Sunteți pe pagina 1din 90

MAE 570 AeroPropulsion

Normal and Oblique Shock


Waves and Expansion Waves
Shock Waves

The air inlet of some supersonic fighter jet is designed such that a shock
wave at the inlet decelerates the air to subsonic velocities, increasing the
pressure and temperature of the air before it enters the engine.
2
Normal Shock Wave
Normal shock in a Laval nozzle
Fluid crossing a stationary shock front
rises suddenly and irreversibly in
pressure and decreases in velocity.

The Mach number in the nozzle just upstream


(to the left) of the shock wave is about 1.3.
Boundary layers distort the shape of the normal
shock near the walls and lead to flow
separation beneath the shock.

Schlieren image of a normal shock in a Laval


nozzle.

3
Shock Waves
A shock wave is a discontinuity in a (partly) supersonic flow fluid.
Fluid crossing a stationary shock front rises suddenly and irreversibly in pressure
and decreases in velocity.

A detached oblique shock occurs in front of all blunt-


nosed bodies, whether two-dimensional,
axisymmetric, or fully three-dimensional.
A detached oblique shock is seen in front of a
sphere traveling at supersonic speeds.

Shadowgram of a sphere in free flight through air, M = 1.53. 
Flow is subsonic behind the part of the bow wave that is ahead 
of the sphere and over its surface back to about 45°. 
At about 90° the laminar boundary layer separates through an 
Photo by A. C. Charters, as found in  oblique shock wave and quickly becomes turbulent. 
Van Dyke (1982).
Oblique Shock Waves
When the space shuttle travels at supersonic speeds through the
atmosphere, it produces a complicated shock pattern consisting of
inclined shock waves called oblique shocks.
Some portions of an oblique shock are curved, while other portions are
straight.

Schlieren image of a small model of


the space shuttle Orbiter being tested
at Mach 3 in the supersonic wind
tunnel of the Penn State Gas
Dynamics Lab.
Several oblique shocks are seen
in the air surrounding the spacecraft.
Photo by G. S. Settles, Penn State
University. Used by permission.

5
Normal Shock Waves
The shock waves occur in a plane normal to the direction of flow.

For flow through a normal shock, with no direction change, area change, or
work done.
The continuity, momentum, and energy equations are:

Continuity: 1u1 = 2u2


Momentum: p1 – p2 = 1u1(u2 – u1)
Energy: h01 = h02 T01 = T02
s2 – s 1 ≥ 0
Subscripts 1 and 2 indicate initial and final states, respectively.

The flow process through the shock wave is highly irreversible.


Normal Shocks
Momentum Equation

p1  p2   2u22  1u12 u 2 
p
RT
( M  a) 2  p
RT
M  RT  2
 pM 2

p1  p2   2u22  1u12  p2M 22  p1M 12


p1  p1M  p2  p2M
1
2 2
2

p2 1  M 12

p1 1  M 22

Continuity Equation
 2 u1  p2  T1  2 (  1) M 12
     
1 u2  p1  T2  1 2  (  1) M 12
Normal Shock
Static temperature ratio
T02  1 2
T01 = T02 T01  1 2  1 M2
 1 M1 T2 2
T1 2

 1 2
1 M1
T2 2

T1 1    1 M 2
2
2

Static pressure ratio


2 2
T2 p2u2 p2 M 2 a2 p2 M 2 RT2 p2 M 2 T2  p2   M2  p2 T M
          2 1
T1 p1u1 p1M 1a1 p1M 1 RT1 p1M 1 T1  p1   M1  p1 T1 M 2

 1 2
p2 1  M 12 p2
M1 1 M1
  2
p1 1  M 22 p1  1 2
M2 1 M2
2
Normal Shock Wave

Down stream properties in terms of upstream Mach number

(  1) M 12  2
M2 
2M 12    1
Stagnation Pressure ratio

Stagnation Temperatures
p increases
T01 = T02 p0 decreases
u decreases
M decreases
T increases
T0 remains constant
 increases
s increases
Normal Shock
x: upstream conditions, y: downstream properties

Arrows indicate which


vertical axis to use
when reading chart.
Normal Shock Wave
Example
A ramjet engine is an air-breathing propulsion device with essentially no rotating
machinery (no rotating compressor blades, turbines, etc). The basic parts of a
conventional ramjet are sketched in the Figure below. The flow, moving from left to
right, enters the engine where it is compressed and slowed down. Compressed air
then enters the combustor at very low subsonic speed, where it is mixed with a fuel
and burned. The hot gas then expands through a nozzle. The net result is the
production of thrust toward the left in Figure. In this figure the ramjet is in a
supersonic freestream with a detached (normal) shock wave attached to the inlet. (A
detached normal shock wave in front of the inlet of a ramjet in supersonic flow is
not ideal; rather, it is desirable that the flow pass through one or more oblique shock
waves before entering the inlet.) After passing through the shock wave, the flow
from point 1 to point 2, located at the entrance of the combustor, is isentropic.

