Introduction To Gas Dynamics All Lecture Slides: Autumn 2009
Introduction To Gas Dynamics All Lecture Slides: Autumn 2009
Introduction To Gas Dynamics All Lecture Slides: Autumn 2009
Autumn 2009
1 Compressible flow
2 Zeroth law of thermodynamics
3 First law of thermodynamics
4 Equation of state ideal gas
5 Specific heats
6 The perfect gas
7 Second law of thermodynamics
8 Adiabatic, reversible process
9 The free energy and free enthalpy
10 Entropy and real gas flows
11 One-dimensional gas dynamics
12 Conservation of mass continuity equation
13 Conservation of energy energy equation
14 Reservoir conditions
15 On isentropic flows
Gasdynamics Lecture Slides
16 Eulers equation
17 Momentum equation
18 A review of equations of conservation
19 Isentropic condition
20 Speed of sound Mach number
21 Results from the energy equation
22 The area-velocity relationship
23 On the equations of state
24 Bernoulli equation dynamic pressure
25 Constant area flows
26 Shock relations for perfect gas Part I
27 Shock relations for perfect gas Part II
28 Shock relations for perfect gas Part III
29 The area-velocity relationship revisited
30 Nozzle flow converging nozzle
Gasdynamics Lecture Slides
Compressible flow
Compressible flow
An extreme example of
compressible flow in action is the
re-entry flow. Another is shown
here on the left as a jet fighter
seemingly punches through the
sound barrier. However, more
daily mundane applications can
also be found in flows through jet
Figure 1: Breaking the sound engines, or around a transport
barrier. . . ? aircraft.
dE = Q + W (2)
Thus, for the system considered here, the work term can be written
~ and displacement ~r as:
in terms of force vector F
Z Z Z
W = F ~ d~r = PA dr = P dV (3)
Using specific variables, the first law can now be written in its
differential form as:
1
de = m Q P dv = q P dv (4)
H = E + PV (5)
dP
dh = de + P dv + v dP = q + v dP = q + (6)
Specific heats
Specific heats
e e
de = dv + dT = q P dv (10)
v T
or:
e e
q = dT + +P dv (11)
T v
Hence, the specific heats can be expressed as:
q e
cv = = (12)
dT v T
q e e v
cp = = + +P (13)
dT P T v T P
h h
dh = dP + dT = q + v dP (14)
P T
or:
h h
q = dT + v dP (15)
T P
And now, the specific heats can be expressed as:
q h h P
cv = = + v (16)
dT v T P T v
q h
cp = = (17)
dT P T
e = e (T ) (18)
h = e + Pv = e (T ) + RT = h (T ) (19)
Furthermore, it follows then that the specific heats are also functions
of only temperature:
de
cv = (20)
dT
de v
cp = +P = cv + R (21)
dT T P
Similarly:
dh P
cv = v = cp R (22)
dT T v
dh
cp = (23)
dT
R = cp cv (24)
Z
e (T ) = cv dT (25)
Z
h (T ) = cp dT (26)
cp
= (27)
cv
where is the ratio of specific heats. For diatomic gases, e.g.,
nitrogen and oxygen, which constitute the bulk of the Earths
atmosphere at standard conditions, this value is found to be roughly
1.4.
E
P= (30)
V
Q = T dS (32)
The term adiabatic refers to the idea that no heat passes through the
boundary of the system, Q = 0.
The term reversible indicates that the system can be restored to its
initial state with no changes in both the system itself and its
surroundings, such that no energy is being dissipated. This is
signified as dS = Q/T .
In general, dS Q/T for natural processes. This means that
entropy increase can be attributed to either heat transfer, or to
irreversibilities.
A process that is both adiabatic and reversible is marked by the fact
that the entropy is kept constant, dS = 0.
Though it may sound very restrictive, plenty of important
applications in fluid flow can qualify. Consequently, isentropic
analysis is used rather extensively.
From the first law, the following can be written for isentropic
processes:
e e
dv + dT = P dv (37)
v T
h h
dP + dT = v dP (38)
P T
For constant :
Z
v dT 1 T
ln = = ln (43)
vo ( 1) T 1 To
Z
P dT T
ln = = ln (44)
Po ( 1) T 1 To
1 T
ln = ln (45)
o 1 To
F = E TS (46)
G = H TS (47)
dE = T dS P dV (48)
dF = S dT P dV (49)
dG = S dT + V dP (50)
dH = T dS + V dP (51)
Experiments have shown that for large regions of the flow, these
gradients are small. Consequently, entropy production is very small
and may be neglected in these regions. This, however, do not apply
to regions where large gradients exist, such as across shock waves,
and flows inside the boundary layers, vortex cores and wakes. In
this regions, non-isentropic analysis must be employed.
