perturbations reflected from the shock in
from point B downstream the shock. The discontinuity τ1 is the region BCC 2 on the vortex layer. At this refraction, the
the lower border of the substantially non-isentropic flow perturbations reflected from the non-isentropic layer 5 are
region 5 which can also be named the vortex layer, or the generated in their turn. These reflected perturbations distri-
slipstream of finite width. Another weak tangential dis- bute along the characteristics of the first family, overtake the
continuity τ2 (Fig. 1b and c) originating from the point C of shock situated upstream them, influence its strength and
intersection between the shock and the first straight shape, and make it curvilinear even after point C. So the
characteristic of the wave r1 is the upper border of the shock j4 cannot be formally called straight, and the flow
vortex region. The weak discontinuity CC1 also goes out behind is not exactly uniform and isentropic.
from the point C and encloses the reflected wave r3 if the The main goals of the present study are to define and
flow at point C downstream the shock is really supersonic. analyze the shape and the other features of the interacting
If the Prandtl–Meyer expansion wave is too strong, then shock, to find out the type of the resulted perturbations
the curvilinear shock degenerates into one more weak and their transition criteria, as well as some special
discontinuity inside the expansion fan (Fig. 1a). The dis- features of the considered interaction, and the influence
continuity τ2 does not exist there, and non-interacted part of the ratio of gas specific heats on the problem solution.
r4 of the wave r1 serves instead of the shock j4 which
comes through the expansion fan otherwise. 2. Mathematical models and governing relations
The oblique shock j4 (Fig. 1b–d) or the expansion wave r4
(Fig. 1a) of the same direction as the wave r1 and the shock j2 2.1. Non-uniformities of the two-dimensional gas flow
(direction index χ ¼ 1) as well as the reflected isentropic wave
r3 of the opposite direction (χ ¼ 1) result due to the To characterize not only flow parameters, but also their
interaction. If the flow downstream the shock is subsonic spatial derivatives at the non-uniform stream of a perfect
(Fig. 1d), then the results of interaction sufficiently depend inviscid gas, we introduce “basic non-uniformities” [3]:
upon the perturbation induced downstream. We do not N 1 ¼ ∂ ln p=∂s which expresses the flow non-isobaricity,
Author's personal copy

M.V. Silnikov et al. / Acta Astronautica 99 (2014) 175–183 177

N2 ¼ ∂θ=∂s (the geometric curvature of streamlines), the analogous to (1, 2) are correct for the shock j4
N3 ¼ ∂ ln p0 =∂n which characterizes the flow vorticity (i.e. sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
its non-isentropic features), and N4 ¼ δ=y which expresses the ðJ 4 þ εÞM 20  ð1  εÞðJ 24  1Þ
M4 ¼ ;
type of flow symmetry (N4  0 for the plane flow considered J 4 ð1 þ εJ 4 Þ
here). Here s and n are the natural coordinates counted along "sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi #
and normally to the streamlines, M is the local Mach number; J m ðM 0 Þ J 4 ð1  εÞðJ 4  1Þ
β4 ¼ arctan ; ð3Þ
p is the static pressure, p0 is the flow stagnation pressure of J 4 þ ε J m0 þ ε  ð1  εÞðJ 4  1Þ
the flow, y is the distance to the axis of symmetry in axis-
and also for the voluntary point at the interacting part BC
symmetric flow, δ ¼ 0 and δ ¼ 1 for the plane flow and for the
of the shock considered
axis-symmetric one, correspondingly, and γ is the ratio of gas
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
specific heats (γ ¼ 1:4 everywhere if not specially mentioned).
_ ðJ þ εÞM 2  ð1 εÞðJ 2  1Þ
The geometrical curvature K s of the shock convention- M¼ ;
Jð1 þ εJÞ
ally remarked here as the additional non-uniformity
"sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi #
(N 5  K s ). We admit that the curvature K s 40 for the J m ðMÞ  J ð1  εÞðJ 1Þ
shocks convex downwards (shocks BE in Fig. 1b, and BC β ¼ arctan : ð4Þ
J þε ð1 þ εÞM 2  ð1 εÞðJ 1Þ
in Fig. 1c), and K s o 0 for shocks convex upwards (e.g., see
the shock EC in Fig. 1b, and the shock after the point E in The angle s of shock inclination relative to the flow
Fig. 1a). velocity vector upstream the given point is also bound
with shock strength (J) and local Mach number upstream
(M)
2.2. Dynamic compatibility conditions at stationary shocks
J ¼ ð1 þ εÞM 2 sin 2 s  ε: ð5Þ
Among the numerous variables, we choose the shock In the coordinate system admitted here, Eq. (5) deter-
strength J (i.e. the relation of the static pressures down- mines the dependence of shock slope angle ξ (see Fig. 1a–
stream and upstream it [4]) as the main parameter. We d) on the shock strength
denote that J is the strength of the shock inside the
Prandtl–Meyer fan, J 2 ¼ p2 =p1 is the strength of the shock ξ ¼ s þ βðφÞ
1 ð6Þ
j2, and j4 is the strength of the resulting shock j4. Here p1
and p2 are the pressures in zones 1 and 2. that is an equivalent to the quantity ys ' ðxs Þ where ys ðxs Þ is
Mach numbers M2 (downstream the shock j2) and M1 the shock shape looked for. Here βðφÞ 1 is the local flow angle