The ramjet is flying at Mach 2 at a standard altitude


of 10 km, where the air pressure and temperature
are 2.65 x 104 N/m2 and 223.3 K, respectively.
Calculate the air temperature and pressure at point
2 when the Mach number at that point is 0.2.
Total pressure and total temperature of the freestream at M  = 2 can be obtained
using isentropic relations or table/chart.

Stagnation pressure ratio across normal shock

M = 2 => p01/p0 = 0.7209.

At point (1) behind the normal shock wave, the total pressure is

Total temperature is constant across the shock wave:


Flow is (assumed) isentropic between points (1) and (2), hence p0 and T0
are constant between these points.
p0,2 =1.49 x 105 N/m2 T0,2 = 401.9 K
At point 2, where M2 = 0.2, we can calculate static pressure and static
temperature, using isentropic relations or Table in Appendix A.
 /( 1)
p0    1 2 
 1  M 
p  2 

Note: Air pressure and temperature on the order of 1.42 atm and 399 K entering the
combustor are very tolerable conditions for low-speed subsonic combustion.
Properties across Normal Shock

Entropy Change across Normal Shock


Properties across Normal Shock
Total Temperature

Total (Stagnation) temperature is constant across a stationary normal shock


wave. (Note: flow across a shock wave is adiabatic).

Total Pressure

s2 - s1 > 0, hence p0,2 < p01. Total pressure decreases


across a shock and p02/p01 is a function of M only.
Total Properties across a Normal Shock Wave
Normal Shock
Example: Shock Wave in a Converging–Diverging Nozzle (see previous example)
If the air flowing through the converging–diverging nozzle experiences a normal shock
wave at the nozzle exit plane determine the following after the shock: (a) the stagnation
pressure, static pressure, static temperature, and static density; (b) the entropy change
across the shock; (c) the exit velocity; and (d) the mass flow rate through the nozzle.
Assume steady, one-dimensional, and isentropic flow with  = 1.4 from the nozzle inlet
to the shock location.
SOLUTION Air flowing through a converging–diverging nozzle experiences a
normal shock at the exit. The effect of the shock wave on various properties
is to be determined.
Assumptions 1 Air is an ideal gas with constant specific heats at room temperature. 2
Flow through the nozzle is steady, one-dimensional, and isentropic before the shock
occurs. 3 The shock wave occurs at the exit plane.
Properties The constant-pressure specific heat and the specific heat ratio of air are cp !
1.005 kJ/kg · K and  = 1.4. The gas constant of air is 0.287 kJ/kg.K.
Converging–Diverging Nozzles
Solution from previous example
Since the flow is isentropic, stagnation properties are constant. The properties at the
exit plane can also be calculated or obtained by using data from Figure.
M = 2:
 /( 1)
 
pe Te e Ae p  1 
 0.1278  0.5556  0.2300  1.6875  
p0 1  (  1) M 2 
p0 T0 0 A*
 2 

T 1
pe = 0.1278p0 = (0.1278)(1.0 MPa) =0.1278 MPa 
T0 1  (  1) M 2
2
Te = 0.5556T0 = (0.5556)(800 K) = 444.5 K
1 /( 1)
 
e = 0.23000 = (0.2300)(4.355 kg/m3) = 1.002 kg/m3   1 
 
 0 1  (  1) M 2 
Ae = 1.6875A* = (1.6875)(20 cm2) = 33.75 cm2  2 
Normal Shock Wave in a Converging–Diverging Nozzle

(  1) M 12  2 For M1 = 2
M2 
2M 12    1
M 2  0.5774
p02
 0.7209
p01
p2
 4.500
p1
T2
 1.6875
T1
2
 2.6667
1

2 (  1) M 12

1 2  (  1) M 12
For M1 = 2

M 2  0.5774
p02
 0.7209
p01
p2
 4.500
p1
T2
 1.6875
T1
2
 2.6667
1
The fluid properties after the shock (denoted by subscript 2) are related to
those before the shock.

For M1 = 2
p02 p2 T 
M 2  0.5774  0.7209  4.500 2  1.6875 2  2.6667
p01 p1 T1 1
p02 = 0.7209p01 = 0.7207(1 MPa) = 0.721 MPa
p2 = 4.5000p1 = 4.5000(0.1278 MPa) = 0.575 MPa
T2 = 1.6875T1 = 1.6875(444.5 K) =750 K
2 = 2.66671 = 2.6667(1.002 kg/m3) = 2.67 kg/m3

s2 – s1 = (1.005 kJ/kg.K) ln (1.6875) – (0.287 kJ/kg.K) ln (4.5000) = 0.0942 kJ/kg.K

The air velocity after the shock is determined from:


u 2  M 2  RT2  (0.5774) 1.4(287 J / kg .K )(750.1 K )  317 m / s
24
Properties across Normal Shock
Total Temperature

Total (Stagnation) temperature is constant across a stationary normal shock


wave. (Note: flow across a shock wave is adiabatic).