Figure 4: A streamtube
Note that this definition still holds in the case of slowly varying
cross-sectional area along the axis x, provided that the variation is a
function of the axis A = A (x) only, unchanging in time; and that
flow properties, e.g., P and , are uniform across the cross-section,
including the velocity u = u (x), which is normal to the
cross-sectional area. These flow quantities can also be allowed to
vary with time t in case of non-stationary or unsteady flows.
In a one-dimensional incompressible flow, practically all information
is contained in the kinematic relation uA = constant, while pressure
is obtained independently using the Bernoulli equation.
On the other hand, compressible flows impose inter-dependency
between the conservation of mass and momentum, and the
area-velocity rule is no longer straightforward.
Note that the mass flow rate across any cross-section is m = uA.
Also, the mass enclosed within the segment for small enough x
can be set equal to A x.
Gasdynamics Lecture Slides
Conservation of mass continuity equation
And in the case of steady flow situations, the time dependency can
be dropped, and one is left with:
d
dx (uA) = 0 m = uA = constant (56)
w = P1 v1 P2 v2 (57)
According to the first law, the energy balance, which now includes
the kinetic energy contributions, is expressible as:
h2 + 12 u22 h1 + 12 u12 = q
(59)
Note that q as shown in the figure refers to heat from outside the
walls. Heat generated internally is included in h. Also, the energy
equation derived here relate the equilibrium states at sections 1 and
2 in the figure. It is valid even if there are regions on
non-equilibrium in-between, as long as the flow is at equilibrium at
the sections themselves.
In adiabatic flows, there is no heat transfer allowed, q = 0, and
the equation is refered as the adiabatic energy equation:
h + 12 u 2 = constant (61)
cp dT + u du = 0 (63)
cp T + 12 u 2 = constant (64)
Reservoir conditions
ho0 = ho = h + 12 u 2 (66)
Be aware that this is only for the perfect gas, otherwise only the
constancy of the stagnation enthalpy holds in the adiabatic flows.
Also note that often these stagnation or reservoir conditions are
referred to as the total conditions.
On isentropic flows
Refering again to the figure of the flow between two reservoirs, note
that the first law doesnt forbid the flow to be reversed, i.e., from
right to left. This illustrates the application of the second law:
so0 so 0 (68)
On isentropic flows
0 Po0
s s = R ln (72)
Po
Be fully aware that for stagnation conditions to properly exist, it is
not sufficient that the velocity is zero. It is also necessary that
equilibrium conditions exist, and there are non-negligible gradients
to cause currents.
Eulers equation
Newtons second law states that the rate of change of momentum for
a body is proportional to the force acting on it:
d = dt + dx (74)
t x
Be aware of the two parts involved: the local time derivative and the
convective derivative associated with the spatial changes to the
quantity as the particle travels with the fluid velocity u. Together,
they are called the total or material derivative.
Accordingly, the acceleration suffered by the fluid particle in a
one-dimensional flow can be written as:
Du u dx u u u
= + = +u (76)
Dt t dt x t x
Note that the velocity term occurs twice in the expression.
Gasdynamics Lecture Slides
Eulers equation
At present, it is assumed
that the only forces acting
on the fluid are inviscid in
nature, no viscous
elements are included.
Furthermore, gravitational
effects are also neglected.
Thus, one ends up with
only pressure forces.
Figure 8: Pressure on a fluid particle
in diverging stream tube
From the preceding figure, the pressure forces in the x-direction can
be summed up as:
X A P A
Fx = PA + Pm x P + x A+ x (77)
x x x
Eulers equation
u u 1 P
+u = (80)
t x x
one ends up with the conservation of momentum for inviscid fluids,
often also called as the Eulers equation.