(upstream the shock) relate between themselves as it immediately before the considered point of the shock.
follows: Except of J m ðMÞ, we ought to emphasize the strength of
the shock declining the flow to a maximum angle possible
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðJ 2 þ εÞM 21  ð1  εÞðJ 22  1Þ at a given Mach number,
M2 ¼ ð1Þ vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
J 2 ð1 þ εJ 2 Þ u !2
M2  2 u t M2  2
J l ðMÞ ¼ þ þð1 þ 2εÞðM 2 1Þ þ 2;
ðε ¼ ðγ 1Þ=ðγ þ 1ÞÞ. The flow deflection angle β2 also 2 2
depends upon the shock strength:
"sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi # and the strength of the shock reducing the flow velocity
J m ðM 1 Þ J 2 ð1  εÞðJ 2  1Þ downstream it to the critical speed,
β2 ¼ arctan ð2Þ vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

J 2 þ ε J m ðM1 Þ þ ε  ð1 εÞðJ 2  1Þ u !2
M2  1 u M 2
 1
J n ðMÞ ¼ þt þεðM 2 1Þ þ 1:
Here J m ðMÞ ¼ ð1 þ εÞM2  ε is the strength of the normal 2 2
shock in the flow with Mach number M. The relations

Fig. 2. The parametric solution of the interaction problem: (a) general solution; and (b) the region of small Mach numbers.
Author's personal copy

178 M.V. Silnikov et al. / Acta Astronautica 99 (2014) 175–183

Dependencies J n ðMÞ, J l ðMÞ, and J m ðMÞ are given in Fig. 2 Considering the parameters to be used later, Prandtl–
(curves 1, 2, and 3, consequently). The inequality Meyer wave strength is
1 oJ n ðMÞ oJ l ðMÞ o J m ðMÞ is correct for supersonic flow. J 1 ¼ πðM 1 Þ=πðM 0 Þ ð10Þ
To connect the spatial derivatives of the flow para-
2  γ=ðγ  1Þ
meters downstream and upstream the shock, we should where πðMÞ ¼ ð1 þ 0:5ðγ 1ÞM Þ is the isentropic
use first-order (differential) dynamic compatibility condi- pressure function. The Prandtl–Meyer wave deflection
tions. Several forms of that unambiguous conditions have angle is
been published (e.g., in [5,6]; the universal form for strong β1 ¼ χðωðM 0 Þ ωðMÞÞ: ð11Þ
discontinuities in the solution of two-variables quasi-
Here χ ¼ 1 is the wave direction index, and
linear system is given in [7]). qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
To express the correlation between the basic non- ωðMÞ ¼ 1= εarctan εðM 2  1Þ arctan M 2  1 ð12Þ
_
uniformities downstream (N j , j ¼ 1:::3) and upstream (N i ,
i ¼ 1:::5) the shock with its strength J in the flow with is the Prandtl–Meyer function, such as
Mach number M, we use the differential dynamic compat- pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dωðMÞ ð1  εÞ M 2 1
ibility conditions in the following form [3]: ¼ ð13Þ
dM Mð1 þ εðM 2  1ÞÞ
_ 5
N j ¼ C j ∑ Aji N i ð7Þ Analogously, the strength of the reflected wave r3 and
i¼1 flow deflection angle are
where C j and Aji are the factors depending on M, J, shock J 3 ¼ p3 =p2 ¼ πðM 3 Þ=πðM 2 Þ; β3 ¼ χðωðM 3 Þ ωðM 2 ÞÞ; ð14Þ
direction χ, and the flow angle θ (only at axis-symmetric where χ ¼ 1; p2 , M 2 , p3 and M 3 are the pressures and
flow). Relation (7), first published in [3], were successfully Mach numbers in regions 2 and 3 (see Fig. 1b).
used for the analysis of supersonic jet flow out of the plane Later we have to deal with not only the whole waves r 1
[8] and axis-symmetric [9] rocket nozzles. and r 3 but with their parts the shock wave has just
To define from (7) the flow non-uniformities down- interacted. Such is the sector of the wave r 1 bounded by
stream the shock, it is enough to know shock geometrical the characteristics OB and E1 E in Fig. 1b. The flow deflec-
curvature (N 5  K s ) and the non-uniformities upstream it tion angle βð1ϕÞ and the strength J ð1ϕÞ of this part of the wave
(in Prandtl–Meyer flow). are expressed by the formulas
βðφÞ
1 ¼ χðωðMÞ ωðM 1 ÞÞ; J ðφÞ
1 ¼ πðMÞ=πðM 0 Þ: ð15Þ
2.3. Prandtl–Meyer flow non-uniformities and the envelope Here M  MðφÞ is the Mach number at the characteristic
line of its characteristics E1 E determined by the angle φ of its inclination. Similarly,
the parameters of the corresponding sector of the reflected
The two first basic non-uniformities of the Prandtl– wave r 3 after the interaction between the shock and the
Meyer flow are inversely proportional to the distance r to above-mentioned sector of the wave r 1 are determined as
the envelope along the characteristic line:
J ðφÞ ðφÞ
3 ¼ πðM 3 Þ=πðM 2 Þ; βðφÞ ðφÞ
3 ¼ χðωðM 3 Þ ωðM 2 ÞÞ ð16Þ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð1 þ εÞ M2  1 ð1  εÞðM 2 1Þ where M ðφÞ
is the Mach number at the last characteristic of
N1 ¼ ψ ; N2 ¼ χψ ; ð8Þ 3
rM rM 3 the corresponding reflected wave sector.