Total Pressure

s2 - s1 > 0, hence p0,2 < p01. Total pressure decreases


across a shock and p02/p01 is a function of M only.
Pitot-Static Tube
Pitot-static tubes frequently use (pressure) transducers to
measure the difference between the stagnation and static
pressure (ps - p∞).
In the Pitot-static probe, a pressure
transducer measuring ps-p∞ is connected
to an electronic gage that converts this
measurement into the free-stream
velocity, V.

Pitot-static probe
Pitot-static probe, showing the
stagnation pressure hole and two
of five static circumferential pressure
holes.
fig_03_e06a
Measuring Aircraft Speed
Pitot-static probe
A common application of the Pitot-static tube is to measure
the fight speed of an aircraft relative to the air in which the
aircraft is flying.
The Pitot-static tube is
commonly mounted on the
side of the aircraft near the
nose.
In some small aircraft, it is
mounted on the underside of
the wing (e.g., Cessna 172).

Care is needed to obtain the static and stagnation pressure values accurately.
Pitot-Static Tube

Measurement of Flight Speed


(a) Pitot-static tube on the side of a F-15A Eagle fighter aircraft,
(b) a triad of Pitot-static tubes on the side of a B-1b Lancer
bomber.

Pitot-static tubes on military aircraft Credit: Dirk P. Yamamoto.


Measurement of Velocity in Compressible Flow

Subsonic Flow (No shock before Pitot tube)


Measurement of Velocity in Compressible Flow

Supersonic Flow (Normal Shock Wave Present)


The streamline cde crosses the bow shock. Fluid
element moving along streamline cde will first be
decelerated nonisentropically to a subsonic velocity
at point d just behind the shock.
Fluid is then compressed isentropically to zero
velocity at point e. Pressure at point e is not the
total pressure of the freestream but rather the total
pressure behind a normal shock wave, p0,2.
Pressure p0,2 is the Pitot pressure read at the end of the tube.
Because of entropy increase across the shock, there is a loss in total pressure
across the shock, p0,2 < p0,1.
Measurement of Velocity in Compressible Flow

Rayleigh Pitot Tube Formula:


Normal Shock Waves
Measurement of velocity in compressible flow

Example
Air flowing at a Mach number of 2.1 encounters a stationary probe within the flow field
such that a normal shock is generated on the stagnation streamline. The static
temperature and static pressure upstream of the normal shock are 5◦C and 90 kPa,
respectively. Determine (a) the stagnation temperature and pressure upstream of the
shock, (b) the Mach speed, static pressure and static temperature immediately
downstream of the shock. (c) the temperature and pressure at the stagnation point on the
probe.

How would the temperature and pressure downstream of the shock compare with the
temperature and pressure at the stagnation point on the probe? What is the change in
entropy caused by the occurrence of the normal shock?
Measurement of velocity in compressible flow
M1 = 2.1, T1 = 5oC = 278 K, p1 = 90 kPa.

M2 = 0.5613

p2 = 448 kPa

T2 = 492 K

Therefore, the Mach speed, pressure and temperature immediately downstream


of the shock are 0.561, 448 kPa, and 492 K (= 219 oC).
Stagnation Properties
The stagnation pressure and temperature before the shock, (p01 and
T01) and the stagnation pressure and temperature (p02 and T02) after the
shock can be determined as follows

 /( 1)
p01    1 2  p01 = 823 kPa
 1  M1  p1 = 90 kPa, M1 = 2.1
p1  2 

T0,1  1 2
 1 M1 T1 = 278, M1 = 2.1 T01 = 523 K
T1 2
T02 =T01 = 523 K

p01 = 823 kPa, M1 = 2.1


p02 = 554.9 kPa
Therefore the temperature and pressure at the stagnation point on the (stationary) probe are
523 K (= 250oC) and 555 kPa, respectively, which are both higher than the temperature and
pressure just downstream of the normal shoch (492 K and 448 kPa).
Measurement of velocity in compressible flow

Entropy Change
The change in entropy on the stagnation streamline across the normal
shock is given by

p01 823
s2  s1  R ln  (287.1 J / kg.K ) ln  113 J / kg.K
p02 554.9

The flow downstream of the shock (between shock and probe) is assumed to
be isentropic, the change in entropy on the stagnation streamline between a
point upstream of the shock and the stagnation point is 113 J/kg.K.
Oblique Shock Waves
When the space shuttle travels at supersonic speeds through the
atmosphere, it produces a complicated shock pattern consisting of
inclined shock waves called oblique shocks.
Some portions of an oblique shock are curved, while other portions are
straight.