Momentum equation
u (A) + u (uA) = 0 (85)
t x
A (u) + uA (u) = A (P) (86)
t x x
One can then take the sum of the results and apply integration by
parts, which results in the following:
2
dA
(uA) + u A = (PA) + P (87)
t x x dx
Momentum equation
Noting that the time differentiation can be taken out from inside the
integral sign, one now has:
Z 2
2 2
(uA) dx + 2 u2 A2 1 u1 A1
t 1
Z 2 (88)
= (P2 A2 P1 A1 ) + P dA
1
Momentum equation
Figure 10: Summing the forces and fluxes in the control volume
(A) + (uA) = 0
t x
dA
u 2 A =
(uA) + (PA) + P (91)
t x x dx
(eo A) + (ueo A) = (PuA)
t x x
Implicit in the expressions are the assumptions of negligible
transport coefficients, e.g., viscosity and heat conduction; adiabatic
condition at the walls; and that only pressure forces are present.
Isentropic condition
dP
dh + u du = 0 and u du + =0 (92)
Eliminating u du between the two:
dP
dh =0 (93)
Isentropic condition
( + d) (a du) = a a d du = 0 (95)
a2 = dP/d (97)
But that still leaves out the question of how to relate the pressure
changes to the density changes.
Gasdynamics Lecture Slides
Speed of sound Mach number
When combined with the ideal gas equation, the following correct
expression is obtained for the speed of sound in a calorically perfect
ideal gas: s
P p
a= = RT (103)
As mentioned earlier, one can now define Mach number as the ratio
of the flow velocity and the speed of sound Ma = u/a. It is clear
that its value depends on the local flow velocity and the flow
conditions. Generally, Ma is assumed to be a positive value.
Based on the value of Mach number, fluid flows can often be
divided into several (sometimes overlapping) regimes:
sub / super-sonic Mach number is less / greater than 1
transonic Roughly set around 0.8 Ma 1.25
hypersonic Usually said to be Ma > 5, but mostly defined by the
high degree of real gas effects, such as dissociation of
gas molecules
cp cp /cv (cp cv ) R
cp = (cp cv ) = = (104)
cp cv cp /cv 1 1
To /T = (ao /a)2 = 1 + 1
2 Ma2 (106)
And the isentropic relations can also be recast in terms of the Mach
number:
1/(1)
1 2
o / = 1 + 2 Ma (107)
/(1)
1 2
Po /P = 1 + 2 Ma (108)
However, it does not mean that at the throat area the flow is sonic.
It only says that if sonic condition is reached, it is necessarily be at
the throat area.
If sonic condition is not attained, the throat area will simply be the
location for the maximum velocity for a subsonic flow, or the
minimum velocity for a supersonic flow, since du = 0 for Ma 6= 1.
Also, be aware that in the vicinity of the sonic point, Ma 1, the
flow is very sensitive to minute area changes, since the denominator
(1 Ma2 ) becomes very small. This is a major concern in
designing supersonic nozzles and wind tunnels.
(A) + (uA) = 0
t x
dA
u 2 A =
(uA) + (PA) + P (118)
t x x dx
(eo A) + (ueo A) = (PuA)
t x x
An immediate problem can be seen in that there are 4 unknowns
(, u, P, e) and only 3 equations. To close the system, one needs a
state equation.
Thus, one can rewrite it and then integrate to obtain the caloric
equation of state e = e (T , v ):
P
de = cv dT + T P dv
T v
Z Z Z (122)
P
de = cv dT + T P dv
T v
For adiabatic flows, applying the ideal gas law to the energy
equation gives the compressible Bernoulli equation:
1 2
2u + 1 (P/) = 1 (Po /o ) (124)
In fact, as shown before, this applies also for isentropic flows, which
would further allow to be eliminated:
h i
1 2 (1)/
2 u = 1 (Po /o ) 1 (P/Po ) (125)
There are two Cp values that are of special interest: the stagnation
value, and the critical value, defined at the location where the local
flow turns sonic.
The critical Cp is obtainable by setting Ma = 1 as:
P P 2 2 1 2 1
Cp crit = 1 2
= 2
+ Ma 1
2 u Ma +1 +1
(129)
Similarly, setting Ma = 0 gives the stagnation Cp :
Po P 2 1 2 1
Cp stag = 1 2
= 2
1 + Ma 1 (130)
2 u Ma 2
Note that in the third figure, the control volume has been shrank to
just enclose the portion of the shock wave normal to the flow, the
streamline ab, where the one-dimensional flow assumption is
locally valid.
Gasdynamics Lecture Slides
1 u1 = 2 u2 (134)
1 u12 + P1 = 2 u22 + P2 (135)
h1 + 12 u12 = h2 + 21 u22 (136)
These will form the basis for the derivation of the shock relations
for perfect gas in the following section.