(here ψ ¼ 1 for the compression wave, and ψ ¼  1 for the


2.4. Oblique shock – preceding overtaking weak
expansion one). Considering the wave r1, where χ  1, M is
discontinuity interaction
the local Mach number, r is the distance from the volun-
tary point C1 at a straight characteristic to the correspond-
Since the basic non-uniformities are equal to zero in
ing point C0 at the envelope line (see Fig. 3a).
zone 0 and correspond to (8) inside Prandtl–Meyer fan, we
The spatial derivative of any variable flow parameter f
! can determine the discontinuities ½N 1  and ½N2  at the
given in the voluntary direction q inside Prandtl–Meyer
! characteristic line O0 C
wave depends on the angle between q and gradient qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
vector gradf which is normal to the straight characteristics ð1 þ εÞ M20  1 ð1 εÞðM20  1Þ
½N 1  ¼ ψ ; ½N 2  ¼ ψ : ð17Þ
! rM 0 rM30
∂f =∂q ¼ q grad f : ð9Þ
On the contrary, the discontinuities of the non-
The relations (8)–(9) and the well-known functions of
uniformities at the last characteristic OB of the Prandtl–
the isentropic flow allow us to calculate all flow non-
Meyer wave are
uniformities before the interacting shock and all para- qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
meters derivatives along the direction of the oblique shock ð1 þεÞ M 21 1 ð1  εÞðM 21  1Þ
coming through Prandtl–Meyer fan of characteristics situ- ½N 1  ¼  ψ ; ½N 2  ¼  ψ : ð18Þ
rM 1 rM 31
ated upstream. For instabnce, the envelope line yΓ ðxΓ Þ,
sometimes degenerating into the sole point (Figs. 1 and It is worthy of notice from (17)–(18) that the relation
3b), sometimes consisting of two separate sections ½N1  χΓ ðM Þ½N 2  ¼ 0 expresses the dynamical conditions of
(Fig. 3c) can be easily derived from the equation y0 ðx0 Þ of compatibility at weak discontinuities
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi of gas flow para-
the streamlined surface [10]. meters (here Γ ðMÞ ¼ γM 2 = M 2  1; χ ¼ 1 for wave r 1 ). It
Author's personal copy

M.V. Silnikov et al. / Acta Astronautica 99 (2014) 175–183 179

Fig. 3. Discriminant curves of Prandtl–Meyer waves: (a) the smooth envelope of the expansion wave; (b) the expansion wave with a centered sector; and
(c) the transition of the expansion wave into the compression one and the corresponding discontinuity of the envelope curve.

can be treated as the specific case of so-called generalized 2.5. Problem solution based on the assumption of zero
Chester–Whitham invariant conservation [11,12]. curvature of the resulting shock
The problem of the interaction of the stationary shock
with the preceding overtaking weak discontinuity has the The calculations carried out by the second-order
following solution [3] in variables accepted here: method of characteristics at utmost asymptotic refining
of the numerical grid have shown that the shock j4 is of an
_
½K s  C 1 ðΓA11 þA12 Þ þC 2 Γ ðΓA21 þ A22 Þ extremely small curvature (in particular, at point C). For
¼ _ ð19Þ
½N 2  C 1 A15 þC 2 Γ A25 example, in case of the interaction with the centered
expansion fan, its curvature is equal to 2 U 10  6 at
_ _
Here Γ  ΓðMÞ, Γ  ΓðMÞ, M is the Mach number upstream M 1 ¼ 1:5, J 1 ¼ 0:8, and J 2 ¼ 1:4; to 3  10  5 at M 1 ¼ 3,
_
the point of the intersection, and M is the Mach number J 1 ¼ 0:4, and J 2 ¼ 4; to 4  10  6 at M 1 ¼ 5, J 1 ¼ 0:3, and
behind the shock at this point (for example, M ¼ M 1 and J 2 ¼ 10. The intersection with the weak discontinuity O0 C
_
M ¼ M 2 at point B). diminishes the shock geometric curvature approximately
Eqs. (17)–(19) allow us not only to define the disconti- by 1200 times in the first case, by 800 times in the second
nuities of shock curvature at points B and C, but also to case, and by 4500 times in the third one. So we can
determine the shock curvature just after the point B quite conclude that the curvature of the shock j4 is very small
exactly (K s ¼ ½K s , (19)). Declaring that the curvature of the compared with the curvature of the interacting shock BC
shock j4 is equal to zero right after the point C, we can and can be accepted to be equal to zero. Then the
determine the shock curvature of the interacting shock BC curvature of the shock BC at point С can be easily
just before this point (K s ¼  ½K s ). But, as it was shown in determined from (17) and (19): K s ¼  ½K s .
Section 1, this assumption is not absolutely exact and The assumption of zero shock curvature at the volun-
needs some proof. tary point of its exit out of the Prandtl–Meyer fan is
Author's personal copy