Schlieren image of a small model of


the space shuttle Orbiter being tested
at Mach 3 in the supersonic wind
tunnel of the Penn State Gas
Dynamics Lab.
Several oblique shocks are seen
in the air surrounding the spacecraft.
Photo by G. S. Settles, Penn State
University. Used by permission.

36
Oblique Shocks
A shock wave can form at an oblique angle to the oncoming supersonic
stream. Such a wave will deflect the stream through an angle , unlike the
normal-shock wave, for which the downstream flow is in the same direction.

Consider straight oblique shocks, produced


when a uniform supersonic flow (M1 > 1)
impinges on a slender, two-dimensional
wedge of half-angle .

At that point, since the fluid cannot flow


through the wedge, it turns suddenly
through an angle called the deflection
angle . Turning angle  = 

Information about the presence of the wedge cannot travel upstream in a supersonic flow,
the fluid “knows” nothing about the wedge until it hits the nose.
Oblique Shocks

Since the Reynolds number of supersonic flows is typically large, the boundary
layer growing along the wedge is very thin, and we ignore its effects.

The flow therefore turns by the same angle as the wedge:

=

If we take into account the displacement thickness effect of the boundary layer,
the deflection angle  turns out to be slightly greater than wedge half-angle .
An oblique shock of shock angle   Velocity vectors through an
formed by a slender, two­dimensional  oblique shock of shock angle 
wedge of half­angle . The flow is turned  and deflection angle .
by deflection angle  downstream of the 
shock, and the Mach number decreases.

Unlike normal shocks, in which the downstream Mach number is always


subsonic, M2 downstream of an oblique shock can be subsonic, sonic, or
supersonic, depending on the upstream Mach number M1 and the turning angle.
Oblique Shock Waves
If a plane shock is inclined at an angle to the flow, the fluid passing through
suffers not only a sudden rise in pressure and decrease in speed but also a
sudden change of direction.
The Figure below illustrates an oblique shock s-s in one-dimensional flow.

In passing through the shock, the fluid is deflected


Oblique Shocks
If a plane shock is inclined at an angle to the flow, the fluid passing through
suffers not only a sudden rise in pressure and decrease in speed but also a
sudden change of direction.
The Figure below illustrates an oblique shock s-s in one-dimensional flow.

In passing through the


shock, the fluid is
deflected

Momentum normal to the shock


Continuity: 1u1n = 2u2n

Energy: T01 = T02


Momentum parallel to the shock
Continuity: Energy:
1u1n =2u2n T01 = T02

Momentum normal to shock Momentum parallel to shock

p1  p2   2u 22n  1u12n 0  1u1n (u2t  u1t )


 u 2t  u1t
Velocity parallel to the shock is 
the same on both sides of it. 

An oblique shock becomes a normal shock relative to a coordinate system


moving with velocity u1t = u2t. This fact permits us to use the normal shock
equations to calculate oblique shocks.
M 1n  M 1 sin 

M 2 n  M 2 sin(   )

Like normal shocks, the Mach number decreases across an oblique shock,
and oblique shocks are possible only if the upstream flow is supersonic.

Unlike normal shocks, in which the downstream Mach number is always


subsonic, M2 downstream of an oblique shock can be subsonic, sonic, or
supersonic, depending on the upstream Mach number M1 and the turning angle.

All the equations, shock tables, etc., for normal shocks apply to oblique shocks
as well, provided that we use only the normal components of the Mach
Recall Equations for Normal Shock

(  1) M 12  2
M2 
2M 12    1

2 (  1) M 12

1 2  (  1) M 12
Oblique Shock

(  1) M  2
2 2 (  1) M 12n
M 2n  1n 
2M 12n    1 1 2  (  1) M 12n

p2 2M 12n    1

p1  1

T2 2 M 1n    1
2
 [2  (  1) M 1n ]
2

T1 (  1) 2 M 12n

 /( 1) 1 /( 1) T01 = T02


p02  (  1) M 12n   (  1) 
   
p01  2  (  1) M 12n   2M 2
1n    1 
Given the up-stream Mach number M1 and the wave angle , we could calculate
M1n = M1sin 
and subsequently calculate M2nusing:

(  1) M 12n  2
M 2n 
2M 12n    1

The condition u2t = u1t, can be expressed: M2ta2 = M1ta1

M 2t  M 1t T1 / T2

We note that

2 2M 1n    1
2 Thus, the downstream Mach
T2
 [2  (  1) M 1n ] number M2 and the deflection
T1 (  1) 2 M 12n angle  can be determined.
Relation between M1,  and 