Gasdynamics Lecture Slides
Shock relations for perfect gas Part I
21 u12 22 u22
1 1
P2 P1 = = 21 u12 (137)
1 2 1 2
This is the Rayleigh line equation in the P v diagram for a given
pre-shock condition.
Further solving for 1 u1 gives:
1
1 1 1 2
21 u12 = (P2 P1 ) = (P2 P1 ) (138)
1 2 2 1
Eliminate u2 term from the energy equation:
1 2 1 1 u1 2 1 2 2 1
1
h2 h1 = u1 = 1 u1 (139)
2 2 2 2 21 22
P = RT and h = cp T
(141)
cp P P
h= =
R 1
1 u12 1 u12
P2 u2 2
1= 1 = 1 (145)
P1 P1 u1 + 1 P1
The second law then states that the entropy change across the shock
wave can be found from the ratio of the stagnation pressures:
s2 s1 T2 P2 Po2
= ln ln = ln
R 1 T1 P1 Po1
Po2 P2 T2 1
= (148)
Po1 P1 T1
1 " +1
#
P2 1 1 + 1 (P2 /P1 ) 1
= +1
P1 1 + (P2 /P1 )
Note that for adiabatic flows, the stagnation pressure also allows one
to measure the available energy in the flow.
2 2 (153)
1 + Ma 1
Ma22 = 2 2
1 Ma 1 1
This figure shows the entropy change s/R across the shock as a
function of the upstream Mach number Ma1 .
Furthermore, based on
these shock relationships,
an upper bound exists for
the density ratio:
2 +1
lim =
Ma1 1 1
(156)
Put this in the equation of entropy change across the shock wave:
1 " +1
#
s2 s1 P2 1 1 + 1 (P2 /P1 ) 1
= ln +1
R P1
1 + (P2 /P1 )
(158)
1
" #
1
2 1 1
= ln 1 + +1 1 + +1 (1 + )
ln (1 + ) = 12 2 + 13 3 14 4 + (159)
3
2
3(+1)2
2
Ma1 1 = 12 +1
2 (P2 /P1 1)3
Again, it shows that upstream flow must be supersonic. However, a
weak shock produces a nearly isentropic change of state.
Gasdynamics Lecture Slides
For example, note that the gas at rest, the left half of the right
figure, has its stagnation properties equal to its static properties, as
opposed to its counterpart on the left figure.
Gasdynamics Lecture Slides
The area-velocity relationship revisited
uA = u A
! +1
2 2 +1 1
A /o
= M2 = Ma 2 2
1 2
A /o 1+ 2 Ma
(161)
For flows that are entirely subsonic, the term A refers to the
fictitious area where the flow would reach Ma = 1, it does not
necessarily exist. However, if sonic or supersonic conditions are
attained, then it would correspond to the actual throat area.
Be aware that this relationship is valid only for isentropic flows due
to the use of /o relation.
Gasdynamics Lecture Slides
When the flow at the throat reaches sonic speed, the mass flow can
no longer be increased, and the flow is choked.
As the back pressure is lowered even further, a region of supersonic
flow begins to appear, starting from the throat down the diverging
part of the nozzle. This will terminate in a normal shock within the
nozzle, which is located farther downstream as the back pressure
further decreases, until it reaches the exit plane of the nozzle. Ergo,
no isentropic solution is possible for this range of back pressures.
This is due to the fact that the back pressure Pb is set higher than
what the exit pressure Pe would be if the entire nozzle beyond the
throat goes supersonic. Thus, a discontinous pressure increase, a
shock wave, is required to match the exit condition.
When the supersonic flow from the nozzle discharges directly into
the receiver, such that a normal shock wave stands at the exit, the
subsequent subsonic flow can be decelerated isentropically by the
use of a diffuser. Essentially, its a diverging extension of the nozzle.
The aim is to recover as much stagnation pressure as possible in the
receiving chamber, instead of simply obtaining the Pb at the exit.
Another configuration utilizes a long duct ahead of the exit. The
recompression then relies on the interaction of shocks and boundary
layers. For supersonic flows, this surprisingly results in a fairly
efficient and practical method.
Fanno flow refers to the flow through a constant area duct, where
the effect of wall friction w is considered.