180 M.V. Silnikov et al. / Acta Astronautica 99 (2014) 175–183

equivalent to the possibility of using the analogous rela- curvilinear shock


tion at all inner points of the interacting shock J ðφÞ ðφÞ ðφÞ ðφÞ ðφÞ ðφÞ
ð23Þ
1 J 2 J 3 ¼ J; β 1 ðMðφÞ; J 1 Þþ β 2 ðM 1 ; J 2 Þ þ β 3 ðM 2 ; J 3 Þ ¼ βðMðφÞ; JÞ:
_
C 1 ðΓA11 þA12 Þ þC 2 Γ ðΓA21 þ A22 Þ Here MðφÞ is Mach number at a straight characteristic
Ks ¼ _ N2 ; ð20Þ
C 1 A15 þC 2 Γ A25 coming into the given point of the shock;
where Γ  Γ ðMÞ, and the quantity N 2 is given by (8). Since J ðφÞ
1 ¼ πðM 1 Þ=πðMðφÞÞ is the strength of the Prandtl–Meyer

K s ¼ ys''=ð1 þ ðys'Þ2 Þ3=2 , where ys ðxs Þ is the shape of the wave part which have just interacted with the shock;
shock, (20) allows us to determine the shape of the shock J ðφÞ ðφÞ
3 ¼ πðM 3 Þ=πðM 2 Þ is the strength of the corresponding

in the interaction region. Eqs. (5–6) binding shock shape part of the reflected wave r 3 ; M ðφÞ
3 is the Mach number at

and strength and (8  16) for flow parameters derivatives the last characteristic of this reflected part; βðφÞ ðφÞ
1 and β3 are

inside the Prandtl–Meyer fan lead to the following shock the flow deflection angles at the named parts of the waves
strength variation versus local Mach number r 1 and r 3 . So the system (23) determines the local shock
strength J and deflection angle β in any inner point of the
dJ=dM ¼ 2ðJ þ εÞf 3 =M; interacting shock. Differentiating (23), we can get the
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffisffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi _ !
ð1  εÞ M 2  1 J m ðMÞ  J C 1 ðΓA11 þ A12 Þ þ C 2 Γ ðΓA21 þ A22 Þ expressions analogous to (20) for the shock curvature,
f 3 ¼ 1þ 1 _ ;
2
1 þ εðM  1Þ J þε M sin ðs  αÞ U ðC 1 A15 þ C 2 Γ A25 Þ and the expressions analogous to (21) for the shock
ð21Þ strength variation.
Condition (22) was introduced in [16] to solve the
where α ¼ arcsinð1=MÞ is the Mach angle. To exclude problem of the interaction between the shock and
trigonometric expressions, Prandtl–Meyer wave of opposite direction (see also [17]).
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðJ þεÞðM 2  1Þ  J m ðMÞ  J 3. Flow analysis, results and discussion
sin ðs αÞ ¼ pffiffiffiffiffiffiffiffiffiffi
M2 1 þ ε
3.1. The accuracy of the methods proposed
Eq. (21) is most convenient to solve the problem
because only local Mach number M and shock strength J
The results achieved due to the “differential” model
participate in this equation. The integral curves of (21) are
(20)–(21) and the “integral” one (22)–(23) demonstrates a
shown in Fig. 2a and b as slim lines. The shock strength
very high accuracy.
decreases under the influence of the rarefaction wave
The errors ΔJ ¼ J J e (J e 's are the correct strengths of the
corresponds to the motion downward these curves; the
interacting shock, and J's are the analytically predicted) in
increase of the strength J at the interaction with the
the proposed models are small enough relatively to both the
compression wave corresponds to the motion upward.
shock strength total variation (J 2  J 4 ) and the pressure
Mach number at sonic line 1 (J ¼ J n ðM Þ there) or on
variation in the reflected wave r 3 . For example, J 3 ¼ 1:014 at
abscissa axis (that corresponds to the shock degeneration)
M 1 ¼ 3, J 1 ¼ 0:4 and J 2 ¼ 4, and the errors of the resulting
can be considered as the only parameter.
shock strength according to the above-mentioned models are
In fact, the assumption of zero curvature of the resulted
ΔJ 4 ¼ 1:8  10  4 and ΔJ 4 ¼ 7  10  5 , correspondingly. The
shock implied the neglect to the influence of the perturba-
model [1] based on the neglect to the reflected waves at all
tions refracted at the entropy layer on the shock para-
(J 3  1) is qualitatively less accurate (ΔJ  0:02C0:4).
meters. These assumptions concerning the moving shock
So, owing to very high order of smallness of the
waves in non-steady flow underlie Chester, Chisnell, and
neglected perturbations reflected from the vortex layer
Whitham's approximate analytical methods [11]. Analo-
and influencing the shock from behind, the solutions
gous assumptions concerning shocks in steady flow were
proposed here are accurate enough to study the type of
introduced in [13–15] and other studies.
the reflected wave r 3 and other special features of the
interaction considered.
2.6. The solution based on the pressure equality and flow
collinearity downstream the resulting waves 3.2. The inflections of the interacting shock