  1 2 2 
21  M 1 sin  
2
tan(   )   
(  1) M 12 sin  cos 

 M 12 sin 2   1 
tan   2 cot   2 
M
 1 (  cos 2 )  2 
Development of Relation between M1,  and 

u2 n 1 tan(   ) 2  (  1) M 12n 2  (  1) M 12 sin 2 


   
u1n  2 tan  (  1) M 1n
2
(  1) M 12 sin 2 

2  (  1) M 12 sin 2  2  (  1) M 12 sin 2 
tan(   )  tan  
(  1) M 1 sin 
2 2
(  1) M 12 sin  cos 

  1 2 2  In a typical application, we know


21  M 1 sin   either  or , but not both.
2
tan(   )   
Fig. 3.10 (next slide) is a plot of this equation
(  1) M 12 sin  cos  for  = 1.4.

The relation for M1,  and  can also be written


as:
 M 12 sin 2   1 
tan   2 cot   2 
M
 1 (  cos 2 )  2 
FIGURE 3.11 Stagnation pressure ratio versus inlet Mach number, with turning angle as parameter.
Curves above dashed line hold for M2> 1, and curves below hold for M2 < 1. (From Shapiro [1].)
FIGURE 3.12 Exit Mach number
versus inlet Mach number, with
turning angle as parameter [1].
Curves above dashed line
correspond to small  and curves
below correspond to large 
(From Shapiro [1].)
Oblique Shocks
Example: 
Supersonic air at M1 = 2.0 and 75.0 kPa impinges on a two­dimensional wedge of half­
angle  = 10°. Calculate the two possible oblique shock angles, weak and strong, that 
could be formed by this wedge. For each case, calculate the pressure and Mach number 
downstream of the oblique shock, compare, and discuss.
Assumptions 1 The flow is steady. 2 The boundary layer on the wedge is very thin.

  1 2 2 
21  M 1 sin  
2  M 12 sin 2   1 
tan(   )    OR tan   2 cot   2 
(  1) M 1 sin  cos   M 1 (  cos 2 )  2 
2
Because The boundary layer on the wedge is very thin
(assumption 2), we approximate the oblique shock deflection
angle  to be equal to the wedge half-angle, : M1
 =  = 10°.
With M1 = 2.0 and  = 10°, we solve the following equation
for the two possible values of oblique shock angle :

2 cot  ( M 12 sin 2   1)
tan  
M 12 (  cos 2 )  2
M1

2 cot  (2 2 sin 2   1)
tan 10  2
o

2 (1.4  cos 2 )  2

Two possible oblique shock


weak = 39.3° and strong = 83.7°. angles formed by a two-
dimensional wedge of half-angle
 = 10°.
From these values, we use the following to calculate upstream
normal Mach number M1n,

M 1n  M 1 sin  M 2 n  M 2 sin(   )

Weak Shock:

M 1n  M 1 sin   2.0 sin 39.3o  1.267

Strong Shock:

M 1n  M 1 sin   2.0 sin 83.7 o  1.988


Oblique Shocks
We substitute these values of M1n into the following equation to calculate the
downstream normal Mach number M2n.

(  1) M 12n  2 Weak shock, M2n = 0.8032, and


M 2n 
2M 12n    1 Strong shock, M2n = 0.5794.

We also calculate the downstream pressure for each case, using


2(1.4)(1.267 2 )  1.4  1
p2  (75 kPa)  128 kPa
p2 2M 12n    1 1.4  1

p1  1 p2  (75 kPa)
2(1.4)(1.9882 )  1.4  1
 333 kPa
1.4  1
Downstream Mach number
M 2n 0.8032
Weak Shock: M2    1.64
M 2n sin(   ) sin( 39.3  10)
M2 
sin(   ) Strong Shock: M2 
M 2n

0.5794
 0.604
sin(   ) sin(83.7  10)

The changes in Mach number and pressure across the strong shock are much greater than the changes across the weak shock.
Comments
For both the weak and strong oblique shock cases, M1n is supersonic and M2n is
subsonic.

However, M2 is supersonic across the weak oblique shock, but subsonic across
the strong oblique shock.

We could also use the normal shock tables in place of the equations, but with
loss of precision.
Oblique Waves
Example
Air flowing with a Mach number of 2.5 with a pressure of 60 kPa and a
temperature of -20oC passes over a wedge which turns the flow through an angle
of 4o leading to the generation of an oblique chock wave. This oblique shock wave
impinges on a flat wall, which is parallel to the flow upstream of the wedge, and
is “reflected” from it. Find the pressure and velocity behind the reflected shock
wave.
Upstream of the wave:

The condition downstream of the initial wave (region 2) are M1 = 2.5 and  = 4o.