However, it is further assumed that the wall friction does no work,
and does not change the total enthalpy level. Thus no heat transfer
from the friction effects.
This model is applicable to flow processes which are very fast
compared to heat transfer mechanisms. Thus, heat transfer effects
are negligible. This assumption can be used to model flows in
relatively short or insulated tubes.
Wall friction will cause both supersonic and subsonic Mach
numbers to approach Mach 1.
It can be shown that for calorically perfect flows the maximum
entropy occurs at Ma = 1.
Due to adiabatic
assumptions,
the total temperature
To = T 1 + 1 2 Ma
2
is constant. This results in the following:
" #
1 2
d Ma2
dT 2 Ma
+ =0 (175)
T 1 + 1
2 Ma
2 Ma 2
As will be shown later, the wall friction causes the flow to either
speed up or slow down toward Mach 1. Thus, sonic condition is
chosen as the reference point, corresponding to the maximum length
of the conduit.
Gasdynamics Lecture Slides
Starting from the equation of the first law for perfect gas:
dT d dT dP
ds = cv R = cp R (180)
T T P
Assuming sonic condition as a reference point, and applying the
continuity and energy equations:
s s 1 T u
= ln + ln
R 1 T u
1 T To T
= ln + ln (181)
1 T To T
1 T 1 +1 T 1 1
= ln + ln ln
1 T 2 2 T 2 2
Other relationships can be derived for the Fanno flow as well. From
the momentum equation:
!
1 1 2 2
dP + 2 Ma d Ma
= 2 1
P 1 + 2 Ma2 Ma2
1
! (182)
1 1 2 2
= 2
+ 1
d Ma
2 Ma 1 + 2 Ma2
Again, assuming sonic condition as a reference point:
1 2
P 1 1 1+ 2 Ma
ln = ln Ma2 ln +1
P 2 2 2
v (183)
u +1
P 1 u 2
= t
P Ma 1 + 12 Ma
2
1 u12 + P1 = 2 u22 + P2
m
P2 P1 = (u2 u1 )
A
(185)
Assume:
Energy balance dictates:
perfect gas
steady flow h2 + 21 u22 = h1 + 12 u12 + q
constant area q = ho2 ho1 (186)
no wall friction = cp (To2 To1 )
Gasdynamics Lecture Slides
Flow with heat addition Rayleigh flow
Combining:
T P 2 Ma2 (1 + )2 Ma2
= = (189)
T P2 2 2
1 + Ma
Gasdynamics Lecture Slides
In addition:
u P /T (1 + ) Ma2
= = = (190)
u P/T 1 + Ma2
" #" #
2 2 1 2
To (1 + ) Ma 1 + 2 Ma
= 2 +1
(191)
To 1 + Ma2
2
" 1 2
#
1
Po 1+ 1+ 2 Ma
= (192)
Po 1 + Ma2 +1
2
It would be instructive to plot the flow on the T s diagram. To do
so, one would need to express P/P in terms of T /T , since:
s s T P
= ln ln (193)
R 1 T P
The following figure shows a Rayleigh line curve. Note that the
second law dictates that the flow is driven toward Mach 1.