The equality of static pressures and the collinearity of The relation (20) determines the geometric curvature
the flows on both sides of vortex layer 5 is a necessary (not of the interacting shock in an explicit form though this
sufficient, though) condition for the straightness of the formula cannot be applied to the shock with the subsonic
shock j4 . These conditions can be formulated as the system _
flow downstream. In the critical case (J ¼ J n ðMÞ, M ¼ 1) the
binding wave strengths J i and deflection angles βi (i ¼ 1:::4) shock curvature is finite.
J 1 J 2 J 3 ¼ J 4 ; β1 ðM 0 ; J 1 Þ þ β2 ðM 1 ; J 2 Þ þ β3 ðM 2 ; J 3 Þ ¼ β4 ðM 0 ; J 4 Þ: As it follows from (20), rather strong shocks interacting
with the expansion wave are convex downwards (see the
ð22Þ
section BE of the shock in Fig. 1b). The loss of the shock
Shock strengths J 2 and J 4 , corresponding deflection strength under the influence of the expansion wave leads
angles, and Mach numbers are tied together with (1)–(4), to shock inflection, and the curvature becomes negative.
and the same flow variables upstream and downstream The negative curvature of the shock degenerating into a
Prandtl–Meyer waves are bound with (10)–(12). weak discontinuity (Fig. 1a) asymptotically strives to zero.
Like Eq. (19) in the previous model, Eq. (22) can be The sign of shock curvature inside a compression wave
generalized for the inner points of the interacting changes in a reverse order.
Author's personal copy