Using oblique shock chart:  = 26.6o.

Using normal shock relations/chart:


Conditions in region 3 are determined by conditions in region 2:
2.334

This gives:

Using normal shock relations:

After the reflection, the pressure is 104 kPa and the velocity is 749 m/s.
Detached Shock Wave
For any value of Mach number M > 1, the possible values of  range
from  = 0° at some value of  between 0 and 90°, to a maximum value
 =  max at an intermediate value of , and then back to  = 0° at  = 90°.

Straight oblique shocks for  or  outside of this range cannot exist.


Example: At M = 1.5, straight oblique shocks cannot exist in air with shock
angle  < ~42°, nor with deflection angle  greater than about 12°.

If the wedge half-angle is greater than max,


the shock becomes curved and detaches M1
from the nose of the wedge, forming what
is called a detached oblique shock or a
bow wave.

A detached oblique shock occurs upstream of a two­
dimensional wedge of half­angle  when  is greater 
than the maximum possible deflection angle . 
Detached Oblique Shock
When supersonic flow impinges on a blunt body—a body without a
sharply pointed nose, the wedge half-angle  at the nose is 90°, and an
attached oblique shock cannot exist, regardless of Mach number.

A detached oblique shock occurs in front of all blunt-nosed bodies,


whether two-dimensional, axisymmetric, or fully three-dimensional. For
example, a detached oblique shock is seen in front of the space shuttle
model in and in front of a sphere.

Shadowgram of a diameter sphere in free flight through air at 
M = 1.53. The flow is subsonic behind the part of the bow 
wave that is ahead of the sphere and over its surface back to 
about 45°. At about 90° the laminar boundary layer separates 
through an oblique shock wave and quickly becomes turbulent. 

The fluctuating wake generates a system of weak disturbances 
that merge into the second “recompression” shock wave.
Photo by A. C. Charters, as found in Van Dyke (1982).
Atmospheric Entry

A blunt shape is the most effective


heat shield.
detachment of an oblique shock

Still frames from schlieren videography illustrating the detachment of an oblique


shock from a cone with increasing cone half-angle  in air at Mach 3. At (a)  =
20° and (b)  = 40°, the oblique shock remains attached, but by (c)  = 60°, the
oblique shock has detached, forming a bow wave.
Photos by G. S. Settles, Penn State University. Used by permission.

65
Airplane flow patterns as speed increases.

A supersonic airplane does not necessarily generate an oblique shock that is


attached to its nose—we may instead have a detached normal shock ahead
of the airplane!
As an airplane accelerates to its supersonic cruising speed the flow will
progress from subsonic, through supersonic with a detached normal shock, to
attached oblique shocks that become increasingly “pressed” against the
airplane’s surface.
Supersonic Airfoils

In contrast to subsonic flow designs, these airfoils must have sharp leading
edges, which form attached oblique shocks or expansion fans.

Rounded supersonic leading edges would cause detached bow shocks, greatly
increasing drag and lowering lift.
Example
A jet aircraft inlet operates at 50,000 ft at M0 = 2.5. If the two-shock inlet has a
wedge angle of 15 degrees such that an oblique shock is established on the
leading edge and normal shock occurs at the inlet as shown below find the air
pressure and temperature behind the normal shock.
Assumptions: Here we assume the flow is uniform with a negligible bounday layer.
Given the geometry we can assume the flow must conform to the wedge shown
after the oblique shock and then pass through the normal shock. The procedures
is similar to the problem above.

Solution: The basic sequence here is to:


1. Determine the shock angle of the oblique shock. a. Using the oblique shock
chart - or - relationships we get 1
2. Use the shock angle to find the Mach number normal to the oblique shock
3. Use the normal shock relations/tables to determine the outlet Mach number
normal to the shock and the pressure (and temperature) ratio(s).
4. Use appropriate relations to find the outlet Mach number
5. For the second shock we only need the normal shock relations .
Oblique Shocks
M1 = 2.5;  = 15o; Solve for ; Guess  = 40o

 M 12 sin 2   1 
tan   2 cot   2 
M
 1 (  cos 2 )  2 

Refined guess:  = 36.95o From chart:  ~ 37 deg.