T (1 + )2 Ma2
= 2
T 1 + Ma2
2 2
(196)
d (T /T ) (1 + ) 1 Ma
2
= 3 =0
d Ma 2
1 + Ma
mass: (2 u2 ) (1 u1 ) = 0
2 u22 + P2 1 u12 + P1 = 0
momentum: (197)
h2 + 12 u22 h1 + 12 u12 = 0
total enthalpy:
s2 s1 1 T2 2 1 T2 u2
= ln ln = ln + ln (198)
R 1 T1 1 1 T1 u1
For Rayleigh line, the momentum equation, along with ideal gas
law, gives:
2 2
0 = 2 u2 + 2 RT2 1 u1 + 1 RT1
2 1+Ma21 1
0 = (u2 /u1 ) (u2 /u1 ) + Ma
Ma21 2 (T2 /T1 )
1
(200)
r
1+Ma 2 4Ma12
u2 /u1 = 2Ma21 1 1 2 (T2 /T1 )
1 (1+Ma21 )
1
ux2 uy2
Setting ho = h + 2 + , the governing equations for the
steady inviscid 2-D flows considered here are:
(ux ) + (uy ) = 0
x y
ux2 + P +
(uy ux ) = 0
x y
(202)
uy2 + P = 0
(ux uy ) +
x y
(ux ho ) + (uy ho ) = 0
x y
Oblique shocks
From geometry:
u1n = u1 sin
= u1t tan
(203)
u2n = u2 sin ( )
= u2t tan ( )
Oblique shocks
1 u1n = 2 u2n
1 u12n + P1 = 2 u22n + P2
(205)
u1t = u2t
1 2
2 u1n + 1 (P1 /1 ) = 12 u22n + 1 (P2 /2 )
Oblique shocks
Thus, one has the following relationships for the oblique shocks:
u2n 1
+1 2
Ma21n 1
2 2 Ma 1n
= = =1+
1 u1n 1 + 1
2 Ma 2
1 n
1 + 1
2 Ma 1n
2
P2 2
2 1
2
Ma21n 1
= +1 Ma1n 2 = 1 + +1
P1
(210)
T2 h 2 2
1 i h 2 2
1 i
= +1 Ma1n +1 +1 1/Ma1n + +1
T1
1 1
Mat2 ut /a2 a2 T2 2
= = =
Mat1 ut /a1 a1 T1
Expanding:
The following is the chart showing the relationship. Note that there
are a maximum wedge angle max allowed for attached oblique
shock. A greater value would result in detached shock. This
maximum wedge angle divides the solution for into two branches.
Mach lines
From the earlier discussions, it has been mentioned that for the
weak oblique shock, the wave angle corresponding to an
ever-smaller deflection angle approaches a non-zero limit , such
that lim0 = = arcsin (1/Ma1 ).
At this point, the flow essentially experiences no changes
whatsoever. There is nothing unique about the point, it could be
anywhere in the flow. In fact, the angle is simply a characteristic
angle associated with the Mach number Ma1 . Thus, the name Mach
angle, which is the angle of the characteristic line emanating from
some point in the flow, relative to the direction of the flow.
Mach lines
(a) (b)
(c)
This equation shall be used as the starting point for the relationship
for weak oblique shock waves.
For a closer look, one can define the wave angle as a slight
increase from the Mach angle , such that = + for 1.
With this approximation, one can obtain the following expressions:
q
1 + Ma21 1
sin sin + cos =
Ma1
q
Ma21 sin2 1 + 2 Ma21 1 (222)
" #
2
( + 1) Ma1
= 2
4 Ma1 1
Hence, for a small finite deflection angle , the direction of the shock
wave differs from the Mach wave by an amount = O ().
For each wave associated with each , one has the followings:
P = O () and s = O 3 .
du d
= p (228)
u Ma2 1
This equation applies continously through the isentropic turn, and
when integrated, will give a relation between Ma and .
Moving further away from the wall, the Mach lines will ultimately
converge to form a shock wave. This is due to the fact that an
intersection of two Mach lines would imply an infinitely high
gradient, since the Mach number is double-valued at this point.
Physically, however, as the lines approach each other, the gradients
would become large enough to invalidate isentropic assumption.
Substituting 2 = Ma2 1:
Z "
Ma2 1 2 2
#
+1 d
(Ma) =
0 1 + 1
+1
2 1 + 2
Z Ma2 1 " #
1 1
= d (232)
0 1 + 1
+1
2 1 + 2
q q
+1 1 2
= 1 arctan +1 Ma 1
p
arctan Ma2 1
Detached shocks
The common theme between these two series of figures is that the
deflection angle for the detached shock is greater than the value
allowed in the --Ma equation.
Detached shocks
f0
d f (234)
= = g ()
d 1+f2 1+f2
1 +1
g () =
1 2 1 2/Ma2 + 2/Ma2 +
Detached shocks
Detached shocks
The oblique shock wave and the simple isentropic wave provide the
basic tools for analysing many cases in supersonic flows, by patching
together the solutions, of which some are shown in this figure.
Shock-expansion theory
Shock-expansion theory
The shock waves and the expansion waves also interact with each
other. Here in the figure, it is shown that expansion waves intersect
the shock wave, curving and weakening it, and ultimately reducing
it to mere Mach wave in the far-field.
For most part, the reflected waves are very weak, and thus do not
affect the shock-expansion result for the pressure distribution.
Gasdynamics Lecture Slides
Reflection and intersection of oblique shocks
When the two intersecting shocks are from the same family, it
results in the merging of the two, shown in this figure. Note that a
wave of the opposing family is needed to equalize the pressure
across the slipstream.