M.V. Silnikov et al. / Acta Astronautica 99 (2014) 175–183 181

Shock curvature direction changes when shock para- As it is seen in Fig. 2a, one of the integral curves of (21) has
meters cross curve 4 in Fig. 2a. Curve 4 goes out the point r point v of contact with the curve φ2 . To achieve its coordinates
at the “critical” (“sonic”) line 1. The finite limit of the shock (M v ¼ 2:282, J v ¼ 3:434), we have to equate the slopes of the
_
curvature determined from (20) at M-1 (J-J n ðMÞ) leads curve φ2 and the integral curve of (21). The thick integral
to the equations curve 5 of (21) descending from the point (M ¼ 2:670,
J ¼ J n ðMÞ ¼ 6:440) at the “sonic” line 1 to the point
ð1  εÞM 6r  ð5 9εÞM 4r þ 12ð1  2εÞM 2r 16ð1  εÞ ¼ 0; (M ¼ 1:477, J ¼ 1) contacted the curve φ2 at the point v. All
J 3r 2J 2r 3εJ r  ð1  εÞ ¼ 0 the integral curves situated to the right of curve 5 correspond
to at least one reflected wave type change point.
for Mach number Mr ¼ 1:676 and the shock strength _
Analyzing the “differential” model (20)–(21) for N 1 ¼ 0,
J r ¼ J n ðM r Þ ¼ 2:361 at the point r where the inflection line
we transformed it to the same Eq. (24) which was
starts. Inflection line 4 has the horizontal asymptote
achieved earlier concerning wave r3 change of type. The
(J ¼ J A ¼ 6:360) at M-1 where J A is the root of sixth-
coincidence of the solutions subsequent to “differential”
order algebraic equation.
model (for the points situated immediately after the
shock) and “integral” model (for the reflected wave)
3.3. The change of type of the reflected expansion or
reveals that the flow type seems not to change along the
compression wave
second-family reflected characteristics, even in zone 5.
Flow type change in the reflected wave can be illu-
Change of flow type (expansion or compression) in the
strated by the influence of the expansion wave on the
wave r 3 can be analyzed based on (23). The second-family
shock with initially “sonic” strength J 2 ¼ J n ¼ 8:282 at
characteristic where the wave r 3 changes its type corre-
M 1 ¼ 3 (the integral curve KK 1 in Fig. 2). The sector KK 2
sponds to the extremum of current strength J ð3ϕÞ of the
of this curve corresponds to the flow expansion in the
reflected wave part. The solution (wave type change
reflected wave (from M 2 ¼ 1 to M ðφÞ 3 ¼ 1:0208). The flow
criterion) depends only on the local Mach number M and
compression in the reflected wave begins at M ¼ 2:860 and
the shock strength J
J ¼ 6:543 (point K 2 ). Shock inflection point K 3 (M ¼ 2:572,
3 J ¼ 4:258) is situated in this sector. The Mach number in
∑ Di M 2i ¼ 0;
i¼0 the reflected wave diminishes to M ðφÞ 3 ¼ 1:0007 at the end
D3 ¼ J 2 ð1 þ εÞ2  4εðJ þ εÞ2 ; D2 ¼ 4εð1  εÞðJ þ εÞðJ 2 1Þ of the compression sector K 2 K 4 (M ¼ 2:089 and J ¼ 2:016
at point K 4 ). But the flow at the reflected compression
2ð1  ε2 ÞJ 2 ðJ  1Þ  4ð1  2εÞðJ þ εÞ2 ;
wave does not decelerate to the critical speed because the
D1 ¼ ð1  εÞ U ½4ð1 2εÞðJ 2  1ÞðJ þεÞ þ4ðJ þ εÞ2 new expansion sector begins after the point K 4 till the very
þð1  εÞJ 2 ðJ  1Þ2 ; D0 ¼  4ð1  εÞ2 ðJ þεÞðJ 2  1Þ: shock degeneration (J-1 at M ¼ 1:630; point K 1 ). Mach
ð24Þ number in the reflected wave strives to 1:0016, and the
integral strength J 3 of the reflected wave strives to 0:9981.
Eq. (24) determines the curves ϕi (i ¼ 1; 2) at ðM; J Þ-plane
(Fig. 2,a,b). If r 1 is the expansion Prandtl–Meyer flow then 3.4. The contemporary degeneration of the resulting shocks
the regions I and II correspond to the reflection of and waves
compression waves, and the region III to the reflection of
rarefaction waves. The contrary conclusions are correct if Inasmuch as some sectors of the reflected wave r 3
r 1 is the compression wave. realize the expansion flow, and some sectors the compres-
Curves φi emerge from the points F i at the axis J ¼ 1. sion one, they also mutually compensate each other some-
The corresponding Mach numbers (M F 1 ¼ 1:245, M F 2 ¼ times. Supposing both J 3 ¼ 1 and J 4 ¼ 1 in (22), we define
2:540) are determined by the relation the special type of the interaction between the expansion
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi pffiffiffi ffi wave and the oblique shock. These conditions determine
M F 1;2 ¼ 2ð1 7 εÞ=ð1 7 2 εÞ: the curves hi (i ¼ 1; 2) in Fig. 2a and b. When the shocks
These special Mach numbers were discovered earlier in which correspond to these curves degenerate to weak
the problems of overtaking shock–shock and shock – weak discontinuities, they leave mutually annihilating distur-
discontinuity interactions [18] and of the systems which bances in the reflected wave contemporarily.
consist of Prandtl–Meyer waves and subsequent shocks with The curves hi begin at the above-mentioned points Fi.
extreme static pressure at fixed total flow deflection [19]. We can indicate the coordinates of the marginal left point
The curve ϕ1 reaches the “sonic” line 1 at the point s α of the curve h2 (M a ¼ 2:230, J a ¼ 2:081), the point b of its
(M s ¼ 1:305, J s ¼ J n ðM s Þ ¼ 1:466) with the coordinates intersection with the line φ1 (Mb ¼ 2:506, J b ¼ 4:611), the
determined by the equations points d (M d ¼ 1:464, J d ¼ J n ðM d Þ ¼ 1:804) and e (M e ¼
3:013, J e ¼ J n ðM e Þ ¼ 8:358) where the curves hi finish at
4εM 6s þð3 þ 2ε  9ε2 ÞM 4s þ4ð1  8ε þ 6ε2 ÞM 2s 16ð1 εÞ2 ¼ 0; “sonic” line 1.
4εJ 3s þ 3ð1  εÞ2 J 2s ð5  2ε þ 5ε2 ÞJ s þð1  3ε  ε2 ε3 Þ ¼ 0:
3.5. Variation of Mach numbers downstream the shock and
The coordinates (M u ¼ 2:089, J u ¼ 1:989) of the point u subsonic flow pockets
where the curve ðφÞ2 has a vertical tangent are determined
by more complex expressions (for example, J u ðεÞ can be Two contradictory factors influence on the variation
_
presented as a root of an eighth-order algebraic equation). of the Mach number M immediately downstream the
Author's personal copy