M 1n  M 1 sin   2.5 sin 37  1.503

(  1) M 12n  2 (1.4  1)1.5032  2


M 2n    0. 7
2M 1n    1
2
2(1.4)1.503  1.4  1
2
p2 2M 12n    1 2(1.4)1.5032  1.4  1
   2.468
p1  1 1 .4  1

 /( 1) 1 /( 1)


p02  (  1) M 12n   (  1) 
     0.929
p01  2  (  1) M 12n   2M 1n    1
2

2 2M 1n    1
2
T2
 [2  (  1) M 1n ]  1.322
T1 (  1) M 1n
2 2
Mach number for flow along the wedge:
M 2n 0.7
M2    1.873
sin(   ) sin( 36.95  15)

Finally, the flow must pass through a normal shock wave before
entering the engine inlet:

Recall, for a normal shock (y denotes downstream and x denotes upstrean):

2 2
M x2 
M 22  2
 1 1.8732 
M   1
2
y
2
M x2  1 M3   1.4  1  0.601
 1 2 2(1.4)
M 22  1 1.8732  1
 1 1.4  1

Static pressure ratio across normal shock:

py 2  1 p3 2   1 2(1.4) 1.4  1
 M x2   M 22   1.873   3.928
px   1  1 p2   1   1 1.4  1 1.4  1
Flow across a normal shock:
 1 2  1 2
Ty 1 Mx 1 M2
2 T3 2
   1.587
Tx 1    1 M 2 T2 1    1
M 32
y
2 2
 /( 1)
  1 2 
p0 y py  1 2 M y 
  
px  1    1
p0 x
 M x2 
 2 

 /( 1)
  1 2 
1
p03 p3 
M3 
  2   0.78
p02 p2  1    1
 M 22 
 2 
Final static temperature and pressure are:

T3 T2 p3 p2
p3  p1  2355.51 lb/ft 2
T3  T1  822.6 R p2 p1
T2 T1
p3  16.358 psi

p03 p03 p02   1 2 


 /( 1)

  0.724 p01  p1 1 
 2
M1 

 28.833 psi
p01 p02 p01

The actual total pressure loss p03  p01  0.724  20.879 psi
p03
 e s / R
p01

p03
s   R ln  0.022 Btu/(lb.R)
p01
Oblique Shocks
Example
Consider a Mach 3 flow. It is desired to slow this flow to a subsonic speed. Consider
two separate ways of achieving this: (1) the Mach 3 flow is slowed by passing directly
through a normal shock wave; (2) the Mach 3 flow first passes through an oblique
shock with a 40o wave angle, and then subsequently through a normal shock. These two
cases are sketched in Figure 9.14. Calculate the ratio of the final total pressure values
for the two cases, that is, the total pressure behind the normal shock for case 2 divided
by the total pressure behind the normal shock for case 1. Comment on the significance
of the result.
Prandtl–Meyer Expansion Waves
Review: Oblique shock waves occur when supersonic flow is turned
into itself.
Prandtl-Meyer Expansion Wave
Oblique shock waves occur when supersonic flow is turned into itself.
Expansion waves occur when supersonic flow is turned away from itself.

The expansion fan is continuous expansion region that can be visualized


as an infinite number of Mach waves, each making the Mach angle  with
the local flow direction.

The expansion fan is bounded upstream by a


Mach wave which makes angle 1 sin-1(1/M1)
with upstream flow, and bounded downstream
by another Mach wave which makes angle 2 (=
sin-1 (1/M1) with respect to downstream flow.

Goal:
Develop to calculate the changes in flow properties across expansion waves.
Prandtl–Meyer Expansion Waves
The flow does not turn suddenly, as through a shock, but gradually—each
successive Mach wave turns the flow by an infinitesimal amount.

Since each individual expansion wave is isentropic, the flow across the entire
expansion fan is also isentropic.

The Mach number downstream of the expansion increases (M2 > M1), while
pressure, density, and temperature decrease, just as they do in the supersonic
(expanding) portion of a converging–diverging nozzle.

Prandtl–Meyer expansion waves are inclined at


M2 the local Mach angle .

M1 > 1 The Mach angle of the first expansion wave is

 1   1 
1  sin 
1
  2  sin 
1

 M1   M2 
Prandtl–Meyer Expansion Waves
Consider a situations where supersonic flow is
turned in the opposite direction, such as in the
upper portion of a two-dimensional wedge at an
angle of attack greater than its half-angle .

We refer to this type of flow as an expanding


flow, whereas a flow that produces an oblique
shock may be called a compressing flow.

As previously, the flow changes direction to


conserve mass. However, unlike a compressing
flow, an expanding flow does not result in a
shock wave.

Rather, a continuous expanding region called


an expansion fan appears, composed of an
infinite number of Mach waves called Prandtl–
Meyer expansion waves.

80
Prandtl–Meyer expansion waves
The flow does not turn suddenly, as through a shock, but gradually—each
successive Mach wave turns the flow by an infinitesimal amount.

Since each individual expansion wave is isentropic, the flow across the entire
expansion fan is also isentropic.

The Mach number downstream of the expansion increases (M2 > M1), while
pressure, density, and temperature decrease, just as they
do in the supersonic (expanding) portion of a converging–diverging nozzle.