The other figure shows what occurs when the expansion waves
intersect an oblique shock wave. Along with the shock weakening,
there are also weak reflected waves.
Gasdynamics Lecture Slides
The flow field for a cone in supersonic flow at zero angle of attack is
not as simple as the one for wedge, where a uniform flow region is
found between the shock and the wedge. In this 3-D case, uniform
flow behind the shock would violate the continuity equation.
However, the flow possesses a property that can be exploited for the
analysis of the flow field, that is the cone can be considered to be
semi-infinite.
The solution has been given, first by Busemann, and later by Taylor
and MacColl. The procedure consists of fitting an isentropic conical
flow to the conical shock. The 3-D flow equations are rewritten in
terms of a single conical variable , which is then solved
numerically.
At the shock, the simple oblique shock jump relation can be used,
since locally it always applies to any shock surface. And the flow
behind the shock must be matched with the isentropic conical flow,
which then determines the solution.
Note that behind the shock, the streamlines are curved. And
compared to the flow over a wedge, the compression is lower due to
the 3-D relieving effect. Also, the detachment occurs at a much
lower Mach number.
~u = hU + ux , uy , uz i (236)
Irrotational flow
~ = 12 ~u
k = 21 ijk (uj /xi ) (239)
Irrotational flow
2 :
Expanding the equation for steady inviscid flow, and dividing by a
uy
Ma2 ux uz
1 x + y + z
uy2
h i
ux2 1 ux2 uz2
= Ma2 2
U
+ 2
2ux
U + 2
U
+ 2
U
+ 2
U
ux
x
uy2 uy2
h i
1 ux2 uz2 uy
+ Ma2 2
U
+ 2
2ux
U + 2
U
+ 2
U
+ 2
U y
uy2
h i
uz2 1 ux2 uz2
+ Ma2 2
U
+ 2 + 2ux
U + 2
U
+ 2
U
uz
z 2
U
h i h i
uy uz uy
Ma2 ux ux ux 2 uz
+ U x + x + Ma U 2 z + y
h i h i
uy uy
+ Ma2 U
ux
y + x
2
+ Ma U 2 uz ux uz ux
x + z
h i h i
ux uy uy
+ Ma2 uz
U
ux uz
z + x
2
+ Ma U 2 ux
y + x
(249)
Gasdynamics Lecture Slides
Note that this equation is valid for both subsonic and supersonic, as
well as transonic flow regimes, but not the hypersonic flow regime,
which requires different approximation to be used due to the large
value of Ma .
And if the transonic regime is excluded, a fully-linear equation then
can be used, obtained by neglecting the entire right-hand-side.
Gasdynamics Lecture Slides
~u = 0 (254)
(258)
Using binomial theorem, the following is obtained:
2
2 2
u +u +u 2 2
2 2
u +u +u 2
Cp = U2ux + x U 2y z Ma4 U2ux + x U 2y z +
(259)
Gasdynamics Lecture Slides
Pressure coefficient
For 2-D and planar flows, consistency with the first-order (linear)
perturbation equation can be maintained by retaining only the first
term:
Cp = U2ux (261)
On the other hand, for axi-symmetric flows, or flows over elongated
bodies, it is necessary to include also the transverse velocity terms
uy and uz :
uy2
h i
2ux uz2
Cp = U + 2
U
+ 2
U
(262)
Boundary conditions
For the 2-D case, it is only necessary to keep the first term, leading
to the following boundary condition:
The previous expression is still valid for a planar 3-D case, since it
requires f /z 0. In this case, the boundary condition becomes:
Note that this does not apply to axisymmetric cases, since the
transverse perturbation velocity cannot be expanded in power series
in the neighbourhood of the axis.
Finally, the far-field boundary condition can be set by requiring the
perturbation velocities to die out, or approaching some finite value,
depending on the nature of the problem.
Gasdynamics Lecture Slides
2 2
1 Ma2
x 2
+ y 2
=0 (270)
2 1 2
+ 2 =0 (272)
x 2 m y 2
X00 1 Y00
= 2 = k2 (273)
X m Y
This gives two ordinary differential equations, whose solutions are:
2 1 2
2 =0 (276)
x 2 y 2
This equation is in the form of what is commonly referred to as the
second order wave equation. The solutions are simply the sum of
two arbitrary functions f and g :
(x, y ) = f (x y ) + g (x + y ) (277)