182 M.V. Silnikov et al. / Acta Astronautica 99 (2014) 175–183

interacting shock. On the one hand, as this curvilinear M


shock weakens inside the expansion fan, its strength J
diminishes, and that leads to a Mach number increase 3.5
behind the shock. But, on the other hand, the Mach
3 8 7
number upstream the shock decreases, and it leads to a 6
3
decrease in the Mach number downstream. Mach number
10
just downstream the shock increases as a rule at large and
moderate Mach numbers upstream and diminishes at 2.5
5
small Mach numbers.
_
The equation dM=dM ¼ 0 analyzed together with (4) and 2
(21) determines curve 6 (i.e. the line kn in Fig. 2b) of the local 1
Mach numbers maxima just behind the interacting shock. 9 13
1.5
Here M k ¼ 1:127, and the coordinates (M n ¼ 1:257, J n ¼ 4
1:376) are the only real roots of the equations 12 11
2
1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9
ð1  εÞM 6n  ð4 5εÞM4n þ ð7  8εÞM 2n ð5  4εÞ ¼ 0;
Fig. 4. Variation of the specific Mach numbers in the practically impor-
J 3n  ð1 þεÞJ n  1 ¼ 0
tant range of γ‘s. According to the citation order: Mr (curve 1), M F 1 and
A local decrease in the downstream Mach number that M F 2 (curves 2 and 3, correspondingly), Ms (curve 4), Mu (5), Mv (6), Ma (7),
Mb (8), Md (9), Me (10), Mk (11), Mn (12), and Mg (13).
corresponds to the region Okn can lead to the flow
transition downstream the shock through the critical
_
speed (M ¼ 1). It can become earlier than the shock wave reflected wave type change curve φ2 moves sufficiently to
degenerates into a weak discontinuity. So the integral the right.
curves of (21) situated in the area between “sonic” line 1 The curve kn of the maximum downstream Mach
and dotted integral line 7, or Og (M g ¼ 1:391, J g ¼ 1:640) numbers also exists at all 1 oγ o 1 moving a little to the
both start and finish at the “sonic” line (to the right of larger Mach numbers. The curve h2 and the region 3b
point g and to the left of it, correspondingly). Considering corresponding to shock degeneration and critical flow
only the supersonic flow downstream the shock, we can deceleration in the reflected waves also shift to the right.
conclude that the integral curve starting from point n The rules determining the direction of shock curvature
finishes at the same point at once. remain the same for other γ's. Mach numbers and shock
As the shock j2 interacts with the preceding expansion strengths corresponding to the shock inflection increase:
wave r1, the Mach number immediately behind the inter- for example, Mr -1:659 and J r -2:206 at γ-1; Mr ¼ 1:676
acting shock is always larger than at the corresponding pffiffiffi
and J r -2:361 at γ ¼ 1:4; M r - 3 and J r -3 at γ-1. The
characteristic of the reflected wave. So, when the flow inflection curve 2 always has the horizontal asymptote.
immediately after the shock sometimes decelerates to a So the increase in specific heats ratio widens the
critical and even subsonic velocity, it means that the expansion wave reflection region (as r1 is the expansion
supersonic flow in the reflected mainly compression wave wave) and shifts the majority of the special curves and
r3 have decelerated to the critical velocity sooner. points plotted here to the region of larger Mach numbers
According to the model (22), the critical flow decelera- and shock strengths.
tion in the reflected compression wave is possible for The dependence of some specific Mach numbers upon
incoming shocks which correspond to the dashed regions the specific heats ratio in the practically important range
“A” and “B” adjacent to the curves 8a (Fig. 2b) and 8b (1 oγ o 2) is shown in Fig. 4. One can see, for example,
(Fig. 2a) if the incoming expansion wave r1 is strong that some specific small Mach numbers (M k , M n , M s , M g ,
enough. Points d and e at the “sonic” line whose coordi- M d , and M r ) are weakly dependent on γ in the range
nates are calculated in Subsection 3.4 (M d ¼ 1:464, shown there.
M e ¼ 3:013) are the endpoints of “A” and “B” regions, The applications of the presented results can be found
correspondingly. in developing optimal design of supersonic inlets and
nozzles; the analysis of supersonic jet engines and per-
3.6. The influence of the ratio of gas specific heats spective detonation engines because of the sequential
presence of compression and expansion flow regions
To characterize the influence of specific heats ratio on [20]. The available numerical tools need theoretical solu-
the solutions derived here, it seems enough to study the tions serving as verification basis.
variation of the special curves and points plotted above.
So, starting points
pffiffiffi Fi of the curves φi merge at γ-1 4. Conclusions
(MF 1 ¼ MF 2 ¼ 2 then). A distance between these curves
increases with γ. So the region II if Fig. 2, a corresponding Two reliable analytical solutions for overtaking Prandtl–
to the expansion wave reflection at shock interaction with Meyer wave–oblique shock interaction were obtained. These
incoming expansion wave widens then. The first of the solutions allow us to estimate both the flow parameters and
start points Fi moves to smaller Mach numbers region. their spatial derivatives downstream the shock as well as the
So region I, wherein the reflected wave r 3 is of another distinctive features of the interacting shock and reflected
type than incoming wave r 1 , becomes narrower. Another Prandtl–Meyer wave. So the geometrical curvature and the
Author's personal copy