Prandtl–Meyer expansion waves are inclined at


the local Mach angle .

The Mach angle of the first expansion wave is

 1   1 
1  sin 
1
  2  sin 
1

 M1   M2 
Prandtl–Meyer expansion waves
Determination of M2

The turning angle  across the expansion fan can be calculated by integration,
making use of the isentropic flow relationships.

For an ideal gas:

Turning angle  across an expansion fan

  v( M 2 )  v( M 1 )

where v(M) is an angle called the Prandtl–Meyer function

v( M ) 
  1 1    1 2
 1
tan 


( M  1)   tan 1 M 2  1 
  1 
Expansion Wave
Example
Supersonic air at M1 = 2.0 and 230 kPa flows parallel to a flat wall that
suddenly expands by  = 10°. Ignoring any effects caused by the boundary
layer along the wall, calculate downstream Mach number M2 and pressure p2.
Assumptions: Flow is steady; Boundary layer on the wall is very thin.

The total deflection angle is equal to the wall expansion angle:


 =  =10o

v( M ) 
  1 1    1 2
 1
tan 


( M  1)   tan 1 M 2  1 
  1 

v ( M 1  2) 
1. 4  1
1. 4  1
 1.4  1
tan 1 

 
(2.0 2  1)   tan 1 2.0 2  1  26.38o
 1.4  1 

  v( M 2 )  v( M 1 ) 10o  v( M 2 )  26.38o  v( M 2 )  36.38o

v( M 2 )  36.38o

36.38 
  1 1    1 2
o

 1
tan 

( M 2  1)   tan 1 M 22  1  
  1 

Solving for M2:

M2 = 2.385
We use the isentropic relations to calculate the downstream pressure

 /( 1)
p0    1 2 
 1 M 
p  2 

 /( 1)
  1 2 
 p 2  p0 1  2 M 1 
p2    p1  p
 /( 1) 1
 p0  p1   1 2 
1  2 M 2  Since this is an expansion, Mach
number increases, pressure decreases.

[1  (0.2)2.0 2 ]3.5
p2  (230 kPa)  126 kPa
[1  (0.2)2.3852 ]3.5

We could solve for downstream temperature, density, etc., using the isentropic relations.
Expansion Wave
Example: Expansion Wave on a Supersonic Airfoil

An airplane travels at a speed of 600 m/s in air at 4°C and 100 kPa. The
airplane airfoil has a sharp leading edge with included angle  = 6° and an angle
of attack  = 6°. Find the pressures on the upper and lower surfaces of the
airfoil immediately after the leading edge.

Find: Pressures on upper and lower surfaces.


(a) Upper surface—isentropic expansion

a  RT 

u 600m / s
M1    1.80
a 334m / s

For M1 = 1.80, the Prandtl-Meyer function 1 is obtained from

v( M ) 
  1 1    1 2
 1
tan 


( M  1)   tan 1 M 2  1 
  1 

v1 ( M  1.80) 
1.4  1
1.4  1
 1.4  1
tan 1 

 
(1.80 2  1)   tan 1 1.80 2  1  20.7 o
 1.4  1 
The Prandtl-Meyer function value on the upper surface, is then

vu - v1 =   vu= v1 + u = 20.7° + 3° = 23.7°

vu ( M ) 
  1 1    1 2
 1
tan 


( M 2,upper  1)   tan 1 M 22,upper  1 
  1 

Solving (example, Excel’s Goal Seek function) M2,upper = 1.9

To find p2,upper we note that for isentropic process p0 is constant


 /( 1)
p0    1 2 
 /( 1)  p2,upper    1 2 
 1 M  p2,upper    p0 p0  1  M1   p1
p  2   p 0   2 
 /( 1)
  1 2 
 p2,upper  p0 1  2 M 1  [1  (0.2)1.80 2 ]3.5
p2,upper    p1  p1 p2,upper  100kPa  85.8 kPa
 p0  p1   1 2 
 /( 1) [1  (0.2)1.90 2 ]3.5

1  M 2 , upper 
2
p2,upper = 85.8 kPa
(b) Lower surface—oblique shock

For M1 = 1.80 and l = 9°, we obtain l, from

2 cot   ( M 12 sin 2    1)
tan    => l = 42.8o
M 12 (  cos 2  )  2

=> M1n,lower = M1sin l = 1.80 x sin(42.8) = 1.223

p2,lower 2  1
 M 12n ,lower 
p1  1  1

p2,lower 2(1.4) 1.4  1


 (1.2232   1.58
p1 1.4  1 1.4  1

p2,lower = 1.58(p1) = 1.58(100 kPa) = 158 kPa

S-ar putea să vă placă și