M.V. Silnikov et al. / Acta Astronautica 99 (2014) 175–183 183

inflection points at the interacting shock were determined [5] G. Emanuel, H. Hekiri, Vorticity jump across a shock in nonuniform
analytically. flow, Shock Waves 21 (1) (2011) 71–72.
[6] G. Emanuel, H. Hekiri, Vorticity and its rate of change just down-
A simple expression determining the rarefaction/com- stream of a curved shock, Shock Waves 17 (1–2) (2007) 85–94.
pression flow type change immediately behind the shock [7] A.V. Omelchenko, On the correlation of derivatives at a strong
as well as in the reflected wave were derived also. It was discontinuity, Comput. Math. Math. Phys. 42 (8) (2002) 1246–1257.
proven that the reflected Prandtl–Meyer wave can change (in Russian).
[8] V.N. Uskov, M.V. Chernyshov, Differential characteristics of flow filed
its type, not once somewhere. The criteria of the full in plane overexpanded jet in vicinity of nozzle lip, J. Appl. Mech.
mutual compensation of the different sectors of the Tech. Phys. 47 (3) (2006) 366–376.
reflected wave and the criteria of the shock degeneration [9] M.V. Silnikov, M.V. Chernyshov, V.N. Uskov., Two-dimensional over-
were derived. Non-monotonic variation of the Mach num- expanded jet flow parameters in supersonic nozzle lip vicinity, Acta
Astronaut. 97 (2014) 38–41.
bers just downstream the curvilinear interacting shock [10] V.N. Uskov, M.V. Chernyshov, Conjugation of Prandtl–Meyer wave
was discovered, and the possibility of subsonic flow pock- with quasi-one-dimensional flow region, Math. Model. 15 (6) (2003)
ets was revealed. The influence of gas specific heats 111–119. (in Russian).
[11] G.B. Whitham, Linear and Nonlinear Waves, Wiley Interscience, New
ratio on the deduced solutions was found quantitatively
York, 1974, 656.
remarkable. [12] A.V. Omelchenko., The generalized Chester-Whitham invariant,
The most evident applications of the presented results Techn. Phys. Lett. 27 (11) (2001) 891–893.
are: the optimal design of supersonic inlets and aerody- [13] W.E. Moeckel, Interaction of oblique shock with regions of variable
pressure, entropy, and energy, NACA TN 2725 (1952) 35.
namic bodies for supersonic flight; the analysis of super- [14] A.L. Adrianov., A Model of the Flow Downstream the Curvilinear
sonic jet engines because of the sequential interchange of Stationary Shock, Computational Center of RAS Siberian Branch,
compression and expansion flow regions divided by the Krasnoyarsk, 1997, 12 (in Russian).
shocks in supersonic jet flows. [15] Yu. S. Vasil'ev, B.E. Gelfand, M.V. Silnikov, Impulse-energy technol-
ogies assisted with mitigation systems, Izvestiya Akademii Nauk,
The work was performed with the financial support of Energetika, 2004, 7–19.
the Russian Foundation for Basic Research (Project 13-03- [16] H. Li, G. Ben-Dor, Oblique shock–expansion fan interaction–analy-
00003). tical solution, AIAA J. 43 (2) (1996) 418–421.
[17] V.R. Meshkov, A.V. Omelchenko, V.N. Uskov, The Interaction of the
Stationary Shock with the Expansion Wave of Opposite Direction,
References Vestnik Saint Petersburg University: Mathematics, Saint Petersburg
State University Press, Saint Petersburg, Russia, 2002, 99–106 (in
[1] R. Courant, K.O. Friedrichs, Supersonic Flow and Shock Waves, Wiley Russian).
Interscience, New York, 1948, 464. [18] W.D. Hayes, R.F. Probstein, Hypersonic Flow Theory, Academic Press,
[2] G.G. Cherny., Gas Dynamics, Nauka, Moscow, 1988, 424 (in Russian). New York, 1959, 464.
[3] A.L. Adrianov, A.L. Starykh, V.N. Uskov, Interaction of Stationary Gas- [19] A.V. Omelchenko, V.N. Uskov, An optimal shock-expansion system in
Dynamic Discontinuities, Nauka, Novosibirsk, 1995, 180 (in Russian). a steady gas flow, J. Appl. Mech. Techn. Phys. 38 (3) (1997) 204–210.
[4] N.N. Smirnov, V.F. Nikitin, S. Alyari Shurekhdeli, Investigation of self- [20] N.N. Smirnov, V.F. Nikitin, Modeling and simulation of hydrogen
sustaining waves in metastable systems, J. Propuls. Power 25 (3) combustion in engines, Int. J. Hydrogen Energy 39 (2) (2014)
(2009) 593–608. 1122–1136.

View publication stats

Vous aimerez peut-être aussi